| Or OF MIC THE UNIVERS, MICHIGAN THE 1811 LIBRARIES FLIGHT TEST MANUAL VOLUME 1 Performance VOLUME II Stability and Control VOLUME III Instrumentation Catalog VOLUME IV Instrumentation Systems GENERAL EDITOR COURTLAND D. PERKINS Professor and Chairman Aeronautical Engineering Department, Princeton University ASSOCIATE EDITOR DANIEL O. DOMMASCH, 1953-56 ENOCH J. DURBIN, 1956- Aeronautical Engineering Department, Princeton University Mr P Hej for and on behalf of AGARD Advisory Group for Aeronautical Research and Development, North Atlantic Treaty Organization. 1959 PERGAMON PRESS NEW YORK LONDON PARIS LOS ANGELES LIMITEDSITV OC MICUICAN LIDDADIES 19 PERGAMON PRESS, INC. 122 East 55th Street, New York 22, N. Y. P.O. Box 47715, Los Angeles, California PERGAMON PRESS LTD. 4 and 5 Fitzroy Square, London, W.1 Engin. Long TL 671.7 .N57 1959 V. PERGAMON PRESS, S.A.R.L 24 Rue des Ecoles, Paris Ve Second revised edition 1959 Copyright 1959 ADVISORY GROUP FOR AERONAUTICAL RESEARCH AND DEVELOPMENT NORTH ATLANTIC TREATY ORGANIZATION This volume was prepared under U.S. Air Force Contract 18(600)-1323, administered by the Air Force Office of Scientific Research of the Air Research and Development Command. 1 Library of Congress Card No. 59-13097 Lithographed in the United States by Edwards Bros., Ann Arbor, Mich. 87-189163 THE AGARD FLIGHT TEST PANEL (May 1959) Present Members Mr. Alec F. ATKIN (United Kingdom) Mr. R. P. DICKINSON (United Kingdom) Mr. M. N. GOUGH (United States) Lt. Col. J. J. BERKOW, USAF (United States) Ten. Col. Dott. Ing. Enzo BIANCHI (Italy) Prof. F. HAUS (Belgium) F/Lt. H. E. BJORNESTAD (Canada) Mr. Jean IDRAC (France) Mr. Yavuz KANSU (Turkey) Prof. Dr. phil. H. BLENK (Germany) Prof. Dr. G. BOCK (Germany) Lt. Col. P. N. BRANDT-MOELLER (Denmark) Mr. P. LECOMTE (France) Cdr. W. H. LIVINGSTON, USN (United States) Col. de BUEGER (Belgium) Dr. Anton J. MARX (Holland) Ing. G. CIAMPOLINI (Italy) Mr. Tor MIDTBO (Norway) Col. D. CHRISSAITIS (Greece) Mr. A. D. WOOD (Canada) Panel Executive: Lt. Col. J. A. WOIDA, USAF (June 1958 to present) Past Members Major H. AASS (Norway) Major H. NESSET (Norway) Major W. RICKERT, USAF (United States) Ing. G. CERZA (Italy) Col. Dr. Ing. F. COLUMBA (Italy) Lt. Col. J. L. RIDLEY, USAF (United States) Mr. Handel DAVIES (United Kingdom) Cdr. L. M. SATTERFIELD, USN (United States) Ing. en Chef B. DAVY (France) Cdr. R. J. SELMER, USN (United States) Lt. Col. G. B. DOYLE, USMC (United States) Mr. T. E. STEPHENSON (Canada) Mr. R. R. DUDDY (United Kingdom) Brig. Gen. M. STRATIGAKIS (Greece) Major E. TUSTER (Norway) Ing. en Chef J. FOCH (France) Mr. N. E. G. HILL (United Kingdom) Brig. Gen. Fuat ULUG (Turkey) Mr. P. A. HUFTON (United Kingdom) Major H. UNSAL (Turkey) Panel Executives: Col. J. J. DRISCOLL, USAF (May 1952 March 1954) Cdr. Emil P. SCHULD, USN (September 1954 - 1958) Acting Panel Executive: Mr. R. A. WILLAUME (France) (April 1954 - August 1954) RECORD OF REVISION (This sheet is prepared for your convenience to keep a record of number and date of revisions.) REVISION NUMBER: DATE: ENTERED BY: Science and technology have a big part to play in our efforts to improve the economic and military strength of NATO. The work of the Advisory Group for Aeronautical Research and Development is therefore of great importance. It is always difficult to achieve a harmony of view between technical experts of many nations. The Flight Test Manual is a sigaal example of how these difficulties can be overcome, and a happy augury for the future. Isman THE RIGHT HONORABLE LORD ISMAY Vice-Chairman, North Atlantic Council Secretary General, North Atlantic Treaty Organization Air Force-USAPE, Voba, Ger-15-7 FOREWORD In this first volume of the NATO Flight Test Manual, prepared under the auspices of the Advisory Group for Aeronautical Research and Development, I am pleased to note the cooperative etfort of the best scientific talent of NATO nations coacemed with the problems of flight testing. I am strongly convinced that a cooperative effort of this na- ture will accomplish two general purposes: it will give opportunity to individuals, in a country where scientific and technical accomplishments have been attained, for mutual exchange of ideas with their counterparts in other countries; and it will fumish inval. uable guidance to flying personnel and engineers of NATO countries as yet without real opportunity to develop their own theories and practices. a I sincerely trust that this venture in NATO technical cooperation will set an encour aging example for similar efforts in other domains of aeronautical activity. Famii hortet taco LAURIS NORSTAD, General, United States Air Force Air Deputy, Supreme Headquarters, Allied Powers, Europe NATO NORTH ATLANTIC TREATY ORGANIZATION (Organisation du Traite de l'Atlantique Nord) FLIGHT TESTING VOLUME I PERFORMANCE Edited by Daniel O. Dommasch Princeton University PREFACE TO THE SECOND EDITION The second edition of the AGARD Flight Test Manual is being brought out under new "ground rules" which will undoubtedly go a long way towards improving the develop- ment and distribution of new material in the years to come. This new edition published by Pergamon Press will be available for public sale and will have the great advantage of a well-organized procedure for getting out new material at regularly scheduled intervals and an efficient system for ensuring that the users of the manual can receive this new material when it is available. This new edition contains all of the additions and correc- tions introduced into the first edition from time to time, as well as nearly seven hundred pages of new material, bringing it up to date as of the fall of 1959. COURTLAND D. PERKINS Princeton University 28 September 1959 PREFACE The Advisory Group for Aeronautical Research and Development of NATO found there was a great need for a flight test manual covering performance, stability and control, and instrumentation of aircraft chat could be used by design, development or research engineers, test pilots, and instrumentation personnel of the participating nations in order to expand their knowledge, improve their methods, and standardize their techniques. Although various member nations of NATO have their own separate publications covering the subjects con- tained herein, AGARD recognized the need for the compilation, revision, and enlargement of this material for the benefit of all NATO nations. The authors generously contributed their time and knowledge in the writing of the various chapters. The high quality of their contributions to this manual will insure its success and will further AGARD's mission in the field of flight testing and instrumentation. AGARD was fortunate to find in the person of Professor Courtland D. Perkios ao editor of high competency in the field, who was willing to devote time and effort to consolidate the individual contri. butions of the authors into an integrated techoical publication. To authors and editors I wish to extend my gratitude and appreciation on behalf of AGARD. The members of the Flight Test Panel, together with AGARD Executives Colonel Jobo J. Driscoll and Mr. Rolland Willaume, and the AGARD Clerical Staff are to be congratulated for their cireless efforts, andl especially wish to express my thanks to Messrs. Jacques Foch Bernard Davy, and Jean Idrac, of France; P.A. Hufton, and N.E.G. Hill, of the United Kingdom; Le. Col. G.B. Doyle, United States Marine Corps; Ms. A.J. Marx of the Netherlands; Li. Col. J.L. Ridley and Major Walter Rickert, of the United States Air Force, all of whom participated in the final review of the text of this manual and supplied valuable suggestions as to the suitability of the material. THEODORE VON KARMAN Chairman Advisory Group for Aeronautical Research and Development Acknowledgements In the preparation of a book of this type, a great deal of work must be carried out by people whose names cannot be prominently displayed in the final published product, but without whom the book could never have been assembled. I should like therefore to express my appreciation to the members of the small editorial group at Princeton who worked very hard to prepare the manuscript under the pressure of a very right time schedule. In partic- ular I would like to express my gratitude to Mrs. Dorothy S. Webster, who typed most of the rough and final copy and acted as editorial assistant; Mr. Robert Westover, who prepared a very excellent set of illustrations and to Mr. Sylvester Hight for his help in translating sev- eral of the French papers, A great deal of credit must go to Professor D.O. Dommasch who acted as volume editor for this project. It was only through a tremendous amount of effort on his part that this first volume of the Flight Test Manual was brought into final form within the scheduled period of time. Finally, I would like to express my appreciation to Dr. Carl L. Fredericks and his staff who prepared the final manuscript copy and cooperated with our editorial group at all times. Courtland D. Perkins VOLUME I, PERFORMANCE TABLE OF CONTENTS Page PREFACE ACKNOWLEDGEMENTS INTRODUCTION TO VOLUME I Chapter 1 AIRSPEED, ALTITUDE AND TEMPERATURE MEASUREMENTS TERMINOLOGY CHAPTER FOREWORD 1:1 1:2 1:3 1:4 1:5 1:6 1:7 1:8 1:9 1:10 1:11 1:12 Introductory Comments The NACA and ICAN Standard Atmospheres Altitude Definitions Speed Definitions Machmeters Speed Measurement Pressure and Temperature Pickups Development of Airspeed Equations Methods of Measurement and Errors Methods of Calibrating the Airspeed System Pressure Measurement The Determination of Speed and Pressure Altitude From The Instrument Readings The Establishment of The Pressure Error Correction From Flight Test Results The Estimation of The Pressure Error Under Other Than Test Conditions Examples of Measured Pressure Errors Temperature Measurement Theoretical Considerations Flight Calibration of Temperature Measuring Systems Concluding Remarks on Temperature Measurement Introduction to Lag Measurements in Pressure Systems Development of Equations Determination of The Lag Constant Design of Systems Concluding Remarks Sample Determinations of AV cal and hp 1:1 1:1 1:3 1:4 1:9 1:10 1:11 1:12 1:16 1:17 1:18 1:19 1:13 1:21 1:14 1:15 1:16 1:17 1:18 1:19 1:20 1:21 1:22 1:23 1:24 1:25 1:24 1:28 1:35 1:35 1:37 1:38 1:39 1:40 1:50 1:54 1:54 1:55 REFERENCES 1:57 TABLE OF CONTENTS (Continued) Page Chapter 2 THRUST AND POWER DETERMINATION TERMINOLOGY 2:1 2:2 2:3 2:4 2:5 2:6 2:7 2:8 2:1 2:2 2:6 2:7 2:8 2:9 2:9 2:9 2:10 2:11 2:12 2:13 2:14 2:15 2:16 2:17 2:18 2:19 2:20 Introductory Comments Power Measurement; Internal Combustion Engines Theory of Reciprocating Engine Power Corrections Corrections Applicable to Power Charts Torquemeter Power Corrections Power Measurement - Compound Engines Jet Thrust Measurement - Introductory Comments Discussion and Region of Application of the Methods of Jet Thrust Measurement The Jet Flow Measurement Method The Climb Performance Method The Measurement of the Useful Thrust of Turbo-Propeller Engines Measurement of Shaft Horsepower The Estimation of Jet Thrust Concluding Remarks on Turbo-Propeller Thrust Determination Ramjet Thrust Measurement Concluding Remarks on Ramjet Thrust Measurement Measurement of Rocket Thrust in Flight General Analysis of Jet Thrust Measurement The Measurement of the Thrust of Jet Engines Thrust Measurements with Afterburning 2:10 2:10 2:13 2:15 2:17 2:17 2:18 2:18 2:21 2:22 2:23 2:28 2:32 REFERENCES 2:35 Chapter 3 A SURVEY OF PERFORMANCE REDUCTION METHODS SUMMARY TERMINOLOGY 3:1 3:2 3:3 3:4 3:5 3:6 Introduction The Problems Methods of Performance Reduction Experimental Methods Analytical Methods Summary of Conclusions REFERENCES 3:1 3:1 3:2 3:2 3:5 3:7 3:8 Chapter 4 PERFORMANCE OF TURBOJET AIRPLANES TERMINOLOGY 4:1 4:2 Introductory Comments Introduction to Level Flight Performance Testing of Turbojet Airplanes 4:1 4:1 TABLE OF CONTENTS (Continued) Page 4:3 4:4 4:5 4:6 4:7 4:8 4:9 4:10 4:11 4:1 4:6 4:13 4:16 4:19 4:23 4:26 4:32 Turbojet Engine Parameters Airplane-Engine Combination Characteristics Reduction of Thrust Required Data to Basic Form Thrust Available Range and Endurance in Level Flight The Cruise Climb The Nominal Best Rate of Climb Airspeed Time to Climb for Jet Aircraft Determination of Maximum Energy Storage Schedule and Time to Climb to a Given Energy Height Proof of Construction Used to Determine Optimum Energy Climb Speed from Acceleration Run Data Descent Performance of Turbojet Aircraft Measurement and Correction of Turbojet Take-off and Landing Characteristics Turning Performance of Turbojet Aircraft Concluding Remarks 4:37 4:12 4:40 4:42 4:13 4:14 4:15 4:16 4:42 4:43 4:46 REFERENCES 4:47 Chapter 5 PERFORMANCE REDUCTION METHODS FOR TURBO-PROPELLER AIRCRAFT SUMMARY FOREWORD 5:1 5:2 5:3 5:4 5:5 5:6 5:7 5:8 5:9 Introduction Differential Reduction Methods The Basic Performance Equations Fuel Flow in Level Flight Equations for Climb Reduction Data Required in Addition to the Basic Performance Quantities Required Accuracy of the Coefficients Possible Simplifications of the Method Reduction of Level Speeds and Fuel Consumptions by Performance Analysis Concluding Remarks 5:1 5:1 5:2 5:7 5:9 5:11 5:13 5:14 5:14 5:16 5:10 REFERENCES 5:16 TABLE I 5:17 TABLE II 5:18 APPENDIX, USE OF LOGARITHMIC DIFFERENTIALS IN PERFORMANCE REDUCTION 5:19 TABLE OF CONTENTS (Continued) Page Chapter 6 DATA REDUCTION AND PERFORMANCE TEST METHODS FOR RECIPROCATING ENGINE AIRCRAFT TERMINOLOGY - 6:1 6:2 6:3 6:4 6:5 6:6 6:7 6:8 6:9 6:10 6:11 6:12 6:13 Introductory Comments Level Flight - Power Required Power Available in Level Flight Determination of V max Range and Endurance Testing (Cruising Flight) Pilot Technique in Level Flight Performance Testing The Instrumentation Required for Level Flight Testing Climb and Descent Testing - Sawtooth Climbs Climbs to Ceiling Pilot Technique for Climb Tests Measurement and Correction of Take-off Characteristics Measurement and Correction of Landing Characteristics Concluding Remarks 6:1 6:1 6:5 6:16 6:17 6:25 6:26 6:26 6:29 6:35 6:37 6:45 6:46 REFERENCES 6:47 Chapter 7 TOTAL ENERGY METHODS FOREWORD Part I OPTIMUM CLIMB THEORY AND TECHNIQUES OF DETERMINING CLIMB SCHEDULES FROM FLIGHT TEST 7:1 7:2 7:2 7:1 7:2 7:3 7:4 7:5 7:6 7:7 7:8 7:9 7:10 Introduction A Simplified Theory End Conditions An Exact Approach Miscellaneous Cases Effect of Operating Conditions Derivation of the Optimum Climb Schedule for the Particular Case Instrumentation and Analysis Standardization of Rate-of-Climb Curves References 7:5 7:5 7:8 7:9 7:9 7:11 7:15 7:15 - APPENDICES 1. Flight Test Technique for Accelerated Levels II. Analysis of Airspeed Indicator Time History from Level 7:16 Acceleration 7:17 TABLE OF CONTENTS (Continued) Page Part II ANALYSIS OF THE OPTIMUM ENERGY CLIMB SCHEDULE 7:19 7:11 7:12 7:13 7:14 7:15 7:16 7:17 7:18 7:19 Introduction Definitions Theoretical Analysis of the Optimum Climb Schedule Geometric Representation Properties of the Graph Discussion of Assumptions Corrections Practical Determination of the Optimum Energy Climb Schedule Conclusions 7:19 7:19 7:20 7:21 7:23 7:24 7:26 7:27 7:31 Part III CORRELATION OF THE ENERGY CLIMB ANALYSES 7:32 7:32 7:20 7:21 Review of Preceding Analyses The Third Method of Expressing the Conditions for Optimum Energy Storage Presentation of the Third Set of Conditions in Terms of V and h Summary 7:22 7:23 7:33 7:35 7:36 APPENDIX A Procedures for Determining the Energy Climb Performance of Turbojet Aircraft 7:38 Chapter 8 TAKE-OFF AND LANDING PERFORMANCE TERMINOLOGY Part I FLIGHT TEST ANALYSIS Summary Take-off Definitions Take-off Measurements Basic Take-off Equations under No Wind Conditions Parameter Theory Take-off Performance Reduction Summary of Equations Landing Definitions Landing Measurements Basic Landing Equations Landing Performance Reduction Summary of Equations Concluding Remarks 8:1 8:2 8:3 8:4 8:5 8:6 8:7 8:8 8:9 8:10 8:11 8:12 8:13 8:1 8:1 ,8:2 8:4 8:7 8:9 8:12 8:15 8:15 8:16 8:17 8:18 8:19 TABLE OF CONTENTS (Continued) 3 Page Part II THEORY 8:14 8:15 8:16 8:17 8:18 8:19 8:20 8:21 8:22 8:23 General Principles The Ground Run The Transition and Climb Piloting Technique in the Airborne Phase Application to Flight Test Procedure Critical Speed During Take-off The Landing Maneuver Estimation of Landing Distance The Choice of the Landing Approach Airspeed Sources of Variation in Total Landing Distance 8:20 8:20 8:22 8:28 8:30 8:30 8:32 8:32 8:36 8:38 REFERENCES 8:39 Chapter 9 SPECIAL TESTS TERMINOLOGY 9:1 9:2 9:3 9:4 9:5 9:6 9:7 9:8 Introductory Comments Introductory Comments on Air Flow Visualization Methods of Flow Investigation Flight Testing Other Techniques Calibration of an Angle of Attack Measuring System Measurement of Sideslip Angle Airbrake Evaluation 9:1 9:1 9:2 9:8 9:10 9:11 9:12 9:14 REFERENCES 9:21 Chapter 10 PERFORMANCE TESTING OF HELICOPTERS SUMMARY TERMINOLOGY Introductory Comments Flight Test Program Ground Resonance Control and Center of Gravity Limits Airspeed Calibration Level Flight Performance Measurement Climb Performance Determination Descent Performance Take-Off and Landing Performance Miscellaneous Tests 10:1 10:2 10:3 10:4 10:5 10:6 10:7 10:8 10:9 10:10 10:1 10:2 10:4 10:4 10:6 10:7 10:12 10:17 10:18 10:23 TABLE OF CONTENTS (Continued) Page 10:11 10:12 10:13 Calculation Methods Blade Stall Limitations Concluding Remarks 10:26 10:30 10:31 REFERENCES 10:32 Chapter 11 THE EFFECT OF THE GROUND ON THE PERFORMANCE OF A HELICOPTER TERMINOLOGY 11:1 11:2 11:3 11:4 11:5 Introductory Comments Theoretical Analysis for a Single Rotor lelicopter Theoretical Analysis for Tandem Rotor Helicopters Flight Tests Comparison of Theory and Experiment 11:1 11:1 11:5 11:6 11:7 REFERENCES 11:9 Chapter 12 THE TRANSITION PERFORMANCE OF A HELICOPTER FOLLOWING A SUDDEN LOSS OF POWER TERMINOLOGY 12:1 12:2 12:3 12:4 Introductory Comments Theoretical Analysis for Single-Engine Helicopters Multi-Engine Helicopters Test Techniques 12:1 12:1 12:6 12:7 REFERENCES 12:10 APPENDICES TO VOLUME I A-1 A-3 A-5 Appendix 1 Part I, ICAN Standard Atmosphere Part II, Properties of the NACA Standard Atmosphere Appendix II Graphical Relations Between Calibrated Airspeed, Mach Number and Pressure Altitude for Subsonic Speeds (NACA Standard Atmosphere for Altimeter and Airspeed Indicator Calibration) Appendix III Graphical Relation Between True Airspeed, Altitude and Calibrated Airspeed Under Standard Conditions (NACA Standard Atmosphere) Appendix IV Mach Number Temperature Relations for Various Recovery Factors Appendix V Relations Between True Speed V, Calibrated Airspeed V Mach Number M, and Pressure Altitude h р A-8 A-9 cal A-10 VOLUME I, PERFORMANCE 3 Contributing Authors Idrac, J. Centre d'Essais en Vol France Bottle, David W. Aeroplane & Armament Experimental Establishment, United Kingdom Cheeseman, I. C. Aeroplane & Armament Experimental Establishment, United Kingdom Davy, Bernard France Langdon, G. F. Aeroplane & Armament Experimental Establishment, United Kingdom Lean, D. Royal Aircraft Establishment United Kingdom Davidson, T. W. Naval Air Test Center United States Navy Douwes Dekker, F. E. National Aeronautical Research Institute The Netherlands LeDuc, Rene France Lightfoot, Ralph B. Sikorsky Aircraft Division United Aircraft Corporation Dommasch, Daniel O. Princeton University Down, H, W. Naval Air Test Center United States Navy Lush, Kenneth J. Air Force Flight Test Center United States Air Force Foch, Jacques Centre d'Essais en Vol France Moakes, John K. Aeroplane & Armament Experimental Establishment, United Kingdom Renaudie, J. F. Centre d'Essais en Vol France Gray, W.E. Royal Aircraft Establishment United Kingdom Gregory, J. D, L. Aeroplane & Armament Experimental Establishment, United Kingdom Schwarzbach, Jerome M. Naval Air Test Center United States Navy Guenod, M. Centre d'Essais en Vol France Shields, Robert T. Aeroplane & Armament Experimental Establishment, United Kingdom Soisson, Jean Centre d'Essais en Vol France Hesse, Walter J. Naval Air Test Center United States Navy Hufton, Phillip A. Royal Aircraft Establishment United Kingdom Utting, Ivan E. Aeroplane & Armament Experimental Establishment, United Kingdom i AGARD FLIGHT TEST MANUAL VOLUME I INTRODUCTION Flight testing of piloted aircraft is conduc- ted with several fundamental purposes in mind: (a) To determine the actual characteris- tics of the machine (as contrasted to the computed or pedicted characteristics); field has lain buried in various manuals, reports and the like, to be dug up only as a last resort. This situation has made it diffi- cult for the engineer just entering the field to find his way and has, in certain instances, produced an air of mystery where none should have existed. Certainly the lack of a compre- hensive work on the subject has frequently led to unnecessary and wasteful duplication of effort. (b) To provide developmental informa- tion; (c) To obtain research information. Thus, in assembling this volume, it is felt that we not only serve the purpose of collec- ting and assimilating the offerings of the several NATO nations, but also serve the purpose of providing as unified as possible a basic reference work in the flight testing field. These three purposes have remained the reasons for flight testing work since the day the Wright brothers made their first flight about a half century ago, and there seems to be little reason that they will change in the future. Indeed, as “the proof of the pud- ding lies in the eating”, so also must the proof of an airplane lie in the flying, for in spite of what developments are yet to come, engineering will always remain in part an empirical science, subject to verification by actual test. Thus, although the aeronautical sciences have undoubtedly enjoyed vast devel- opment during the past fifty years, this development has not eased, but rather intensi- fied the need for flight resting work. In the original plan for preparation of these NATO flight test manuals, the material, which now appears in this first volume and the stability and control work of the second volume, were to have been spread over three volumes; the first dealing with fundamental analyses; the second, with engineering appli- cation; and the third, with the routine conduct of tests. It was later decided that the lack of a unified approach resulting from such a spread was not desirable, and accordingly, all work on performance testing has been combined to form a single volume. Since flight testing is a necessary and an integral part of aeronautical development, the science of flight testing has not stood still and, along with other things aeronautical, has made great strides forward during the past five decades. Unlike other, and perhaps better defined branches of aeronautics, the field of flight testing has not been the subject of numerous anthological texts or handbooks, and the great wealth of information in this Moreover, since the manner in which the routine test procedures are organized in detail (if any flight test program can really be considered a routine matter) depends to a great extent on the type of engineering per- sonnel available at a given test center and, of course, on the equipment available, it was decided to minimize emphasis on such things ii source by the airplane's or pickup's pressure field). as data analysis sheets and detailed plans of approach to any given problem. Care has been taken, however, to point out certain pitfalls into which inexperienced personnel may be lured if they are not careful. The second problem, that of reducing test data to standard conditions, is closely related to the third problem involving the design of the test program since the conditions under which the data are obtained and the selection of these conditions determines the manner of reducing the data. Thus, the second and third problems are related and are generally con- sidered simultaneously. Most of the material, contributed by the various authors, pertained to what might be called acceptance flight testing of aircraft; that is to say, it was more concerned with determining what the flight characteristics of a given airplane actually were, rather than with the breakdown of these characteristics into data useful directly in research and developmental work. It should be apparent, however, that the same procedures useful for acceptance tests also are applicable to research and developmental testing, the major difference being the manner in which the final corrected data are presented. Thus, it has been unnecessary to differentiate here between the ultimate purposes for conducting flight tests. These latter two problems are subject to theoretical analysis using several different techniques. When the number of variables which cannot be controlled during the test is small, test methods based on dimensional analysis prove very useful provided the analy- sis considers all the variables which are significant under all circumstances. 1 Finally, we should like to comment on the major problems encountered in flight testing work, broadly these are: (a) Measurement of the actual values obtained during test; In certain instances when a well-defined relationship exists among two or more of the test variables, the so-called analytic methods of anaysis may be used as a basis for a test procedure, as is the case with recipro- cating engine airplanes. Where other methods are not usable, finite difference techniques are available; however, when these are employed, one must generally schedule tests to obtain data initially under conditions as close as possible to standard to reduce the size of the corrections to a minimum. (b) Determination of what the measured values would have been under some arbitrary set of standard conditions; (c) Design of the test program to provide the desired results for the least cost in time and money within the limitations of avail- able manpower and equipment. When reducing data, it is generally pre- sumed that one or more of the test variables such as pressure altitude, density altitude, temperature altitude, velocity (true, calibra- ted or equivalent) or Mach number are "standard", and the data reduction process then revolves about these standards. has given rise to such terms as density altitude methods, etc., which usage, in some cases, is not too descriptive of the actual reduction process involved. In solving problem (a), we must consider the nature of the instrumentation (its accura- cy, reliability, size, sensitivity, etc.), the nature of the quantity to be measured (whether it exhibits local variations, whether auxiliary conditions affect the instrumentation, and so forth) and the effects of disturbances pro- duced by the airplane or by the measuring equipment on the measured quantity (for example, the errors produced at a static In planning any test program, it is neces- sary to consider the variables of pilot tech- nique and of weather. Obviously, one cannot control atmospheric conditions; however, it is iii this procedure is the lack of knowledge con- cerning the functioning of the power plants during the tests, since no data are available determine if the engines are actually performing according to specification. necessary to consider how such disturbances as air turbulence, temperature inversions and so forth affect the test data and plan accor- dingly. With regard to pilot technique, this variable produces the most pronounced scat- ter of data when the pilot is asked to control a number of quantities simultaneously, such as during landing and take-off maneuvers. Thus, if the test program is planned so that the pilot has to maintain rigorously only one item such as altitude, less error due to pilot technique may be expected than if he must control altitude, airspeed, flight path angle, and so forth, all simultaneously. Consequently, no special attention is given to the topic of side by side testing, but rather, in the first eight chapters of this volume, the three basic problems of flight testing are considered and, in the last chapter (nine), certain special topics are investigated. It has frequently been proposed that one way to eliminate certain variables and to simplify the task of data reduction, when com- parison only of airplane characteristics is concerned, is to fly two airplanes simultane- ously through a given set of maneuvers, and then perform the same set of maneuvers interchanging pilots. The major drawback to Because the flight testing field is not a static one, it is anticipated that new and revised ideas will present themselves in the future and for this reason, a loose leaf binding has been adopted for this book to permit changes to be made readily as neces- sary. Comments and submission of new material to the editor by all readers of this text will be welcomed. iv ) ( VOLUME I, CHAPTER 1 CHAPTER CONTENTS Page TERMINOLOGY CHAPTER FOREWORD 1:1 INTRODUCTORY COMMENTS 1:1 1:2 THE NACA AND ICAN STANDARD ATMOSPHERES 1:1 1:3 ALTITUDE DEFINITIONS 1:3 1:4 SPEED DEFINITIONS 1:4 1:5 MACHMETERS 1:9 1:6 SPEED MEASUREMENT 1:10 1:7 PRESSURE AND TEMPERATURE PICKUPS 1:11 1:8 1:12 DEVELOPMENT OF AIRSPEED EQUATIONS METHODS OF MEASUREMENT AND ERRORS 1:9 1:16 1:10 METHODS OF CALIBRATING THE AIRSPEED SYSTEM 1:17 1:11 PRESSURE MEASUREMENT 1:18 1:12 THE DETERMINATION OF SPEED AND PRESSURE ALTITUDE FROM THE INSTRUMENT READINGS 1:19 1:13 THE ESTABLISHMENT OF THE PRESSURE ERROR CORRECTION FROM FLIGHT TEST RESULTS 1:21 1:14 THE ESTIMATION OF THE PRESSURE ERROR UNDER OTHER THAN TEST CONDITIONS 1:24 1:15 EXAMPLES OF MEASURED PRESSURE ERRORS 1:28 1:16 TEMPERATURE MEASUREMENT 1:35 1:17 THEORETICAL CONSIDERATIONS 1:35 1:18 FLIGHT CALIBRATION OF TEMPERATURE MEASURING SYSTEMS 1:37 1:19 CONCLUDING REMARKS ON TEMPERATURE MEASUREMENT 1:38 1:20 INTRODUCTION TO LAG MEASUREMENTS IN PRESSURE SYSTEMS 1:39 1:21 DEVELOPMENT OF EQUATIONS 1:40 1:22 DETERMINATION OF THE LAG CONSTANT 1:50 1:23 DESIGN OF SYSTEMS 1:54 1:24 CONCLUDING REMARKS 1:54 SAMPLE DETERMINATIONS OF AVcal and Ahp 1:25 1:55 REFERENCES 1:57 TERMINOLOGY с Speed of Sound in Free Stream co Speed of Sound at Sea Level Ср Specific Heat at Constant Pressure (perfect gas) CP Pressure Error Coefficient (= - AP/¿PV?) - Су Specific Heat at Constant Volume (perfect gas) f( ) Function of h Tapeline Altitude ho Density Altitude hp Pressure Altitude ht Temperature Altitude Ahp Altimeter Pressure Error Correction (= hp-hip) J Mechanical Equivalent Heat L Tube Length, Ft. M Mach Number ( = V/c) n Normal Load Factor or Exponent for Polytropic Process P Pressure P; Impact Pressure (equal to total pressure for M< 1) Ps Static Pressure PI Total Pressure, Lbs./Sq.Ft. APS Static Pressure Error Correction (Ps -p's) AP Pitot Pressure Error Correction (Pt-P't) Q Instrument Volume, Cu, Ft. a Incompressible Dynamic Pressure (= {P V2) 9c P1-Ps = "Compressible" Dynamic Pressure "' TERMINOLOGY (Continued) ДЧc Dynamic Pressure Error Correction (9c-9'c) R Perfect Gas Constant r Tube Inside Radius, Ft. s Entropy S Wing Area T Absolute Temperature t Time, Sec. Tt Total Temperature V True Airspeed Vcol Calibrated Airspeed Ve Equivalent Airspeed (= Vo) Vi Indicated Airspeed AVcal A.S.I. Pressure Error Correction ( = Vcal-V'cal) W Aircraft Weight A Increment 8 Relative Pressure (= Ps/Pso) Y Ratio of Specific Heats (Cp/cv = 1.4 for air) ) X Lag Constant, (sometimes called time constant or characteristic time, corre- sponds to RC in a simple RC electrical circuit), Sec. Viscosity of Air, Lb./Sec. per Sq. Ft. T W Pi Relative Density ( = P/ Po) Acoustic Lag, Sec. $ Temperature Ratio ( = T/To) P Air Density, Slugs/Cu, Ft. TERMINOLOGY (Continued) Superscripts and Subscripts Observed or indicated conditions are signified by the use of a prime (1) superscript whereas the true or free stream conditions carry no superscripts. The subscript (o) rep- resents standard sea level conditions and the subscript (1) represents conditions aft of a shock wave. (t) Denotes at a given time t (t + o) Denotes at a given time t to (9) First derivative of a quantity with respect to time. FOREWORD This first chapter of Volume 1 of the AGARD Flight Test Manual is devoted to the im- portant topic of the speed, pressure and temperature measurements required in flight testing work. The chapter is divided into five basic parts: (1) Introduction, Standard Atmosphere, Altitude and Airspeed Definitions (2) Speed Measurement (3) Pressure Measurement (4) Temperature Measurement (5) The Effects of Lag Error The various portions of the chapter have been contributed by several nations, and slight differences in the manner of approach to the problems by each group are illustrated in dif- ferent sections. No attempt has been made to reduce the presentations to a single approach; however, to avoid confusion, the terminology has been standardized as far as possible. A certain amount of overlapping of the work of the various contributors exists; however, the retention of this overlap is felt to be a desirable feature of the chapter since it adds to the clarity of the individual sections. Detailed instructions on the mechanics of conducting calibration checks are not included in this chapter. However, it should be noted that before checking any system in the air, a careful prior ground check should be conducted to determine if system leakage exists, whether the instruments are functioning properly, and so on. Moreover, as is the case with any type of flight test work, careful planning and proper test scheduling can save much time, money and effort during the actual conducting of the tests. Items such as these are the responsibility of the test engineering staff, and the decision as to what comprises the proper method of approach in a given case depends primarily on the equipment available and the accuracy required. Finally, turbulent air conditions should be avoided during calibration checks of pressure sensing systems, since there is no way to correct for the effect of turbulence. 1:1 INTRODUCTORY COMMENTS In this first chapter we shall consider the theory and the procedures required for determination of speeds, pressures and tem- peratures as well as the definition of the two fundamental standard atmospheres. 1:2 THE NACA AND ICAN STANDARD ATMOSPHERES The performance characteristics (and in certain cases, the handling qualities) of an airplane depend on the nature of the atmos- phere through which it flies. Because the nature of the actual atmosphere changes from day to day, it is not possible to define airplane performance precisely without first exactly defining the state of the atmosphere at the time we expect to realize the specified performance. This thought leads quite nat- urally to the idea that to standardize per- formance information we must work with some type of standardized atmosphere where- in there exist given values of temperature, density and pressure at each altitude. Both the NACA and ICAN standard atmos- pheres approximate the standard conditions existing at latitude 40° N and are identical up to a standard altitude of 35,332 feet. Above this altitude, the temperature is as- sumed to be a constant -67°F (-55° C) in the NACA atmosphere. On the other hand, in the ICAN atmosphere it is assumed that the temperature does not stabilize until an alti- tude of 36,088 feet is reached, at which point the temperature is taken to be -69.7°F (-56.5° C). Several standard atmospheres have been defined, and two are in use by the NATO nations. These are the NACA (National Ad- visory Committee for Aeronautics) standard atmosphere (used in the United States) and the ICAN (International Commission of Air Navigation) standard atmosphere used by most of the European nations. In both standard atmospheres the sea level conditions are taken to be Po 0.002378 slugs/cu.ft. Po 2116 lbs. per sq. ft. = 760 mm. Hg = 29.921 in. Hg. T. = 59°F : 518.4°R = 288° K = 15°C Because standard atmospheres are, of necessity, somewhat arbitrary statistical averages, the conditions in the actual atmos- phere seldom, if ever, agree with those of a “standard" atmosphere. This means that measurement of such items as ambient tem- perature or pressure at a point in the actual atmosphere cannot by themselves define (for instance) the altitude of the airplane. Indeed, as we shall see, only in the standard atmos- phere are we able to define a “standard altitude." 80 32.17 ft./sec2 In both, it is assumed that the equation of state or perfect gas law holds р р RT • 1:1 The major part of this first volume of the NATO Flight Test Manual is devoted to methods of determining from flight test in the actual atmosphere, what the performance would be at standard altitudes in the stand- ard atmosphere. For this purpose, it is first necessary to determine the ambient conditions in the actual atmosphere as well as such items as true airspeed and Mach number of the airplane relative to the air. where T is in absolute temperature units. With p in lbs. per sq. ft., p in slugs per cu. ft. and T in Rankine units, we have R 1718 ft.? /sec20R. - Again, in both atmospheres it is assumed that the temperature lapse rate is 3.57°F per 1000 feet of standard altitude. This 1:1 provides the equation In the ICAN atmosphere we have T(°F) = 59° -(3.57°h/1000) 1:2 р 8 : Po : e-d, 1:11 Both standard atmospheres are assumed to be dry (relative humidity zero) and non- turbulent. Therefore, the equation of balance is Р : 1.33e- Po 1:12 o To : 0.7519 1:13 dp = -pg dh : 1:3 where From Eqs. 1:1, 1:2, and 1:3 we find that up to the isothermal region the following equations hold for either standard atmos- phere - di : 1.498 + h-36088 20790 1:14 6.256 sen - (17)*** р Po T To (0) 5.256 (1-0.00000689h)5.256 1:4 Eqs. 1:4 through 1:14 completely define the ICAN and NACA standard atmospheres. Tab- ulated data for these atmospheres are given in the appendices to this volume. 4.256 - (5.) . (0)*.286 =(1-0.00000689h) *.266 1:5 2 T = 1-0.00000689h To 1:6 In addition to the two basic standard at- mospheres considered above, certain varia- tions of these are also used on occasion. The variations simulate either hot or cold days and in general presume the same pres- sure variation with altitude as exists in the basic standard atmosphere, but with different temperature and density variations. The location of the tropopause (or start of the isothermal region) is, of course, different from the standard locations when the varia- tions are considered. In the NACA atmosphere at altitudes in excess of 35,332 feet, we have р -0-1 1:7 1.32e-1 1:8 Although it is well known that the iso- thermal or stratospheric region of the actual atmosphere is of limited extent, no account of this is taken in the definitions of the ICAN or NACA atmospheres, and it is assumed that isothermal conditions extend indefinitely. A 3 Foro : 0.7508 Actually the assumption of constant tem- perature is satisfactory up to altitudes some- what greater than 20 miles. Since present day piloted aircraft do not attain altitudes as great as these, the existing standard atmos- phere definitions are adequate as of this time and for the purposes of this manual. 1:9 where d=1.452 + n-35332 20950 1:10 1:2 1:3 ALTITUDE DEFINITIONS At any given altitude in the standard at- mosphere there exist fixed values of tem- perature, pressure, and density, which are the standard values for that altitude. In the actual atmosphere, however, we cannot ex- pect to find (except by coincidence) that any of the quantities (temperature, pressure, or density) at a given distance above sea level will have values equal to those of the stand- ard atmosphere at the same height. 1:12 for h, the resulting relations would de- fine an altitude which is actually a measure of ambient density. This altitude is called density altitude, symbolized by hd or Hd. Because the power required characteristics of a given airframe at speeds lower than drag divergence are a function only of density (insofar as the atmospheric conditions are concerned) density altitude is frequently used as a base reference when the perform- ance of the propeller-driven airplanes is being considered. (c) Temperature Altitude, he Therefore, since many performance char- acteristics depend on ambient conditions rather than on actual or tape line altitude, we frequently find that we are not concerned with true height, but rather with such things as pressure altitude, density altitude and temperature altitude. Before defining these quantities, we note that tape line altitude is significant in evaluating climb performance and obstacle take-off and landing perform- ance. The altitude in the standard atmosphere corresponding to a given ambient tempera- ture is called temperature altitude, sym- bolized by hy or Ht. In common with pres- sure and density altitude, temperature alti- tude is not a measure of actual altitude, se, but rather is a measure merely of am- bient temperature, (a) Pressure Altitude (d) Standard Altitude The altitude read on the dial of an alti- meter set to 29.92 inches of mercury is called pressure altitude. In the United States, altimeters are calibrated according to Eqs. 1:4 and 1:7. Therefore, if we solve these equations for h, the resulting relations de- fine the pressure altitude in terms of ambient pressure. Pressure altitude, symbolized by hp or Hp, is actually a measure only of am- bient pressure and clearly is not a precise measure of the actual distance above sea level. Any altitude in the standard atmosphere is, by definition, a standard altitude; i.e., the pressure, density and temperature all have their proper values simultaneously. In the actual atmosphere we would be at a stand- ard altitude, if for a given pressure altitude, we had the standard atmosphere value of temperature altitude (and of course, by virtue of the equation of state under these circum- stances, we would also have a density alti- tude equal to the pressure altitude). In the NATO European countries, alti- meters are calibrated according to Eqs. 1:4 and 1:11; however, the significance of pres- sure altitude is precisely the same as in the United States. This does not mean that we would know how far we actually were above sea level. Rather it simply means that we could find some altitude in the standard atmosphere where the conditions were the same as those existing about us in the actual atmosphere. As we have noted previously, it would be the result of pure coincidence if we were actually to locate a standard altitude point in the actual atmosphere. (b) Density Altitude, hd If we were to solve Eqs. 1:5 and 1:8 or 1:3 difficult to measure air density while aboard an airplane, and accordingly, it is general practice to measure only temperature and pressure. The density, if required, is then computed using the equation of state or suit- able charts prepared for this purpose, Because we cannot generally hope to con- duct tests at "standard altitudes" even though we must finally present our test data in terms of "standard altitude", we must devise ways and means to correct data ob- tained in the actual atmosphere to standard altitude conditions. For instance, if we con- ducted our tests at a given pressure altitude, in general, both temperature and density would be non-standard, and if we wished to consider that the pressure altitude was the standard pressure altitude, we would have to make data corrections for the non-standard density and temperature. 1:4 SPEED DEFINITIONS Similarly, if we presumed that we could consider the density altitude as standard, corrections would have to be imposed for non-standard temperature and pressure. Just as the definitions of altitude given above are associated with standard atmos- pheric conditions, so also are the definitions of measured airspeed. The conventional airspeed indicators in use by all the NATO nations are actually differential pressure gages calibrated ac- cording to the law of frictionless adiabatic flow, presuming that the ambient conditions are those of the standard atmosphere at sea level. For subsonic flow the calibration equation is derived from the isentropic flow relations and for supersonic flow the calibra- tion equation used assumes the existence of adiabatic flow with a normal shock located just forward of the total pressure pickup. For subsonic speeds, therefore, > A data reduction procedure is frequently identified by the atmospheric quantity held constant during correction. Thus, the first case of the preceding paragraph represents a “pressure altitude" method of data re- duction, whereas the second represents a “density altitude" procedure. Obviously, we also have a third basic procedure and this is the "temperature altitude" method. v2 / 20 veoma o 1) 516-1) [ +- Dp-Ds Ds (e) True, Tape Line or Geometric Alti- tude 1:15 and for supersonic speeds - Dri-Ds Ps (y+l) v27 4114-1) YIY 2c2 The actual distance of the airplane above sea level is called either true, tape line or geometric altitude. Normally, true altitude can only be determined by radar equipment. It cannot be established by instantaneous reading of the pilot's flight instruments, although it can be computed with some pre- cision on the basis of data recorded at various altitudes by a climbing airplane. This latter topic is discussed in Chapters 4 and 7 of this volume in connection with en- ergy climb computations. [**] [v- ) Y+1 1-y+2 1 ] *zrya)"17-1 1:16 . Pol Because the equation of state relates pressure, density and temperature, the meas- urement of any two of these quantities auto- matically determines the third. It is rather when Pr : free stream total pressure : total pressure aft of normal shock Ps : free stream static pressure Y : specific heat ratio free stream sonic speed 8 1:4 or YPSO /Vcal Because the airspeed indicator only senses the quantity Pt-Ps or Pt, -Ps, the indicator dial is calibrated according to the assump- tion that c has its sea level value of co and that Ps, when it appears alone, has the value Pso Therefore, taking Psoas 2116 psf, , y = 1.4 and co : 1117 fps. The calibration equations are . oli ) 1/Vcal + 4 Со ac 2 Со . But .2855 Pt-ps Vcal : 2495 +1 -1 YPSO 2116 ce Po 1:17 2 so that for subsonic flow where Vcal is in feet per second, and Povcal Vcal .2835 Véal 4c = -Cina 2 4 Pt, Ps 2116 1272.5 1:19 cal 1:18 for supersonic flow. Eq. 1:19 is an excellent approximation to 1:17 throughout the Mach number range to which it applies. Therefore, for practical purposes the British and American defini- tions of calibrated airspeed may be con- sidered the same. Inasmuch as the other members of NATO use the same basic cali - bration equations presented herein, we con- clude that except for the units used to mark off the dial, the term calibrated airspeed has the same meaning in all the NATO countries. In Great Britain a simplified version of Eq. 1:17 is employed. The equation is ob- tained from 1:15 as follows: 1 Let ac : Dr-Ds. Then from Eq.1:15, letting c = = co and p = PSO . - ac pt -ps Y-1 Vcal 27 Y/17-1) DSO o {[+ they } ( + 2 Со 1:18A Several other airspeeds are frequently referred to in flight test literature. These are: (1) equivalent airspeed, (2) true air- speed and (3) indicated airspeed. Before Mach numbers in excess of about 0.4 were attained, equivalent airspeed was generally referred to as indicated airspeed. However, since the indicator reading at higher Mach number differs considerably from the equiv- alent airspeed, the two definitions (indicated airspeed and equivalent airspeed) are no longer interchangeable. The relationship be- tween the various airspeeds is given below, V; • Indicated airspeed as shown on the instrument dial, when cor- rected for lag errors, instru- ment error and position error, becomes calibrated airspeed. We next expand the bracketed quantity using the binomial expansion, whence, re- taining three terms only, 2 y Vcal 1 + () 2 Со ac : pso @ly + y (Vcal 8 CO Wca * - 1:5 Vcal - Calibrated airspeed, when cor- rected for the difference be- tween sea level pressure and actual ambient pressure, be. comes equivalent airspeed. Normally it is quite difficult to measure lag errors while in flight, particularly for the case of acceleration runs, Ground checks are generally relied upon to establish sep- arately the lag constants for the Pitot and static systems. Ve : Equivalent airspeed, when cor- rected for the difference be- tween ambient and sea level density, becomes true airspeed. VT : True airspeed At this point we shall consider briefly the relation between the speeds listed above. A detailed treatment is given later in the chapter. Since the lag in a given system depends on the rate-of-change of pressure being applied to the system, and since in flight this rate of change must be established on the basis of indicator readings, it is gen- erally not possible to assume that the over- all lag error correction can be made with a precision of more than 80 to 90%. Good design practice, however, can reduce the system lag errors under any set of con- ditions to a small percentage of the quantity being measured, in which case precise cor- rections are not required for practical work. The topic of lag errors and their determina- tion is considered in detail in sections 1:20 to 1:25 of this chapter. 9 Relation Between V; and Vcal On the basis of a calibration conducted on the ground, the inherent error in a given airspeed indicator or altimeter may be de- termined as a function of the indicator read- ing. This error is called instrument error, and is subject to change due to wear of the instrument and other similar factors. Since instrument error changes with time, frequent calibration is required if precise results are desired. Once corrections for instrument and lag errors have been imposed, position error may be accounted for and suitable corrections made. In this connection we note that under steady level flight conditions, there are no lag errors and when correcting data for cases such as these, position error correc- tions are made immediately following the instrument error corrections. Corrections for instrument error should be imposed prior to any other correction. Following these corrections, the effects of lag should be considered and suitable cor- rections made. In this connection we note that lag errors are present only when the airplane is changing altitude or speed so that the flow about the airplane is of an unsteady nature. For most test work at subsonic speeds it is presumed that all of the position error originates at the static source (static pres- sure pickup). This is a reasonable assump- tion inasmuch as flow perturbations due to the airplane which can affect a properly located total head probe are very nearly isentropic in nature, and accordingly, do not effect the total pressure. Under these circumstances, there is some question as to the exact value of the position error since it is known that the value of the pressure coefficient is affected by unsteady flow conditions. Provided the position error has a reasonably small value, any change due to unsteady flow will be of second order and may be ignored. Of course, if the total head tube is located behind a propeller, in the wing wake, in the boundary layer or in a region where the flow has been influenced by transonic shock for- mation or in a region of localized supersonic flow (such as in the case where the total head tube is located on a wing boom of a 1:6 ) swept-wing airplane where localized super- sonic flow regions may extend past the nose of the boom), there will exist total head position error which must be accounted for. Now calibrated airspeed (Eds. 1:17 or 1:19) is a function only of ac; consequently, at a given pressure altitude, there is a unique relation between calibrated airspeed and Mach number so that at a given pressure altitude Normally, however, it is possible to lo- cate the total head pickup properly and thus avoid these difficulties. It is most desirable to do this inasmuch as such things as lo- calized supersonic flow regions produce rather erratic readings. Ap = f(nw, Vcal). : This is also true under supersonic con- ditions since for Y : 1.4, Eq. 1:16 becomes y = Pt, Ps 166.9 M2 1. 2.5 Ps Position error is determined principally by the value of the pressure coefficient at the location of the static source. As is well known, the pressure coefficient in turn is determined by the value of the airplane lift coefficient and the flight Mach number. For calibration purposes, however, it is neces- sary to determine the relation between posi- tion error and calibrated airspeed. This is accomplished as follows: M2 1:22 We have Comparing this with Eq. 1:18, we see that under supersonic as well as subsonic con- ditions, there exists a unique relation be- tween Vcal and Mat a given pressure altitude. Frequently, weight and load factor effects are neglected when presenting position error data; however, for airplanes carrying large fuel loads and whose weight accordingly may change markedly during the course of a flight, the "W" effects should be taken into account. Ср q 4 = m) 뿡 ​f(CL,M) f M]ofrow,p, m) nW M Үрм? S 2 1:20 where Ap is the position error in lbs. per sq. ft. and n is load factor acting on the air- plane. Since q = y pM²/2, we see that Ap is itself a function only of nW, p and M. Under the assumption that the position error is produced only by pressure coeffi - cient variation at the static source, it is possible to relate altimeter position error directly to airspeed indicator position error (since in most installations both utilize the same static source). This thought is the basis of the altimeter depression method used to calibrate airspeed systems by meas- uring the altimeter error as a function of calibrated airspeed at given pressure alti- tudes. The equation for Mach number under sub- sonic flow conditions is obtained by dividing both sides of Eq. 1:15 by c?, whence 2 P-Ps /y M? 1 0 +,7-10,- Y-1 Ps 1:21 Provided we are dealing with position errors of small magnitude (which is nor- mally the case when the magnitude of these errors is limited by design specifications), it is possible to develop rather simple equa- tions relating altitude, calibrated airspeed and Mach number errors to one another at a and we have that the Mach number is deter- mined by only two quantities, ac and ps. 1:7 Using Eq. 1:23 for Ap, we obtain a direct relation between AVcal and shp, given pressure altitude. The relations be- tween pressure altitude and calibrated air- speed position errors are developed below. Similar relations for Machmeter error are given in section 1:5. 2 2.5 Vcal Anp AVcal : [1+0.2 (copy)" go Со For small errors, we may write the at- mospheric balance equation in the form 1:25 Δρ - - P9 Δhp pg - - 1:23 where the density ratio o has the standard atmosphere value corresponding to the test pressure altitude. where p has its standard atmosphere value for the indicated test pressure altitude. This is the differential form of altimeter calibra- on equations written as a finite difference equation. We may next obtain the differential form of the airspeed calibration equation from Eq. 1:18A. If we presume that the position error arises from static source error only, then the total pressure in 1:18A is a constant and taking the differential of both sides of Eq. 1:18A leads to From Eqs. 1:24 and 1:25 it is seen that the relation between Ap and AVcal is a func- tion only of Vcal whereas the relation be- tween A hp and Vcal depends also on the actual test pressure altitude. It is of course possible to write “exact" equations relating the various position er- rors, and chart solutions of these "exact" equations are available (see Ref. 26, for in- stance). However, greater precision than provided by Eqs. 1:23, 1:24 and 1:25 is not required except when dealing with exces- sively large position errors. The topic of pressure errors is covered in detail by the section of this chapter prepared by Messrs. Shields and Utting and the discussion here serves merely to introduce the topic. Y- 1 dps : PSO fiziki Y-1 (Vcal 2 Со Y-1 Vcal dVcal. co For y : 1.4 this gives Once the increment AVcal has been ob- tained, the indicated speed is corrected ac- cordingly and Vcal is obtained. In practice, position error is determined using the alti- meter depression method, a trailing static bomb, a swiveling static head, the speed course procedure, radar tracking, pacing airplanes and the like. By far the most pop- ular procedure is the altimeter depression method. At the present time, it is generally assumed that no position error exists at su- personic speeds for a properly located static source. 2.5 dps dVcal povcal 1 + 0.2 Ро [1+( or for small finite errors 2.5 Δρ Vcal voodit ve ( [ Po Vcal 1 +0.2 Relation Between Vcal, Ve and V (ve Alcal The equivalent airspeed is of importance for it is a direct measure of the free stream со 1:24 1:8 1:5 MACHMETERS 8 dynamic pressure (q=1/2pV': 1/2 po Ve ) and is frequently used as a basis for reducing flight test data for piston-engined airplanes (see Chapter 6). Equivalent airspeed is de- fined by the relation Ve : Vo In the United States direct reading Mach- meters are available for use as standard panel instruments. These units are similar to existing airspeed indicators with the ex- ception that an additional element is added which senses ambient pressure. This aneroid diaphragm is connected to the instrument linkage in such a fashion that the instrument is capable of solving Eqs. 1:21 and 1:22 directly. 1:26 where Ve s equivalent airspeed V = true airspeed o: P/Po where p is the actual am- bient density. Consequently from Eq. 1:15 Because the Machmeter is a somewhat more complex instrument than the airspeed indicator, less reliance is placed on it for test work than on the airspeed indicator. How- ever, the Machmeter is nonetheless a valuable instrument. 22: (* )3=117- - -1)/ + Y - 1:27 Just as the airspeed indicator and alti - meter are subject to position error, so also is the Machmeter. To determine the rela- tion between the static source pressure error and Mach number error, we proceed as follows: and from Eq. 1:17 we have Dr-ps 1)/y Solving Eq. 1:21 for Pr/Ps, we have for ve ps Y: 1.4 +.)'=17-1 x 8 2 Vool Pt (09 -Ps +.)"" -1)/y 3.6 PSO PI . (1+0.2002" [ 1:28 where 8 = Ps/Pso. ps and taking differentials of both sides hold- ing Pt constant, we have ) dps 1.4 MPS . OM 1 +0.2M2 From Eq. 1:28 we see that to determine equivalent airspeed all we need to know are pressure altitude and calibrated airspeed. To determine true speed, Eq. 1:26 is ap- plied to the value of Ve obtained from Eq. 1:28 (provided the ambient pressure and temperature are known). Charts relating Ve to Vcal are given in an appendix to this volume. Other approaches to the problem of determining true airspeed are given in the portion of this chapter prepared by Mr. Davy. or as a finite difference equation Ap 1.4 MP s il AM 1 + 0.2 M2 1:29 1:9 and Using Eq. 1:23 in Eq. 1:29, we obtain the relation between Ahp and AM, voor ca. Ahp YPs P 1.4 MPs M2 AM (1 + 0.2MP)pg 1:35 1:30 where e has the standard atmosphere value for the test pressure altitude. From Eqs.1:29 and 1:24 we have In the following sections of this chapter, calibration of the airplane's speed, pressure and temperature measuring systems is con- sidered from several different viewpoints. It should be noted that as far as airspeed and altimeter systems are concerned, the prime object of calibration tests is to deter- mine position or pressure error. AM ΔVcal AM. Ap . Ap Alcal Polcal .(1 + 0.2M2 Ncal 12/2.3 we [+02 1:6 SPEED MEASUREMENT 1.4 MPS Со 1:31 The purpose of this portion of Chapter 1 is to explore the currently used methods of determining airplane speed. Although other procedures will eventually be developed, we shall here consider only those which depend on pressure measurements on board the plane and, in some cases, the measurements of ambient air temperature. Relations between Mach number, Vcal, V and Ve are also readily obtainable. To re- late Vcal to M, we take the ratio of Eqs.1:17 and 1:21, whence Pt-Ps -1)/y In this presentation the following defini- tions apply: +1 .)%/y cal pso - có (a) True Airspeed, V M2 (Po-Ps Jy-1)/y ) Ps +1 1.)="12- 1:32 The speed of the airplane's center of gravity with respect to still air is called true airspeed or simply true speed. The true speed is also known as the relative velocity* or aerodynamic speed. We note that airplane ground speed is the vector sum of the wind velocity and the true airspeed. Similarly for Ve and V we have 9 = { pove? - ¿pv? : { xpsMe : 1:33 (b) Calibrated Airspeed, Vcal from which The true speed which an airplane would ve? ур - CO M2 Ро * Editor's note: The reverse of this velocity is referred to as the relative wind, 1:34 1:10 1:7 PRESSURE AND TEMPERATURE PICKUPS have at sea level under standard conditions (hp : 29.92 in.Hg, T = 59°F) for the same difference between total and static pressures (P7-Ps) as exists for a given flight is called calibrated airspeed. In section 1:8 we shall see that to deter- mine speed it is necessary to measure the total pressure Pt, the ambient static pres- sure Ps, and also an air temperature indica- tive of ambient conditions. Calibrated airspeed is of interest for the following reasons: (a) Static Pressure (1) It is measured aboard the air- plane by a simple,precise instrument amount- ing to a differential pressure gage (assum- ing that the correct total and static pres- sures exist at the pickups). Measurement of any other quantity, such as V, Ve (equiv- alent airspeed) or M requires a more com- plex system. Static or ambient pressure (as would be measured by a balloon, for example) is dif- ficult to determine with precision aboard an airplane, and the pickup or static heads which measure it must be calibrated in flight as discussed in section 1:9. (b) Total Pressure, Pt (2) Calibrated airspeed is equal to true speed at sea level under standard con- ditions. (3) It provides a positive measure of take-off and landing speeds regardless of the altitude of the ground or the local barometric pressure (provided the weight is known). Total pressure exists at the end of the total head or Pitot tube, i.e., where the rel- ative velocity has been reduced to zero and stagnation conditions prevail. It is supposed that frictionless flow of the air exists up- stream of the stagnation point, and therefore the stagnation or total pressure represents the maximum pressure encountered at any point on the airplane. This supposition may be used to verify the accuracy of the total head probe since, if two different Pitot tubes give the same pressure for all cases of stabilized flight (in which case, there is no lag), we may be quite sure that both are correct. (4) It provides the pilot with a meas- 4 ure of the dynamic pressure, qc, on which, above all, depend the high speed structural problems. (c) Mach Number The ratio of true speed to the speed of sound in the ambient air (V/c) is called the Mach Number. In meters per second, c: 20.05VT where T is the absolute (Kelvin) с : ambient temperature measured in degrees Centigrade. In feet per second with T in degrees Rankine, c : 49.1VT. If possible, the total head tube should be directed into the airstream; however, it is known from experience that even consider- able angles of yaw do not appreciably affect the measurement of Pitot pressure. On the other hand, care must be taken to insure that the upstream flow is undisturbed and that the total head pickup is not located aft of a shock wave*, in the slipstream or in any At this point we merely mention the equiv- alent airspeed Ve, which is a measure of the incompressible dynamic pressure (9 : 1/2 Pole?). Ve is related to true speed by the relationship Ve : vo. It is not pos- sible to measure Ve directly. *At supersonic speeds a shock always forms ahead of the total head tube; we shall see in section 1:8 how this fact is accounted for. 1:11 region where the total pressure is modified by boundary layer effects, etc. Writing this equation between the free stream and a stagnation point (c) Total Temperature, To v2/2 + JCpT Јерте or va 2Jcp (Ty-T). . 1:37 Total temperature is related to static temperature, T, and results from adiabatic compression of the air to stagnation or total conditions. We shall see later how total and static temperature are related to each other. Static or ambient temperature can be ob- tained from meteorological surveys or meas- ured by the airplane. This formula relates the temperature rise* (T-T) to speed and theoretically allows us to determine speed by measurement of two temperatures; however, this cannot easily be done aboard the plane. 1:8 DEVELOPMENT OF ALRSPEED EQUATIONS (a) Equations for V, M and Vcal at Sub- sonic Speeds In this section the classical thermody- namic notation given below is employed. If no normal shock exists between the free stream and the stagnation point, the compression is isentropic and therefore obeys the law p : Gas pressure P = Density 2 constant. T: Absolute temperature 1:38 J: Mechanical equivalent heat Since air may be considered a perfect gas, the perfect gas law Ср Specific heat at constant pres- sure (perfect gas) р RT 1:39 CV : Specific heat at constant vol- ume (perfect gas) applies. On eliminating p between Eq8. 1:38 and 1:39, we have (between two points 0 and 1) Y : Cp/cv : 1.4 for air To (), Mo Y Y-1 Y 8 R = Perfect gas constant = J(cp-cv) : : 1718 f?/sec2 °R It is assumed that the energy equation (which holds for an adiabatic process (re- versible or not) applies; i.e., 1:40 (1) Equations for true airspeed vº/2 + JCPT : constant. We may also write Eq. 1:37 as 1:36 ve 2 Jept (+- : ) * Editor's note: The original paper used the word isentropic, but since total temperature is the same as long as a process is adiabatic, this more general term was substituted. *Since this applies to any adiabatic process, Eq.1:37 holds subsonically as well as su- personically. 1:12 1 and substituting Eq. 1:40 With y : 1.4, we find that DA V*:23 697 [C:)* -] = (1+0.2 M2)3.5 Ps v : Cp т . 1:46 1:41 Similarly from Eq.1:5 This equation gives the true speed in terms of Ps, PP and T. If we know the total tem- perature, we may also write T+ = 1+0.2M2. T 1:47 1 - 8 ?: 2 J Cp To Ti pt: [:-(43)*""] Ps (/ P+ Eq8. 1:46 and 1:47 show that the Mach number can be obtained in terms of a single parameter, which is 1:42 From the foregoing we see that V is a function of the two parameters, T (or T.) and (Ps/pp) either the ratio of total to static pressure (2) Calculation of Mach number or the ratio of total to static temperature. By definition, M : V/c where c is the speed of sound in the free stream (pressure Ps, temperature T, density p), and (3) Equations for calibrated airspeed ca = you (Y: 1.4) By definition, calibrated airspeed is a function of the total and static pressures, P, and Ps, and corresponds to the true speed under sea level standard conditions. since Eq. 1:44 may be written ale 2 RT Dp-ps v2 = - 2.2.1 ( A * +, 97-117-11 ] 2 y RT Y-1 Ps c2 : YRT. 1:43 Ср Letting 9c * PA-Ps and replacing T by To and p by Po, we have the definition of calibrated airspeed In Eq. 1:41 we shall replace J by (Y/(y-1) R, which is its equivalent. Thus, Р Poly-1)/y v2 : 27 RT ac 2yRTO 2.4 57-1 vCal 8 C +.) 3061-10 Y Y-1 Ро 1:44 1:48 or, using Eq. 1:43 It is easy to show that 2 WAY M2 Ps 2y vě = ( •v%). * : 016-1 (s [ 8 yol Po 1:45 1:13 V2 (b) M2 and it may be seen that Ve cannot be deter- mined in terms of a single parameter as = YRT can Vcal (c) Yol jylly-1) De +7 Po ps M2 (b) Practical Considerations in the Sub- sonic Range 2 (d) Y-1 T.I Assuming that the pressures Pr and Ps are known and have been corrected for position and lag errors, we may compute ac : Pt-Ps which gives us Vcal directly (using a table or graph representing Eq.1:48). Using the ratio Pr/Ps, we may similarly obtain M (from a table or graph representing Eq. 1:46). T+ T 1+ 블 ​M2 2 Therefore, we have three independent variables. The entropy is defined as It may be noted that the pressure altitude hp, calibrated airspeed Vcal and M are all determined by the two quantities P, and Ps; consequently, a relationship exists between hp, Vcal and M. A chart graphically present- ing this relationship is given in the appendix to this volume. (e) s = Cy log(Ps/py) + constant. The properties of the fluid across the shock are related by (1) PV : PV, (continuity) (c) Equations for M, V and Vcal at Su- personic speeds (Normal shock in an adia - batic flow of a perfect fluid) (g) Y-1 2y Ps my verloren y vp + PSI Ya lo psi 2y Considering the uniform one-dimensional flow ahead of a normal shock (no subscript) and behind the shock (subscript 1), the in- variant characteristics of the fluid in these two flows are the two known specific heats, Cp and cy, and as usual we assume (conservation of energy) (h) (PV) V + ps = 1, VSV, Psi (م) R: Jlcp - cv) = constant R y : Cp - CV (conservation of momentum). In each of the two flows, the seven quan- tities (speed V, density p, static pressure P, total pressure Pt, static temperature (abs) T, total temperature Ty and Mach number M) are separately related by the four relations Considering (d), relation (g) shows that Ty : Tt, ; consequently, if we know at least three of the upstream characteristics, re- lations (f), (g) and (h) permit determination of all of the characteristics of the down- stream flow, as functions of given properties of the upstream flow. (This is merely a matter of algebra.) (a) Ps 2 وا RT (for a perfect gas) Using the down stream relations, it is 1:14 demonstrated in many text books that the following relations hold across a normal shock. pickup measures "P,.”, and the static source “Ps”. Thus, using the Rayleigh Pitot For- mula, the flight Mach number M may at once be determined. V = ) ° i = v'y-1)M? +2 (-) M? If the ambient air temperature is known, the true airspeed is obtained from V: MY RT PI Ро (y + 1)M? (y -1) M° +2 T may be determined by atmospheric soundings or by using a properly calibrated low lag total temperature pickup and the equation 2 ym2-ly-1) Psi Ps Y+1 T : Ti M2 (with y : 1.4). I + ylly-1) y+l 5 M2 2 Pt, Ps - 2) [ + "]"! Consequently, the true airspeed is 2y .(M2-1) M (which is known as the Rayleigh Pitot Formula) SYRTI M2 5 2ly-1) (M2-1)(1+y) T - r[+ + (x + 1)^ M? Th Tt As far as the airspeed indicator is con- cerned, it should be noted that at supersonic speeds it measures (Pt, -Pr) where Pr differs from the true static pressure by the position error. The indicator does not measure the difference between free stream total pres- sure Pr or Pi and P, as in the subsonic case where the impact pressure and total pressure are identical (since there is no shock sub- sonically). (x-1) M? + 2 MP, - 2 Mº-tr-) Calibrated airspeed is always referred to as an indicator of the difference between Pt and Ps: This correspondence between speed and pressure difference is based on the fact that calibrated airspeed is the same As is well know, the last equation of the above group demonstrates that if a normal shock exists, then the flow upstream must be supersonic and the flow downstream sub- sonic. In practice, if a standard airspeed system is used with a total head pickup lo- cated well ahead of the airplane and approxi- mately along its centerline, and if a static pressure probe, corrected for lag and posi- tion error, provides the correct ambient static pressure, we shall have that the total * Editor's note: Supersonic Machmeters us- ing the Rayleigh formula for calibration : 1.4 are available for direct Mach number indication. with y 1:15 parameters, angle of attack and the Mach number M. We shall disregard the effect of Reynolds number. as true speed under standard sea level con- ditions, assuming a perfect instrument, no position error and no lag. Thus, up to Mach (1), under standard sea level conditions (P1-Ps = 905 mb - 1117 psf), the corre- spondence must be based on the isentropic or subsonic relation between pressure and speed, whereas above P.-Ps : 1117 psf, we . must use the Rayleigh Pitot Formula. Presuming that a suitable location of the static orifice has been determined which minimizes errors as much as possible, we may determine the amount of residual posi- tion error in terms of Q and M. 1:9 METHODS OF MEASUREMENT AND ERRORS Pr Let Ps be the true static pressure and be the pressure sensed by the pickup. We then define the static position error as dps = 3 Pr - Ps. To determine the pressures P, and Ps, pressure pickups or probes are used. We shall now discuss the characteristics of these pickups.* We further define (a) Total Pressure Ap . Pr - Pr and then form the ratio dos Measurement of total pressure presents no particular problems provided the total pressure probe is placed ahead of any shock waves formed by the airplane, itself. This condition is essential, for it is difficult to imagine any way of correcting for the errors which would result were it not fulfilled. The shock wave due to the pickup itself is, of course, considered in the calibration equa- tions discussed in the preceding section. Δρ which we designate as “relative static er- ror." We recognize that (b) Static Pressure dps Δρ : fla,M) and a dimensional analysis establishes that The static pressure pickup may be located on the side of the fuselage or on a suitable probe. Unfortunately, it is seldom possible to find a location where true ambient pres- sure is sensed because the pressure field at all points in the vicinity of the airplane (at subsonic speeds)** is generally a func- tion of speed and altitude, and the secondary dps mon, ) Ap Pr Δρ SPT 1:49 where m = airplane mass g : gravitational acceleration * The topic of lag errors due to changing pressures is discussed in sections 1:20 through 1:25 of this chapter. Here we con- sider only stabilized conditions. ** Editor's insert S - wing area n : load factor. 1 1:16 The foregoing chart may then be used as the basis of correcting for variations in weight and load factor. This relation may be deduced from dps/Ap : f(a,m), since P/P, is directly related to the uncorrected Mach number and for each value of Ap/Pr, mgn/Spr is directly related to CL and hence too. Eq. 1:14 may also be written (c) Total Temperature dps mgn S Δρ Pr mgn'mon S S 10. The total temperature pickup may be placed at almost any location on the airplane but not in the boundary layer. Since total temperature is invariant through a shock, it is not necessary to avoid the formation of airplane shock in the air ahead of this pickup. 1:50 For a given airplane, S is a constant, and a standard mass, mo, may be established for the given set of tests. Variation of weight will then be accounted for in the load factor, n, which under these circumstances repre- sents the angle of attack, Pr, and Ap which defines altitude and speed. Thus we may write A sufficient flow of air must be maintained through the pickup to avoid thermal lag. How- ever, the flow must not be large enough to prevent the existence of essentially stagnant conditions at the pickup. dps F ܝ 42 ܝ 은 ​용​) O. 1:10 METHODS OF CALIBRATING THE AIRSPEED SYSTEM Consequently, when the relationship be- tween P, P, Ps, and n has been determined by means of the calibration procedure of section 1:10, the following chart may be prepared, When speed is determined by measure- ment of the quantities, Pp, Ps and Ty, we must, in theory, provide a calibration for each of these parameters. At this point we shall consider the quantities Pt and Ps. Two different procedures are used to provide pressure calibration. These are: (1) The speed course method (2) The altimeter depression method wherein we determine the static pressure error dps by determining the flight altitude of the airplane and then measure the true static pressure for compariso: with the in- dicated static pressure. The combination of these two methods may be used to determine total pressure error; however, total head errors (other than lag error) are normally insignificant and gen- erally it is assumed that all position error originates on the static side. -16 Fig. 1:1 1:17 (a) The Speed Course Method This method may be applied at low or high altitudes using some means of precisely de- termining the time of flight between two fixed points. The method is subject to numerous . errors, the worst being due to wind effect which alters the ground-airspeed relation- ship. Runs in opposite directions tend to minimize the wind errors. run through the test air mass is first con- ducted to establish the relation between tape line and pressure altitude at the time of the tests, Obviously, the pacing procedure is also useful, particularly at low speeds with the calibrated pacing airplane having a lower stalling speed than the test vehicle. Thus we may determine positive error under various conditions of speed and alti - tude at load factors near one. Test calibra- tion for load factors greater than one is pos- sible by using radar altitude determination during turning flight. We must also have knowledge of the true temperature of the air through which the plane flies (see Eqs, 1:41 and 1:42), since temperature has an important effect. For these reasons it is often simpler to obtain a precise calibration using the altimeter depression procedure. 1:11 PRESSURE MEASUREMENT (b) Calibration of Static System for Po- sition Errors To determine the speed and pressure al- titude at which an aircraft flies, the "free stream" values of dynamic and ambient ' static pressure are required. The pressures registered by the aircraft pitot-static sys- tem will, in general, differ from these be- cause: To determine the true static pressure at the test altitude, different procedures are employed which depend on the test altitude and speed. (a) the pressure field of the aircraft causes the local pressures at the measuring sources to differ from the “free stream" values, there are errors in registration of the local pressure by the source used (nor- mally a pitot-static head). At low altitudes, we employ tower runs wherein the airplane flies past a tower of known height, and its height with respect to the tower is established by photographic or other means. Comparison of the “actual” pressure altitude at which the airplane is flying, determined by adding the pressure altitude at the top of the tower to the height above the observer, with the airplane indi- cated pressure altitude provides the static error. (This procedure is satisfactory as long as the pilot flies by at approximately the level of the observer. Otherwise, serious errors may result.) The runs are made at various speeds to obtain a plot of position error as a function of speed, The resulting errors are called pressure or position errors and the associated cor- rections to be added to the recorded values to obtain "free stream" values are pres- sure error corrections. Thus we have tion : (1) Static pressure error correc- Ps-P's : APS At altitude the same principle is utilized with the test airplane flying past a calibrated "base" airplane. Radar methods may also be employed to determine high speed errors at altitudes, provided a low speed calibration of the airplane is first made and a low speed (2) Pitot pressure error correc- tion - Pt-P'; : AP (3) Dynamic pressure error correc- tion : 9c-9c = APA-APS - 49c: 1:18 (b) The Determination of Equivalent Air speed from the A.S.1. Reading Generally, static pressure error may be due to both (a) and (b) above, but pitot pres- sure error, when it exists, will be due only to (b) except at supersonic speeds when shock losses occur in front of the head, for which theoretical allowances can be made. The calibration equation for instruments in current use in the United Kingdom is V's. a's + +po. Vacol (i + + 2 : + 1:51 The corresponding errors produced in the readings of the Air Speed Indicator (A.S.I.) and altimeter are termed the A.S.I. pressure error and altimeter pressure error, the latter of course arising only from the static pressure error and the former from dynamic pressure error and hence from both static and pitot pressure errors. where Po and Co are the density and speed of sound under standard sea level conditions, Vcol and co being here in ft. per sec. The reading of the instrument in the ab- sence of pressure error is thus given by Thus we have (1) A.S.l. pressure error correc- Vcal - V'cal : AVcal tion : 96 - tro Vol('++ vente 9c = + co 1:52 (2) Altimeter pressure error cor- rection : hp-h'oAnp In general these errors must be estab- lished by a flight calibration made under the appropriate conditions. In some cases, how- ever, it is possible to extrapolate over a wide range of conditions from a calibration at one height (conveniently ground level) and meth- ods of doing this are considered. From these equations, therefore, the A.S.I. pressure error correction, AVcal can be obtained in terms of the dynamic pressure error, A9c and Vcal or Vicol. This relation is plotted in Fig. 1:2 as AV'cal V8.A 9c for various values of Vical. To obtain the equivalent airspeed Ve * from the "free-stream" instrument reading Vcal, the exact relation between dynamic pressure and speed (of which the calibration equation is an approximation) must be con- sidered. 1:12 THE DETERMINATION OF SPEED AND PRESSURE ALTITUDE FROM THE INSTRUMENT READINGS (a) General The relation between equivalent airspeed and pressure follows from Bernoulli's equa- tion for adiabatic compressible flow The instruments used in the aircraft to record speed and height are of course pres- sure gages, though scaled in knots and feet respectively. A unique relation, therefore, exists according to the scale or calibration equation between the applied pressure and the instrument reading. {ve + Ps P constant 뿅 ​* These comments are based mainly on Ref. 1 and use the approach of that report. * The equivalent airspeed Ve is defined by Ve - Vo where V is the true airspeed and o the relative density. 1:19 soots A.SI PRESSURE ERROR CRARECTION 20 Alcal. KNOTS SOOTS o u -70 -30 -10 OI OS 70 20 40 SO GO DYNAMIC PRESSURE ERROR CORRECTION APLB./SQ FT Fig. 1:2 A.S.1, Pressure Error Correction -10 bos soo BASCO ON (+ Alcal c? 12 Apa por con1 spot mai ves ta veri3 [ + 4 (Won t execu] - Firavaillon te -20 cal { 1 (+ Vcal c2 ) 400 25 Vcal = 200KTS. 300 OOI -30 1:20 which gives (under isentropic conditions) where Po is the pressure under standard sea level conditions and the air density is related to the pressure altitude hp according to the scale on which the altimeter is calibrated, For U.K. instruments the ICAN standard atmosphere* values of Po and p are used. YI(Y-1) Pp B (var de har . : + Y-1 2 va 8 COP ." 1:53 From Eq. 1:56 it follows that where 8 is the ratio of the ambient static pressure Ps to the sea level standard pres- sure. This equation relates 9c to Ve and S, and is often expanded as a power series in Ve/8 co? to give hp Aps - Ps - D's 8 s gp dh hp 1:57 2 4c = { Po ve? (i++ Ve Ve? 1. Ve Sco? '40 8?com + :) 1:54 From this equation the altimeter pressure error correction Ahp can be obtained in terms of static pressure error correction Ap's and h'p or hp. This relation is plotted in Fig. 1:4 as A hp vs Ap's for various values of hp. Eqs. 1:52 and 1:53 or 1:54 give the rela- tion between Ve and Vcal , viz., Vcal? Vcol² Vcal?(1 + (1+ t. vcov :) 1:13 THE ESTABLISHMENT OF THE PRESSURE ERROR CORRECTION FROM FLIGHT TEST RESULTS Experimental Methods vo(+ tv I Vez + Ve 40 8200 8 CO 1:55 Various experimental methods are avail- able for the determination of the pressure error correction. They may be divided into two main classes: This relation, which includes the calibra- tion relation, Eq. 1:52, and which through is a function of the altitude, may conveniently be called the calibration-altitude relation. It is given in Fig. 1:3 as a chart of Ve - Vcal against Vcal at a series of pressure altitudes hp. (a) A comparison of applied pitot and static pressures, P't and p's with an experi- mental determination of the “free stream" pressures, (b) A comparison of the indicated cali- brated airspeed Vical with an experimental determination of the equivalent airspeed Ve. (c) Determination of Pressure Altitude from the Altimeter Reading The calibration equation for the altimeter is of the form * Editor's note: The NACA standard atmos- phere is exactly the same as the ICAN at- mosphere up to 35,332 ft. altitude at which point the stratosphere is assumed to start and the temperature is taken as -67°F. In the ICAN atmosphere the temperature of the stratosphere is taken as -56.5°C (-69.7°F). h Ps - Do --S" 9p dh 1:56 1:21 INDICATED CALIBRATED AIRSPEED, VCal KNOTS 6 100 200 300 400 500 600 SEA LEVEL -5 01- Sooo -15 Fig. 1:3 Scale - Altitude Relation , Ve-Vcal ,KNOTS M=1.0 10000 PRESSURE ALTITUDE hp FEET 20 60000 15000 55000 1500 -25 750000 20000 40009 25000 30000 -30 Se 1 + 1) en (over 1 + 1). No casve + ਕਰ ve? 8c2 + ) *(1 + i en helt (----+ 1:22 PRESSURE ALTITUDE INDICATEO SY ALTIMETER hP FEET 20.000 40,000 50,000 60,000 2500 20,000 2000 ALTIMETER PRESSURE ERROR CORRECTION Ahp FEET. 0.000 1500 SCALEVE H000 Fig. 1:4 Altimeter Pressure Error Correction (ICAN Scale) 500 sos 0000t 0.000 210.000 SEA LEVES ooo2 F30 000 -90 -70 -60 -50 -40 -30 -20 -10 -80 STATIC 20 30 40 SO PRESSURE CORRECTION Aps:LB/SQ.FT. SEA LEVEL TORI-1 het hp -1000 BASED ON APS se.dh 20.000 hp -1500 6000, 30000 - 200g -0.000 1:23 Comparison of Speeds In the first method, the pressures may be measured either directly on a pressure scale, or in terms of pressure altitude on the alti- meter. These two methods will now be con- sidered and it will be shown how the pressure error corrections Vcal and A hp may be deduced from the flight test results. In this method the airspeed may be deter- mined by timing the aircraft over a measured speed course, by radar tracking (for both of which methods an allowance for windspeed is required) or by flying in formation with a calibrated aircraft, Comparison of Pressures This is the basis of the standard method at present used in the U.K. for ground level calibrations. The static pressure error is determined by the aneroid method which con- sists of comparing the reading of a sensitive pressure gage connected to the aircraft static source with another gage registering the air pressure at a suitable reference point past which the aircraft flies, The equivalent airspeed Ve deduced from these tests enables Vcal to be found from the scale altitude relation of Eq. 1:55 plotted in Fig. 1:3. The A.S.I. pressure error is then given directly by AV'cal Vcal · V'col To derive the altimeter pressure error correction, the reverse process of section 1:14 is then applied. The pitot pressure error is usually meas- ured by a differential pressure gage connec- ted between the aircraft pitot pressure source and a venturi pitot head mounted at some suit- able position on the aircraft clear of slip- stream and other disturbances. 1:14 THE ESTIMATION OF THE PRESSURE ERROR UNDER OTHER THAN TEST CONDITIONS General Other experimental methods of deter- mining the static pressure error correction by pressure comparison include the compari- son of hip with the pressure altitude meas- ured by flying the aircraft in formation with a calibrated aircraft, by using a trailing static head (Ref.2), or by flying in a part of the atmosphere recently calibrated (Ref. 3). While it is desirable to measure the pres- sure errors under the conditions to be oper- ationally obtained, this will not always be convenient in practice. By making appropri- ate assumptions, estimates of the error at other conditions of height, weight or normal acceleration may sometimes be made from a calibration over the speed range at one height. According to the scale on which the pres- sure errors are measured, Eqs. 1:51 and 1:57 or Figs. 1:2 and 1:4 may then be used to obtain the instrument pressure error corrections. Thus, if the static pressures are compared on the altimeter scale, Aho is given directly from the experimental results. In general, from conditions of flow sim- ilarity, we should expect AP Cpo teve Ave-flCL, M) or a leaf (CL, m) of ,M. The appropriate A.S.1, correction may then be found by using Fig. 1:4 to deduce Ap's from Oh's and hence obtain the dynamic pressure р error Aq'c-Ap', -Ap's (the pitot pressure error being determined independently, e.g., by venturi-pitot). The A.S.I, pressure error correction is then obtained from Fig. 1:2. 1:58 To derive this relation experimentally for direct application to any flight condition 1:24 angle of attack would thus require level flight calibrations at several heights over the full height range. The appropriate assumptions on which pre- dictions to other conditions can be made from tests at one height depend on the Mach num- ber and are considered below for several ranges of this parameter. CLT-M2 = constant and CP/I-M2 = constant. . Low Mach Number Therefore For M < 0.4, the effects of compressi- bility on pressure error are usually neg- ligible, and without introducing serious er- rors, Eq. 1:58 may be simplified to Vel/nW 1 - M2 • constant en geef ICU) or favor Δη Vcal VnW nW and AP : constant. nW From this relation established over the CL range at one height, the instrument pres- sure error corrections at any relevant height, weight and normal acceleration can be ob- tained. These relations can be used to obtain the pressure error at any altitude, weight and normal acceleration in the region where the Prandtl-Glauert rule holds from a calibra- tion over the full speed range at one height. (b) Change of Altitude There is a unique relation between AVcal and Vcal at all heights for a given weight and normal acceleration. However, Ahp vs. Vcal will depend on height and can be ob- tained by use of Fig. 1:4. Medium Subsonic Mach Numbers (a) General For higher speeds, account must be taken of compressibility effects. Allowance may be made for these up to a moderate Mach number (say, M = 0.75) by use of the Prandtl- Glauert relation derived for thin aerofoils at small angles of attack provided no shock waves have appeared. If the relationship between AVcal and Vcal had been established by test at one height h, it is possible to calculate this relation- ship appropriate to another height h2. Thus Ap = constant and ve? W1-M2 = constant at constant angle of attack, weight and normal acceleration, The process is as follows: (1) Select a value of v'cal, and its 1 appropriate AV'cal, giving Vcali Včali + AVcali and obtain Ap from Eq. 1:52.* From this relation we have at constant * Editor's note: As used here, the quantity Ap represents the overall position error of the airspeed system and therefore is taken equal to 9c-a'c. Consequently from Eqs.1:51 and 1:52 Ap=9c-4c = {povéal [i++ #mo{Vēl [i++. Po prvi [* { 1 VI? 112 cal cal + -1} 1:25 (2) Obtain Ve, from Vcal, and Eq. 1:55. (3) Find the value of Ve ? W1-M2 and from this determine the value of Vez at the new altitude. . constant M and Cl, the quantity nW/p is constant and there is a change of height. This may be read from Fig. 1:5 using the constant Mach number lines (dotted) and the ordinate scale in Vcol. Correction back to the original height, if the Prandtl-Glauert relation is assumed, can be made from Fig. 1:5 by following a constant Mach number line to the corrected Vcol and returning along a constant Vcalo line to the original altitude. (4) From Vez and Ap obtain Včala and AVCal2 by reversing the process of (1) and (2). Repetition of this process over the range of Vcal gives the new relationship between Vcal and AVcal at any required height. (d) Applicability of Method Stage 3 involves a successive approxima- tion and to simplify the calculation, the re- lationship between Vcali and Vcale with altitude at constant angles of attack has been plotted in Fig. 1:5 (full lines which give the change in Vcol with hp at a constant pressure error correction Ap). This application of the Prandtl-Glauert relation has been found to give fairly good agreement with experiment for several types of wing tip and nose-boom pitot-static in- stallations up to about M = 0.75 at low CL values (< 0.2), though discrepancies have occurred from about M = 0.5 at high CL's? . . It is thus desirable to check experimen- tally the applicability of the Prandtl-Glauert relation to a particular type of installation by calibrations at high and low altitudes be- fore using it above about M = 0.6. Unless the pressure error is large, this will give sufficient accuracy when used with applied pressures, i.e., to give the variation of Veal with n'p. Note that although Ap is unchanged when converting a Ap - Vcal chart to another altitude, AVcal at a given Vcal will change as it depends on Vcal as well as on Ap. . Transonic Speeds (c) Change of Weight and Normal Ac- celeration Generally in the transonic range the full generalized expression must be considered, i.e., At constant angle of attack and Mach num- ber we have as for the incompressible case Cp = f(CLM). . Ve? nW : constant and Δρ nW = constant. The effects of Mach number in this range are not predictable by any simple theory and it is unlikely that any reliable method can be devised permitting extrapolation to be made from measurements at lower speeds into the mixed flow region just below M : 1. Pressure error calibrations must therefore be made at high altitude up to the maximum attainable Mach number and in general at several heights to permit separations of the effects of CL and M. These relations follow without any as- sumptions about the effects of compres- sibility. Since this correction is made at 1:26 BASED ON v2 = CONSTANT 600 oog 1-M2 THE AND THE SCALE-ALTITUDE RELATION OF FIG. 3 BETWEEN Ve & Vcal 500 Imzodo M=0.85 M0-89 SEA LEVEL. CALIBRATED AIRSPEED, Vcal., KNOTS 500 400 =o75 LLLLY || || IL CALIBRATED AIRSPEED, Vcal , KNOTS ME 0-65 400 M:0.66 300 FM-o 50 M=0.40 300 -2004 M0301 200 v M20.20 -10,000 o +30.000 +50,000 +60poo +19,000 +20000 +40,000 PRESSURE ALTITUDE, hp, FEET Fig. 1:5 Variation of Calibrated Airspeed with Altitude at Constant Attitudes Assuming the Glauert Law 1:27 by the Rayleigh formula. Y/ (Y-1) Mach number effects become increasingly predominant as height is increased and CL effects less important so that a plot of Cp vs. M obtained at one height might be generally applicable (above, say, M = 0.7) over a mod- erate CL range. It is desirable, however, to check the validity of this assumption for each installation by tests at widely different heights. Since Cp : Ap/1/2yPsM? these calibrations can be presented for use with Figs. 1:2 and 1:4 in terms of A p/Ps vs. Mcal. ( ) M2 اه Y+1 2 2Y M2 Y+1 D's Y-I\1/17-1) Y+1 7- or for air with Y: 1.4 - Supersonic Speeds PH pia - 5/2 166.9 M' ( 7M2-1) At supersonic speeds, pressure error problems may be expected to be of a less serious nature than in the transonic range. This equation may thus be used to extend the scale-altitude relation of Fig. 1:3 to su- personic speeds. (a) Static Pressure 1:15 EXAMPLES OF MEASURED PRESSURE ERRORS Provided the static source is ahead of the fuselage bow wave (as for a pitot-static head on a nose boom) the pressure at the source will be unaffected by the aircraft pressure field and any pressure error will be due to failure of the head itself to register local "free stream" pressure. General The calibration can thus be derived from wind tunnel tests on the head alone or from flight measurements with the same head on another aircraft. Available evidence sug- gests that for a static tube with the pressure holes more than 8 10 diameters behind the nose the effects of the nose shock wave have mainly died out so that the pressure error may be very small at supersonic speeds. Some experimental data extracted from various sources are presented in Figs. 1:6, 1:7, 1:8, 1:9 and 1:10 to indicate the effects of several parameters and illustrate the magnitude and variation of pressure error which may be expected. To facilitate com- parisons, all the errors are presented in the form of a pressure coefficient, Cp : Ap/1/2pV?. Additional data on these and other aspects of pressure error problems together with a comprehensive list of earlier references are given in Ref. 4. (b) Pitot Pressure Errors Arising From the Head Itself (a) Pitot Pressure Error Because of the local shock formed ahead of the pitot entry, the pressure registered by the pitot will be less than that given by Eq. 1:53. Assuming a normal shock at the entry the pitot pressure is given in terms of "free stream" Mach number and static pressure Subsonically, when the airılow up to the head is isentropic, there will be no effect of the aircraft pressure field on the local total 1:28 ANGLE OF YAW OR INCIDENCE -DEGREES o 5 10 15 PITOT -0.02 PRESSURE ERROR • 12 pv? M=0.4 = се -0.04 M:07 M 1:06 0.06 -0.08 PITOT PRESSURE ERROR. FIG. 5A. ANGLE OF YAW OR INCIDENCE - DEGREES 5 15 M: 0.6 -0.02 STATIC PRESSURE ERROR 12 pv2 M: 0.7 \ M:04 = cp -0.04 -0.06 -0.08 STATIC PRESSURE ERROR. FIG. 5B. Fig. 1:6 Effects of Inclination to Flow and Mach Number on Pressure Registered by British Mark VIII A. Pitot Static Head 1:29 DISTANCE OF STATIC HOLES FROM NOSE. METHOD OF TEST NOSE DIAMETER. GROUND LAUNCH POCKET. 8-7 8.3 TRANSONIC TUNNEL 15.0 9-7 } } + 0.03 20-0+ from Various Heads at Zero Yaw Fig. 1:7 Comparison of Static Pressure Errors at High Mach Numbers Cp + 0.01 10.0. 0.8 0.9 0-1 1-1 1-2 1-3 1.4 MACH NO 1:30 .12 •08 YA CHORD 04 YZ CHORD Cp 34 CHORD I CORD Distances Ahead of the Aircraft Wing Fig. 1:8 Calibrations of a Static Tube Located at Various THAT 2 CHORD 112 CHORD. --04 -2 192 2x 80.- 9 OD 1-0 o 1-2 À tul 9-1 8-1 w 1:31 8.1 9.1 1.4 12 DIA. 1/2 DIA 2.1 I DIA ୯ 0.1 CL 8. 9. 2 .12 80 VO -04 80.- dy Fig. 1:9 Calibrations of a Static Tube Located 0.5, 1.0 and 1.5 Body Diameters Ahead of the Fuselage Nose 1:32 HEAD POSITION TYPE OF ARCRAFT POSITION OF STATIC SOURCE ALONG SPAN LENGTH OF BOOM A WING LEADING EDGE SWEPT WING .91 s •63c 8 W .715 660 N 1:05 1.470 0 11 DELTA WING 1.05 . · 155 . NOSE BOOM N .950 SWEPT WING 1.810 WHERE S: SEMI-SPAN. CE LOCAL CHORO. O: FUSELAGE DIAMETER +0.15 +0.10 - - - - А +0.05 STATIC PRESSURE ERROR /Pv² B = Cp -0.05 -0-10 -0-15 0.5 0.9 1.0 1.1 0.7 0.8 TRUE MACH NUMBER Fig. 1:10 Some Examples of Static Pressure Errors Measured in Flight at High Mach Numbers 1:33 pressure, and any pitot pressure error will be due to the failure of the head to register the local total pressure, usually because of inclination to the local flow. is normally zero except at high angles of attack. Static pressure error, however, varies considerably with head location and speed, and it is not possible to make gen- eralizations applicable to all installations and flight conditions. Individual calibrations are the only reliable method of obtaining static pressure errors. However, some ex- amples are given to illustrate the effects of some of the parameters. The effect of incidence or yaw and Mach number on pitot reading is illustrated in Fig. 1:6A by wind tunnel data obtained on the RAF Mark VIII pitot-static head (data from Ref. 5). The error is negligible up to about 5° and in practice it is generally found that aircraft pitot pressure errors are zero above about 200 knots IAS. The effects of yaw may be delayed by modifying the nose geometry to give a sharper-edged entry but for this par- ticular head the aim has been to minimize the dynamic pressure variation with yaw, and the pitot pressure error variation with yaw is similar to that for the static pressure error. (b) Effects of Head Location These effects are illustrated in Figs. 1:8 and 1:9, which represent flight data (Ref.6) obtained up to moderate Mach numbers, on straight-wing aircraft, the results being plotted in terms of CP vs. CL. (b) Static Pressure Error (c) Wing-tip Leading Edge Position The effects of inclination to the flow and of Mach number on the pressure registered by the static holes are illustrated by wind tunnel data (from Ref. 5) for the RAF Mark VIII head in Fig. 1:6B. The effects of Mach number shown up to M = 0.7 are not large and other similar evidence indicates a sim- ilar trend for this type of head up to M = 0.9. Fig. 1:8 shows Cp vs. CL for several lengths of boom at the wing-tip (Ref. 6). At low CL's the static pressure error is fairly small for a boom length of 1/2 chord or more and negligibly small at 2 chords. At high CL's the error becomes negative in all cases and tends to a fairly high value. Part of this trend is evidently the angle of attack effect on the head itself, illustrated in Fig. 1:6B. (d) Nose Boom Position The effects of Mach number in the tran- sonic and supersonic ranges are illustrated in Fig. 1:7 for several heads with zero in- clination to the local flow. These indicate that with the static holes located well back down the tube (more than 8 diameters) the error is small at supersonic speeds but that there is a region around M : 1 where local shock waves affect the registered pressure and the error may become moderately large. There are no data at present available in U, K, on the effects of yaw at supersonic speeds. Fig. 1:9 shows similar data for nose booms. The error tends to be large for a boom less than about 1-1/2 fuselage dia- meters in length, the variation with CL show- ing a trend similar to the wing position data. (e) Effects of High Mach Number Pressure Error for Aircraft Installation Examples of measured static pressure error on several installations over the high Mach number range are plotted in Fig. 1:10 as Cp vs. M, all the data being from meas- urements in level flight between 30,000 and 35,000 feet. (a) General As has been stated, pitot pressure error 1:34 1:17 THEORETICAL CONSIDERATIONS Assuming that the flow is adiabatic, the energy relation Few systematic trends can be discerned but the static error tends to a high value near M : l in all these cases, though it appears possible to delay the rise by lengthening the supporting boom. The two nose-boom in- stallations shown have a constant pressure error coefficient over a large Mach number range (one up to M : 0.96) while the wing installations exhibit various trends. It is possible that wing fences are in some cases affecting these. * D V2 y Y-Ip 2 y . = 9cp To : constant applies to the flow past the temperature pick- up provided it is located at a point on the airplane where the energy content of the air is the same as in the free stream. In the above equation The limited evidence here above M : 1 confirm that the error reduces again to a small value as indicated for heads alone in Fig. 1:7. Y р • Cp/cy : ratio of specific heats = ambient pressure - ambient density 1:16 TEMPERATURE MEASUREMENT P 11 Ср specific heat at constant pres- sure T, At the speeds and altitudes at which air- planes are capable of flying it is possible to assume (with insignificant resulting error) that the flow ahead of any temperature pick- up is adiabatic in nature and that the air obeys the perfect gas law. Thus in theory, at least, the measurement of air temperature poses no particular problems. However, problems do arise in the practical measure- ment of temperature, principally because of equipment limitations. g : total temperature (absolute scale) = gravitational acceleration con- stant We shall here consider the problem of measuring temperature, using equipment carried aboard the airplane, and merely mention that radiosonde equipment or similar means may be used also in certain instances, provided the time lag between the measure- ment and the conducting of the actual tests is small. If the speed is low (of the order of 200 feet per second or less) the term [y/67-1)). (p/p) in Eq. 1:59 is predominant, and little error is introduced by ignoring the V2/2 term. Under these circumstances the total and am- bient temperatures may be assumed the same for all practical purposes. At higher speeds the velocity must be considered. A convenient relation for evaluating ve- locity effects is developed below. The per- fect gas law may be written р : SRT P 1:60 * Editor's note: The eccentricities in wing boom data shown may possibly be due to the formation of shocks ahead of these booms even at subsonic free stream Mach numbers. This phenomenon results from the outward extension of localized supersonic flow regions which build up along the fuselage sides under transonic conditions, where g = gravitational constant I . R = Gas constant (53.3 ft/ºr using English engineering units). 1:35 1 Combining Eqs. 1:59 and 1:60, we have the flow past a pickup element which projects into the air stream, ܙܘo(-,) Y y? (RT) + 2 : gop te There is at least one stagnation point on the temperature head where total conditions exist. Moreover, a boundary layer is built up around the sensing element, and as a first approximation, total temperature exists in the air particles comprising the base of the layer. where T : ambient temperature, But by definition : Ср = R. Thus, at first glance one would imagine that the pickup would sense total temperature. Actually, the sensing element as a whole never does achieve stagnation temperature because of heat flow from the thin boundary layar to the passing air and because of heat flow from the pickup itself, to surrounding structure. Thus, most pickups measure a temperature somewhat less than the total, but greater than ambient. Thus, on dividing by gcp and rearranging, -1 T: - To -6. v2. R 1:61 Thus the total temperature differs from the ambient by a constant multiplied by the true airspeed squared. The losses from the pickup are usually accounted for by inserting a correction (or recovery factor) "e" in Eq. 1:62, which then assumes the form If we note that YgRT c? (the ambient speed of sound squared), Eq. 1:61 may be written To - f[it (39)w:] : + EM 1:63. To = - T[i+(") ne] 1:62 where M = Mach number, For most test work it is assumed that e is a constant, independent of temperature, and generally this is a satisfactory assumption. However, it seems apparent that at super- sonic Mach numbers will become quite temperature dependent, Since Mach number can readily be deter- mined when total and ambient pressures are known (see section 1:4 of this chapter), Eq. 1:62 is a most convenient relation between total and ambient temperatures for use in flight test work. Excepting for the recently developed axial flow vortex thermometer (Ref. 7), existing flight test temperature pickups unfortunately measure neither the total nor ambient tem- perature, but rather something in between. To understand the factors which cause this situa- tion, we shall consider briefly the nature of Although the value of € (which varies from about 0.60 to 0.95 for typical installa- tions) may be determined from wind tunnel tests of a given type installation, it is desir- able to determine e from actual flight runs. The procedure used to accomplish this is discussed in section 1:18. Although vortex thermometers are not at present used at major flight test activities, 1:36 1:18 FLIGHT CALIBRATION OF TEM- PERATURE MEASURING SYSTEMS they have certain features worth discussing here. The vortex thermometer is based on the idea that the pressure and density, and consequently, the temperature, decrease to- ward the center of a free vortex (provided, of course, we do not consider the vortex core itself). This phenomenon is called the Ranque effect and is the basis of operation of the Hilsch tube which may be used to provide small amounts of refrigeratej air. The major factor to determine by flight calibration of temperature measuring sys- tems is the element recovery factor. Nor- mally there are no position errors to worry about, although lag errors do exist and may be appreciable during dives or acceleration runs at supersonic Mach numbers. Under subsonic conditions any lag errors which may exist are customarily ignored. Here we shall concern ourselves only with the determination of recovery factors. Referring now to Eq. 1:63, this may be written Toi ۲۰(7) 1 Te M 2 In a vortex thermometer, spin is intro- duced into the air passing the temperature sensing element and an approximation to free vortex motion is obtained. The spin or vor- ticity is introduced by a rotating vane (gen- erally rotated by the passage of air past the vane) designed to produce sufficient circula- tion at various forward speeds to produce a temperature drop equal to the difference between the total and ambient temperatures of the air stream. Accordingly, a properly designed vortex thermometer reads ambient air temperature regardless of forward speed (at least in the subsonic speed region). 1:64 where Tt; : indicated total temperature. At constant ambient air temperature T, Eq. 1:64 describes a linear relationship be- tween Tt; and M’, provided e is constant. Therefore, if a series of flight runs is made at various flight speeds (covering the speed range of the airplane at a given altitude), a plot of Tti vs. M? has the general nature shown in Fig. 1:11. Two types of vortex thermometers have been investigated, the tangential type and the axial flow type. According to Ref. 7, the axial flow type may be designed to give sat- isfactory results up to 500 mph true air- speed; however, little information is avail- able describing functioning at higher air- speeds. Siode 2 The sensing elements in present use for test work are of two types: ( O (a) the shielded resistance-type con- nected to a voltage compensated Wheatstone bridge; (b) the shielded thermocouple-type con- nected to a galvanometer circuit (with am- plification as required). Minimum Test Speed M2 These units are theoretically satisfactory at all speeds provided the recovery factor is known. Fig. 1:11 1:37 As may be seen from the figure, extra- polation of the test data curve to M? : 0 determines the ambient temperature of the air mass in which tests were run. Knowing the ambient temperature, the slope of the curve of Tr; vs. M2 may be used to deter- mine €. If m = slope, and y is taken as 1.4, then . 5m T 1:65 If the test data exhibit curvature, extra- polation to M2 = 0 will still give the ambient temperature; however, the curvature indi- cates that e = f(M), and, therefore e is not a constant. In conducting temperature system cali - bration tests, it is important that the pilot make his runs at a given altitude and remain in the same air mass during all runs. Tests should be conducted as rapidly as possible to avoid the possibility of the ambient tem- perature changing during the test runs. Nor- mally no more than four points will be re- quired at any one altitude, and it is not important that the airspeed be absolutely stabilized when data are taken, inasmuch as relatively slow rates of change of speed produce negligible lag errors either in the temperature pickup or the airspeed system, Tests should be conducted at several alti- tudes to determine the effect of height on the heat loss characteristics of the pickup. Data to be obtained consist of indicated total temperature, total pressure and static pressure. Data may be recorded by the pilot or if a photopanel or transmitting equipment are installed, these may be uti- lized. No special pilot technique is required, and, if convenient, the calibration may be carried out in conjunction with other tests such as level flight runs.* . Provided the curvature is not excessive, a straight line approximation to the actual data may be used to determine an average value of e, which may then be used for cor- recting data to standard. The amount of curvature which may be ignored in a given instance is a factor which the test engineer must decide upon for himself. 1:19 CONCLUDING REMARKS ON TEMPERATURE MEASUREMENT If the test data exhibit excessive curva- ture, a convenient calibration curve may still be prepared on the basis of the equa- tion obtained by dividing both sides of Eq. 1:64 by T, thus Measurement of air temperature using equipment installed within the airplane is a relatively simple process, and the major problem is the determination of the system recovery factor. If the recovery factor is constant or nearly so, it is readily found from a plot of indicated total temperature against M?. Ti.1+() cm ( + €M2 1:66 where the ambient temperature T is obtained by first plotting the test data as in Fig.1:11 and extrapolating to Me : 0. For a variable recovery factor a tempera- ture calibration plot should be prepared in the form of a graph of Tti/T vs. M?. Knowing T, the ratio Tt;/T may then be plotted against M or M?, and the result is the calibration curve. We note that if e is insensitive to altitude variation, data from all altitudes should fall along the same line when plotted according to this latter scheme. * Note: If temperature system calibrations are conducted simultaneously with accelera- tion runs, proper account should be taken of airspeed system lags which normally are considerably greater than any lags exhibited by the temperature system. 1:38 1:20 INTRODUCTION TO LAG MEASURE- MENTS IN PRESSURE SYSTEMS application might be the measurement of surface pressures of an oscillating airfoil. Equations describing the system for this case are quite complex. Refs. 8 and 9 analyze this case. Both theory and experience show that when a pressure gage is connected by means of tubing to a source of pressure, it does not respond instantly to a change in pressure at the source but lags behind the pressure at the source by an amount called the lag error. (b) Measurement of constantly increasing or decreasing pressures or of oscillating pressures of extremely low frequency. The second major division is further divided into two subdivisions: The pressure at the source is transmitted through the tubing with lag due to: (1) pres- sure drops in the tubing resulting from the viscous friction between the flowing air and the tubing wall; (2) Pressure drops across orifices and restrictions; (3) inertia of the air; (4) the finite time required for a pres- sure disturbance to travel the length of the tube. (1) Measurements where the pres- sure lag is large compared to the pressure in the tube. This would be the case in an intermittent wind tunnel where there is a change of several atmospheres from one pressure to another in essentially step func- tion fashion, Refs. 10 and 11 analyze this case, 1 . After the pressure disturbance reaches the gage, the instrument diaphragm must deflect and move the indicating needle or actuate some other output generator. This results in additional lag due to the instrument damping, inertia and viscous and coulomb friction. (2) Measurements where the pres- sure lag is small compared to the pressure in the tube. Airspeed, altitude and Mach number meas- uring systems currently used in airplanes and considered herein consist of a gage (which is usually of the diaphragm type) con- nected to sources of total and/or static pressure by means of tubing. Sometimes other instruments, orifices and/or restric- tions are in the system. From the above considerations and the fact that in actual installations there is heat transfer between the system and the sur- rounding medium, it is readily apparent that a general mathematical treatment of the response of such a complicated system would involve at least simultaneous second order partial differential equations. Even if solu- tions could be found for these equations, the results would not be usable because of the lack of knowledge of the factors (heat trans- fer characteristics, temperatures along the tubing, etc.) entering into the equations. Analysis of flight test data from these systems comes under the category (b (2) above) of constantly increasing or decreas- ing pressure changes where the lag is small compared with the pressure in the tube, Since lag exists in airspeed, altitude and Mach number measuring systems, either correc- tions must be made for the lag errors, or it must be shown that for a particular case the error is small and can be neglected. In view of the complicated nature of the problem, analyses of pressure measuring systems are usually divided into several specialized categories which allow simplifi- cations to be made in each particular case. The two major divisions are: It is indeed fortunate that drastically simplified approaches have been found which will supply adequate lag corrections over a large range of flight conditions encompassing those presently encountered in performance testing. These simplified approaches will (a) Measurement of comparatively high frequency oscillating pressures. Such an 1:39 be considered in connection with the follow- ing: (a) The influence of system design and operation on lag. (b) Methods of correcting flight data for lag errors. (c) Present examples of lag corrections to flight data. As noted earlier in this chapter, lag is only one of the several errors for which corrections are required to observed values of airspeed, pressure altitude and Mach number. These various corrections are applied in the following order: > and of ambient conditions only, and by ground tests determine the lag as a function of these variables. A lag correction can then be ap- plied to the flight data for each flight value of ambient condition and rate of change of pres- sure, This process takes into account pos- sible changes in the character of flow in the tubing from laminar at low rates of change of pressure to turbulent at high rates. However, the flow in the tubing is laminar over the usual flight test regions for almost all air- speed, altitude and Mach number measuring systems. The assumption that the flow is laminar together with the neglect of air inertia and the assumption that the pressure lag is small compared with the system pressure, permits the use of a simple expression to describe the system. The simplified equation describ- ing the response of a pressure measuring system which is used for airspeed, altitude and Mach number flight data corrections, has been derived from two somewhat different physical pictures. It is observed that the performance of a typical system is similar to a one-degree-of- freedom damped oscillator in that there are conditions under which the system can be underdamped, critically damped or over- damped. Therefore, one approach, described in Refs. 12 and 13, is to neglect the instru- ment lag entirely and to assume that the equa- tion for the oscillator applies to the pressure system. Solutions of this equation for vari- ous forcing functions are discussed in Ref. 14. (a) Instrument scale error. (b) Lag error. (c) Position error. In this section of Chapter 1, V'cal is used to designate the indicated airspeed corrected for instrument error. 1:21 DEVELOPMENT OF EQUATIONS The most accurate method of correction of flight measurements of airspeed, altitude or Mach number for lag error would be to set up the actual airplane system in the lab- oratory and by a trial-and-error process determine the true airspeed, altitude or Mach number history corresponding to the indicated history and ambient conditions of the flight. While this procedure is possible, it is not feasible since it would require excessive time and effort, and, to the authors' knowledge, has not been attempted except for the simple case of a steady dive. Systems are underdamped when the tubing is short and the altitude is low. If buffeting or oscillating pressures are being measured, the underdamped system will exaggerate the amplitude of the oscillation although mean values will usually be given correctly. Also, an underdamped system will be excessively sensitive to step input type disturbances such as would result from encountering a gust. Therefore, underdamped systems are not usually employed for flight testing work. A somewhat less accurate but much sim- pler process is to assume that the lag is a function of the rate of change of pressure For both underdamped and overdamped 1:40 systems which are subjected to constant rate of change forcing functions, it is found that, after the transient has died out, the system behaves as with zero mass with the indicated pressure lagging behind the source pressure by a time a. In this case the sys- tem is described by the equation: to + p': p(t) 1:67 Another factor to be considered is the acoustic lag T (the time required for a pres- sure disturbance at one end of the tube to reach the other). For example, if a step function type pressure change is made to the source end of the tubing, time elapses be- fore the effect of this pressure change is first felt at the other end of the tubing and the gage pressure responds according to Eq. 1:67. . where p' is the indicated pressure p is the applied pressure P(t) = (constant) (t) for a steady rate of ( pressure change. For p(t) = (constant) (t) = kt, the partic- ular solution (steady state solution) to Eq. 1:67 is i in seconds is the tube length divided by the speed of pressure propagation which is given in Ref. 15 as 1000 ft. per sec, for small diameter tubing. The acoustic lag produces a phase shift and is included in Eq. 1:67 to yield: p': klt - 1) = p-ö'd when Ø = k do'(t+ 7) + p'lt + r) - plt) 1:69 so that P-pupa. with the solution for p(t) =kt; p'= k[(1+r)- k -1]. 1:67 A On the basis of an analogy between elec- trical and pneumatic systems, the lag con- stant is theoretically derived for laminar flow in the tubing as: Ref. 15 presents a comparison between the measured response of a system (consisting of a section of tubing and an instrument) to a very rapid change in pressure and the re- sponses computed with and without consider- ing the acoustic lag. When the acoustic lag was included, the calculated response more nearly agreed with the actual response. 0 8μ. d: nor? [ L + #p2 1:68* * Editor's note: The important thing to note here is that i depends on and pas atmos- pheric variables. This can be directly es- tablished by dimensional analysis since we know that i = f (Pw, r, L, Q) for laminar 1 flows with small pressure gradient. Since has the dimension of time, our dimensional equation is In most flight test applications, however, the acoustic lag contribution is small and can be ignored; it is considered here only for the sake of completeness. It is a simple matter to estimate the acoustic lag contri- bution for a particular system and compare this with the pressure lag to determine whether or not the acoustic lag can be ig- nored. Sample calculations ignore the acous- tic lag. M T: M\b (L) (L from which ()° (Lvºrnier de of [66%) . 60W Another approach, given in Refs. 16, 17 and 18, is to compute the lag on the basis that conditions are steady; i.e., P = constant. Instrument effects are neglected. The de- rived equation is exactly the same as the equation derived through the first approach except that the factor (L + Q/#r?) becomes : 1:41 (L/2+ Q/tr?) and again t can be ignored for steady flow. The approximate relation between compres- sible dynamic pressure, total and static pressure, and calibrated airspeed for sub- sonic flow is: ac - -ps 2 = { POV calli+ Pov? allit (... : 1 ( cal 4 Со + 1:70 The further assumption is made that Eqs. 1:67 and 1:69 can be applied to forcing func- tions other than constant rate of change. A very simple method of correcting pressure data for lag results from this assumption. The correction, in terms of pressure, de- pends only on the indicated pressure, the time rate of change of indicated pressure and the lag constants, i and . The time rate of change of indicated pressure and the indicated pressure are measured in flight while and are determined from ground tests. Applying Eq. 1:69 to the static and total pressures of Eq. 1:70 and defining the lag correction as Aq 9c-4c gives: Aqc aclt+r)-akt) +A4 t + ) +(4, -asip's It + ) The lag of the instruments alone has been examined by various investigators. For ex- ample, Ref. 19 has an analytic evaluation of altimeter lag and Ref. 17 has an analytic evaluation of airspeed indicator lag. Ref. 16 deduces the airspeed indicator and alti- meter lag from test measurements while Ref. 20 gives lags of altimeters actually subjected to pressure conditions simulating dives and climbs. 1:71 where it . lag constant of total pressure system and λς lag constant of static pressure system. The sum and substance of all these anal- yses and tests are that by proper design and selection of the instruments, the instrument lag can be made small and may therefore be neglected. However, care must be taken to insure that the instruments actually used have low lag. Instrument lag is partially taken into account by making the necessary ground calibrations on the complete systems including the instruments actually used for flight measurements. This equation shows that the lag correc- tion in terms of Aqc at a given time (t) can be obtained from measurements of a'c at that time and of ac, P's and ac at time (t + ) plus the values of the total and static pressure lag constants. > It then remains to transform Eq. 1:69, which is expressed in terms of pressures, into terms of altitude, airspeed or Mach number the desired quantities. The values a'c and q'c are computed from airspeed measurements and Eq. 1:70. The ambient pressure and time rate of pressure change are obtained from altitude time his- tories, knowing the altimeter calibration equation and the scale error correction. The lag constants are obtained with the relation: (a) Airspeed X 1=106) (。 : λο The airspeed correction could be ex- pressed in terms of either compressible dy- namic pressure, qc, or indicated airspeed. which for the total and static lines become 1:42 (see Eq. 1:68) 1:70, we have DSO V'2 19 = d0 666) 1 cal . da'c = { po 2v Cal dVcal Ps + ac va • 2V caldv.cal + Ź pov?com is ..(post ) Aso leo Po col 4 Ро Ps Со 1:72 1:74 whence Fig. 1:12 presents the variation of the ratio is so and 14/1tol with standard al- titude and speed. iso and 11. are obtained from ground tests as described later in the chapter. da's i'ca sevceloveolit dhe p dv 1+ 2 Со and Usually the acoustic lag is small compared to the total lag and may be ignored under these circumstances. The terms of Eq.1:71 in the rectangle drop out and the lag correc- tion at time t is equal to the sum of the re- maining terms on the right-hand side of the equation taken at time (t) instead of at time (t+), i.e., cal ac da's dt on Povcalcal + 2 Со 1:75 Also, from the atmosphere equation of balance Aqc= d70'c + (14 - Asos' =doen- isüs'. do's : - godno gpaho 1:76 so that 1:73 p's -- gprie 1:77 Letting dac and dVcal become finite dif- ferences in Eq. 1:75, and substituting with Eqs. 1:75 and 1:77 in Eq. 1:73, we have Avcol : dqV cal Eq. 1:71 should be used for airspeed lag corrections when the acoustic lag is to be taken into account. For the more usual case where the acoustic lag is small enough to be ignored, it is usually convenient to use an equation expressed directly in terms of air- speed and pressure altitude and their time derivatives (as well as the lag constants). The equation is based on the assumption of a small error where differentials are equal to differences, and while sufficiently accurate for small corrections has an increasingly large error as the corrections become large. The equation is derived as follows: (Ar-dsigphp Ncoll? Percoilith , Со 1 11 . gorp (Ag-As) - dovical /Vcal cal 2 1:78 where o has its standard atmosphere value at the test pressure altitude. Voorlitz . 1 Taking differentials of both sides of Eq. 1:43 6 5 ASL 4 λ./λεsL .L. a / S.L. STATIC λ./λς. TOTAL 3 001 175 200 250 2 300 350 400 450 500 550 600 Z 650 C Vcal. - kts. SEA LEVEL 10,000 20,000 30,000 40,000 50,000 mp, PRESSURE ALTITUDE - FT. Fig. 1:12 1:44 Fig. 1:13 presents a plot of and for y: 1.4 272.5 go cal dac [tomonga ( +0.2 = f(hp, Vald P.Vcal d Vcal Vcor ( i + Vesty 2 со 1:80 so that 272.5 as a function of calibrated airspeed and pressure altitude. Vcal ac po Vcal - boxeal [0.2 (en in Vcal со 1:81 Still more accurate but somewhat more complex equations are obtained by differ- entiating the actual calibration equation and Eq. 1:73 gives us goholupds! . : z face + 93-067 - -1)/y AVcal = lyVcal vCal . -1 [) Ро Vcoito ay may 1:82 whence We note that the quantity - dac 2 Vcal d Vcal 2 Y PO / 9c Yol Po +1 Vy Y- Y go Ро Ро 1272.5 Ncall Vcal I+0.2 1 ol[ Со or ac has very nearly the same value as the ap- proximation to it f(hp, Vcal) used in Eq. 1:78. : +)" dac - povcal d Vcal ilo Ро 1:79 Now, from the calibration equation y-1) [(y-1) Vºcal ac Po (b) Altitude The pressure altitude may be obtained in terms of either ambient pressure or feet of altitude, If an exact difference method is desired, the altimeter calibration (Ref. 21) is operated upon in the same manner as the dynamic pressure calibration in the previous section, Included herein is a differential correction method which provides a correc- tion in terms of altitude. The equation is +1: + ya 28 so that Eq. 1:79 becomes 1/(Y-1) dac - In the Moon )+]" [ Anplt) = holt + o) - hp (t) +ishpltto). Со · Po Vcal d Vcal 1:83 1:15 650 Vcal [1+į i vode] go VARIATION OF i hp, Vcol) Vcal co WITH PRESSURE ALTITUDE AND CALIBRATED AIRSPEED .20 .040 Fig. 1:13 Variation of .18 .035 .16 .030 .14 .025 Vcalli+ė (col :.12 Altitude and Calibrated Airspeed Со go f( hp, Vcol.) ..10 fl hp, Vcal. 으 ​.020 .08 SEA LEVEL (Pressure Altitude) .015 10,000 000's +--/--/- .06 20,000 .010 10,006 15,000 20,000 SEA LEVEL (Pressure Altitude) = f(hp, Vcall with Pressure 2017 .04 25,000 30,000 30,000 35,000 .005 .02 35,000 40,000 45,000 50,000 40,000 45,000 50,000 55,000 55,000 60,000 60,000 100 150 200 250 300 300 350 400 450 500 550 600 CALIBRATED AIRSPEED, Vcal, KNOTS 1:46 This equation is differentiated and solved for dM, The correction given by Eq. 1:83 is added to the indicated value of pressure altitude to obtain the corrected value. Again, the first two terms of the right side of Eq. 1:83 represent the acoustic lag and may be ignored in most cases; in those cases, Dolly-1)/y -1 i dpt dps dm : ( IPs YM Do Ps 1:85 Ane : isho. Дhp 1:84 Thus the lag correction can be obtained from a history of indicated altitude and the ground measured lag constant. Since the assumption is made that the lag is small compared to the pressures, dM is small compared to M, dps is sn:all com- pared to Ps, and dp, is small compared to Pt. Eq. 1:67 gives the lag of the total and static pressure sides of the system as: (c) Mach Number Ds CD's - Aps - dps . - λsβς Dr-p'- Apr : dpt = lopte 1:86 Since the Mach number is a function of the two pressures P, and Ps, it is evident that a lag correction to Mach number must include functions of p, and Ps and their time derivatives. Calibrated airspeed and pres- sure altitude, or Mach number and cali- brated airspeed, or Mach number and pres- sure altitude might be used. If the Mach number is computed from measurements of calibrated airspeed and pressure altitude, as is usual in flight testing, the lag correc- tions should be applied to these quantities individually prior to the Mach number cal- culation. Eqs. 1:86 are substituted into Eq. 1:85 to give the Mach number lag in terms of total and static pressures and their time deriva- tives, namely, Ps AM: (" ) Pty-1/7 Ps 。 is DS YM 1:87 To determine py/Pa and Ps/Ps we solve Eq. 1:84 for pp/Ps to obtain If, however, the Mach number is obtained from a direct reading Machmeter, the most convenient lag correction would include the indicated Mach number and indicated alti- tude terms. If the Mach number is less than one, the equation, assuming small lag and neglecting the acoustic lag, is obtained as follows: PT Y-1 ylly-1) . MS 블 ​] +1 Ps 2 1:88 and on differentiating, At subsonic speeds the Mach number is related to the static and stagnation pressure (page 36 of Ref. 21) by the equation dpt dps PA Ps 2 Ps? -1) Y 11 M2 y M2 = 2 Y-1 Pilly-1)/y IPs - " ( 23 ) +1]*** (I) (2 MdM) 1:84 2 1:47 : or also, from the equation of state p: P/RT so that dpt dps ៩ Ps ghip hip -1 Ds Ps RT 53.3T = y MOM ".(و ، (2) lu«(). )"+". /( +1 1:91 where T is in degrees Rankine. Using Eq. 1:88 for Ps/P7 In the standard atmosphere dpt T: To-.00357h = 518.4-.00357h y MdM dps + Ps Po Y1 [ m2 (4:1) up to the stratosphere, whence T: 392.4° R at 35,332 feet and above Dr. PS умм in the NACA atmosphere and PT + Ps *[w(37)+] M2 2 T : 389.7° R at 36,088 feet and above 1:89 in the ICAN atmosphere. Substituting Eq8, 1:88 and 1:89 in Eq.1:87, we have AM: M2 +1 CH X 2 YM M Inasmuch as the altimeter calibration equation assumes standard density and tem- perature at a given pressure altitude, the temperature in Eq. 1:91 is the standard at- mosphere value at the test pressure altitude. Substituting Eq. 1:91 in Eq. 1:90 and setting y = 1.4, we find that (letting M = M' and hphp) Ps AtY Mм (dy-1s) + Ps 1039 Y-1 2 M2+ 1:90* Now we have the atmospheric balance equation as AM = {1+.2M") (:2719 [ х 1.4MMA dp = - pgdhp hip (λι - λς) 53.3 T بود 2 1 +.2M2 1:92 and Ø : - pghp ; where, as we have already noted, T has the standard atmosphere value corresponding to hip, and 14 and is are obtained by means of ground calibrations and Eq. 1:72. * Editor's modification from here to Eq. 1:101. For the supersonic condition, the Mach- meter is calibrated according to Eq. 1:22, 1:48 which may be written Now OPII dporu Pri 166.9 M2 Ps 2.5 dps Asps (78) so that 1:95 becomes 1:93 Taking differentials of both sides, we have > ܐ is dm = M[7M2- 1] [Arena 7(2 M2-1) Pol ps dpti Pr, dps 1:96 Ps PS 2 MdM(166.9) Moreover, from Eq. 1:94 2.5(166.9)M2 Х 2.5 3.5 1- 7 - (-) 7 M2 M2 PS Elő - 7M ( 2M2 - 1) (7M2 1) M Ps 2 DM M3 ) or or eti Pς 7 M ( 2M2 |dpt dos -1) 1) Pti Ps M ( 7M2 1:97 Pri Ps Ps 2 d M(166.9) 2.5 M- M and on substituting Eq. 1:97 in 1:96 and sim- plifying 2.5 PRI 7 M M2 OM : D Mar-as) ( 7M? - 1) Ps 712M2-1) + үм 1:98 Using Eq. 1:93, this simplifies to Using Eq. 1:91, which holds at any speed dpt dps 1) DM ( 2M2 7 M ( 7M2 dm = may hip M14-238( 7M2 - 1 ) (-as) ( 1 53.37 17/ 12M2-0) PHI Ds 1) 1:94 1:99 or Letting dM go to AM and writing this in terms of indicated quantities dpt dps] dM M (7M? - 1] 712M2-1) PH Ps HOM Atels AM= 'p- ( 373.11 2M2-1 ( 7M2 - 1) 1:95 1:100 1:49 where T has the standard atmosphere value corresponding to hp . Eq. 1:68 can be used to compute the lag constants of simple systems. Usually a value of n = 1.0 is used for the computation. Ref. 16 presents equations for computing the lag constants of more complicated systems. Whereas at subsonic Mach numbers, it is permissible to use ground calibration values of Ito and iso together with Eqs. 1:62 to obtain it and is , considerable error may result if this procedure is employed super- sonically for the reason that the stagnation temperatures are considerably greater than ambient in the latter case. Experience has shown that the actual lag constants are related to the calculated lag constants by the relation (actual) = KA 1 = (calculated) where K is a correction factor which accounts for the unknown effects of orifices and roughness. This factor varies from 1 for a system consisting of a single instrument and a straight tube to as high as 5 for complicated systems with rough tubing. A possibly satisfactory procedure under supersonic conditions would be to use the indicated total temperature for computation of the absolute viscosity coefficient of the airspeed system. If Hti is this coefficient, then (a) Step Function Input Test Method λς : λςο : iso Centro vtipso PSO p's ) and РО The step function input test method con- sists (as may be surmised from the name) of applying a step function input to the source end of the system and measuring the response. Eq. 1:67 is solved for a step function input and the solution rearranged to obtain; PSO dy a dro Pti t 1:101 d: loge where pt, is the total pressure existing aft of a normal shock computed from the indi- cated Mach number reading. (p - p') at t=0 (p-platt=1 1:102 The corrections given in Eqs, 1:92 and 1:100 are added to the indicated values to obtain the corrected Mach numbers. where p is the pressure p' approaches. 1:22 . DETERMINATION OF THE LAG CONSTANT Eq. 1:102 shows that the lag constant is the slope of the response curve, P-P' vs. t, plotted on semi-log paper and is the time for the pressure difference to drop to 1/e (.368) of its initial value. First, let it be clearly understood that although there are methods of computing the lag constant, flight corrections should never be based on computations but should always be based on experimentally determined val- ues. The computed lag constant is useful only as a means of roughly estimating be- forehand the approximate magnitude of lag error for a given system and flight condition. If the system is underdamped and thus not accurately defined by Eq. 1:102 or because of the instrument contribution, actual time- histories can have an oscillation superim- posed on the logarithmic decay somewhat as shown in Fig. 1:14. 1:50 A sample lag constant determination is shown in section 1:24. (b) Steady v Test Method For these cases the lag constant is obtained from a mean line drawn through the oscil- lation. If the oscillation is large in com- parison to the time constant, large errors will creep in and the lag constant should be determined by the steady time rate of change of pressure method described below. Tests may be made with step functions of different magnitudes and the resulting lag constant plotted against step input magnitude and extrapolated to zero magnitude. Other- wise, the step should be made small enough to avoid violating the assumptions that the flow is laminar and that the pressure change is small compared to the pressure in the tube. Ref. 17 gives the maximum pressure drop per foot of line for which laminar flow can exist as Although this method has not been used to date by the authors, apparatus for these tests is now being built. The method has been used by others, (Refs. 22 and 23) and is useful for underdamped systems; there- fore, it will be briefly described. A constant rate of change of pressure = k) is applied to the source end of the system. Under these conditions, Eq. 1:67A describes the steady state variation of p', and from this equation we have Δρ 3.745.10-6 8p3 Ib. per sq. ft. P-p': pa. L 1:103 Also, Ap/p (ambient) should be less than .03. Thus a plot of p and p' or t will provide two parallel lines as shown in Fig. 1:15, once the transients have been damped out. Pop' D AP p-p' P اد (D-D') TIME, 1 TIME, 1 Fig. 1:14 Fig. 1:15 1:51 From Fig. 1:15 we see that from similar triangles Ap = p-p' for steady state motion so that Eq. 1:67A becomes replaces the instrument by an equivalent volume (Ref. 22) is shown in Fig. 1:16. Another apparatus which includes the test instruments in the airplane (Ref. 23) is shown in Fig. 1:17. Др л Δρ : At (c) Sample Corrections to Flight Data or At:1. Hence a measurement of the time separa- tion of the curve of p and p' at any given pressure value determines the lag constant. Application of the constant p test technique has been limited in the past because of the necessity of removing the airspeed (or alti- meter or Machmeter) system from the air- plane for the test. One such apparatus which Sample lag corrections to airspeed and altitude test data are given in section 1:25. The calculations do not include the acoustic lag correction which is small for the ex- amples presented. It is emphasized that, to correct for lag, time histories of the quantities are needed which in turn requires a photopanel or os- cillograph installation in the airplane. No sample Machmeter Mach number lag cor- rection is included since no actual examples are available at present. System Source Pressure Gage For Measuring Source Pressure System Tubing Tubing 7 7 SOTO Atmosphere To Vacuum Pump I Used To Evacuote System) Added Volume VIIMT 수 ​AD U Tube Water Manometer for Measuring Pressure Log AD Fig. 1:16 1:52 Enclosure Static Source रु. Total Source Altimeter Counter Pitot Head Photopanel Storage Tank Needle Valve Camera b To Vacuum Pump Counter Intervelometer Controlling Counters Altimeter Airspeed Indicator Photopanel Switch Controlling Intervelometers And Cameras Camera Fig. 1:17 1:53 (d) Accuracy of the Correction Although an accurate error analysis has not been made, it is felt that the estimate in Ref. 22 of £20% accuracy is probably quite close to the truth. This accuracy means that the system should be proportioned to make a 20% error in correction lie within the required accuracy of airspeed, altitude or Mach number. For example, if the allowable error is airspeed due to lag is 1 knot, then the lag correction must be held to 5 knots or less. relative to the instrument. In some cases it is possible to locate the instruments (or associated recording equipment) close to the sources of pressure as, for example, lo- cating a photopanel with airspeed meter in a fuselage nose close to a nose boon pitot- static pickup. In other cases it is not pos- sible to locate the instruments close to the pressure sources as, for example, when a trailing static head is required to hang a wingspan or more below and behind the fuse- lage. 1:23 DESIGN OF SYSTEMS In all cases the tubing should follow as direct a path as possible and should be free of sharp bends. The tubing should have a smooth interior finish, the diameter should be as large as practical, and the pressure orifices should have areas equal to the tubing diameters, The general arrangement and location of the static and total pressure sources of air- speed, altitude or Mach number measuring systems are usually dictated by require- ments other than low lag. However, within the specified general arrangement, the de- signer should arrange the variables under his control to either reduce the lag to obtain the required accuracy of the correction, or, more preferably, to eliminate the need for a correction. The system designer usually has some control over the following variables that affect lag: Controlling the relative lag of the total and static pressure sides of the system can be a powerful means of reducing the airspeed and Mach number lag in flight during specified maneuvers. For example, if the lags of the total and static sides are equal, there would be no airspeed lag during steady speed climbs and descents. If the lag of the total side is negligible, the airspeed lag would be negligi- ble during constant altitude increasing speed tests regardless of the static pressure lag error. Ref. 15 considers the effect of relative lag on the Mach number lag error. The alti- meter lag depends only on the lag of the static pressure system. (a) Instrument characteristics. (b) Number of instruments in the system. (c) Tubing length, inside diameter and surface finish. . (d) Relative lag of total and static pres- sure sides of system (this influences the airspeed and Mach number lag). 1:24 CONCLUDING REMARKS Ref. 17 shows that instrument lag is mini- mized by selecting instruments with a high natural frequency, low damping and low fric- tion. Ref. 20 discusses altimeter lag. In general, the number of instruments in the system should be kept to a minimum in order to reduce the volume and complexity of the system (and the lag). The lag errors in airspeed, altitude and Mach number measurements have been dis- cussed and examples of corrections for these errors presented. Maneuvers during which significant lag errors have been encountered have included stalls and high speed dives. Lag errors must be taken into account either by showing that the corrections are small and can be discarded or by correcting the flight data. The pitot-static system should be de- signed with the view of minimizing lag errors insofar as this is practical. The minimum tubing lengths are usually set by the locations of the pressure sources 1:54 1:25 SAMPLE DETERMINATIONS OF AVcal AND Ahp Lag Correction to Altitude: Lag Correction to Airspeed: AT AVcal = it Anp = + so is hip Aso to + Vcal + (aso , hp [fhp. Vcal) ] ito As - xto xto iso PLAA MANEUVER STALL TIME QUANTITY SEC UNITS FLIGHT DATA FROM STEP 0 1 2 3 1 5 6 7 1 ho Ft Flight Data 5, 020 4,850 4,695 | 4, 580 | 1, 535, 4, 740 5,0155, 330 10,006 2 Ft Instr. Corr. 5, 024 4,853 4,698 4, 582 4, 537| 1, 7131 5,019 5, 334 10,000 Sec Lag Check # .556 .556 . 556 . 556 . 556 . 556 . 556 .556 . 50 4 1.15 1. 15 1.17 1.14 1.13 1.14 1.15 1.16 1.36 1: 3 150 + / λς/λso hip Ahp hp 5 Ft/sec. Flight Data -190 -170 -130 -100 0 +300 +320 +320 -20* 6 Ft 3 x 4 x 5 -121 -109 -82 -63 0 +190 + 205 +206 -14 7 Ft 2 + 6 4, 903 1, 744 4,616 4,519 4, 537 4,933! 5, 2245, 540 9, 986 8 Kts V cal Flight Data 558.0 558.0 557.0 553.5 542.0 531.5 523. 1 510.0 100.8 9 V'cal Kts Instr. Corr. 555.8 556.3 554.8 551.3 539.8 529.0 520.3 507.3 100.0 01-1 10 V'col +3.0 0 -3.0 -9.0 -11.0 -10.8 -11.50 -13.0 -1.0* Kts/sec | Flight Data Sec Lag Check 11 . 086 .086 .086 .086 .086 .086 . 086 .086 .086 12 .69 .69 .68 .68 .71 .72 .73 .74 1. 34 Ato λ/λfo dVca! /dhp dy 13 .0127.0128.0130 .0131 .0136 .0139.0145 .0145 .089 1/sec Kts 14 V cal 10 x 11 x 12 +. 178 0 -. 175 -. 527 -.672 -. 669 -.722 -.829 -.115 xto ܘܪܐ ܘܪܬ 15 λς. so Kts 3 x 4 x 5 x 13 -1.537 1.395 -1.066 -. 825 0 +2.641 +2. 973 +2.987 -1.246 iso As hip it no d Vcal dhp dVcal hp dhp 16 -Xto Kts -5x11x 12 x 13 +. 143 +. 129 +.099 +.077 01 -. 258 -. 291 -. 295 +. 205 ito ito 17 AV. cal Kts 14+ 15 + 16 -1.2 -1.3 -1.1 -1.3 -.7 +1.7 +2.0 +1.9 -1.2 18 Kts Vcal 9 - 17 554.6 555.0 553.7 550.0 539.1 530.7 522. 2 508. 2 98.8 * Obtained from plotted time history of stall #Similar to Fig. 1:18 1:55 100 90 80 70 C STATIC SYSTEM 60 TOTAL SYSTEM 50 ) ( AT .368 STEP FUNCTIONS 2 58.0 LBS. / FT. = 21.3 LB/FT:2 λς = 0.52 SEC. AT .368 ( 64.5 LBS./F1.2, :) = = 23.7 LB/FT.9 40 30 is=0.07 SEC. .07 SECA -.52 SEC 20 (9:-9c4 LBS./FT.2 10 9 8 7 6 NOTE : 5 RESPONSE MEASURED BY AIR SPEED INDICATOR 4 3 O 0.25 0.50 0.75 1.00 1.25 1.50 TIME - SEC. Fig. 1:18 1:56 REFERENCES 1. Weaver, “The Calibration of Airspeed and Altimeter Systems,"U, K. Ministry of Supply Report No. AAEE/Res/244, August, 1949. 2. Smith, “The Measurement of Position Error at High Speeds and Altitude by Means of a Trailing Static Head," U. K. Ministry of Supply, RAE Technical Note No. Aero, 2163, June, 1952. 3. Zalovcik, “A Radar Method of Calibrating Airspeed Installations on Airplanes in Maneu- vers at High Altitudes and at Transonic and Supersonic Speeds," U.S.A. NACA Tech- nical Note 1979. 4. Huston, "Accuracy of Airspeed Measurements and Flight Calibration Procedures,” U.S.A, NACA Report No. 919, 1948. 5. Rogers and Berry, "Tests On The Effects Of Incidence On Some Pressure Heads At High Subsonic Speeds,” U. K. ARC Report No. 13,263, July, 1950. 6. Gracey and Scheithauer, “Flight Investigation of the Variation of Static Pressure Error of a Static Pressure Tube with Distance Ahead of a Wing and a Fuselage," U.S.A. NACA Technical Note 2311, March, 1951. . 7. Ruskin, R. E., Schecter, R. M., Dinger, J. E. and Merril, R, D., "Development of the NRL Axial Flow Vortex Thermometer,” NRL Report No. 4008, September 4, 1952. 8. Tabach, Israel, “The Response of Pressure Measuring Systems to Oscillatory Pres- sures," NACA Technical Note No. 1819, February 1949. 9. Iberall, Arthur S., "Attenuation of Oscillatory Pressures in Instrument Lines," U.S. Department of Commerce, NBS Research Paper RP2115, Journal of Research of the NBS, Vol. 45, July, 1950. 10. Kendall, J. M., “Time Lags Due to Compressible-Poiseuille Flow Resistance in Pres- sure-Measuring Systems," NOL Memo No. 10677, May 4, 1950. 11. Sinclair, Archibald R. and Robins, Warner A.,“A Method for the Determination of the Time Lag in Pressure Measuring Systems Incorporating Capillaries," NACA Technical Note No. 2798, September, 1952. 12. Weidemann, Hans, "Inertia of Dynamic Pressure Arrays,” NACA Technical Memo No. 998, December, 1941. 13. DeJuhasz, Kalman J., “Graphical Analysis of Delay of Response in Air-Speed Indicators,” Journal of Aeronautical Sciences, Vol. 10, No. 3, March, 1943. 14. Draper, C. S. and McKay, Walter, "Instrument Analysis," MIT, 1943-44. 15. Huston, Wilbur B., “Accuracy of Airspeed Measurements and Flight Calibration Pro- cedures,” NACA Technical Note No. 1605, June, 1948. 1:57 REFERENCES 16. Charnley, W. J., "A Note on a Method of Correcting for Lag in Airspeed Pilot-Static Systems," RAE Report No. Aero 2156, September 1946. 17. Schwarzbach, J. M., “Lag Correction to Flight Measurement of Airplane Stall Speed," M. S. Thesis for the University of Maryland, 1953. 18. Wildhack, W. A., "Pressure Drop in Tubing in Aircraft Instrument Installations," NACA Technical Note No. 593, February 1937. 19. Head, R. M., "Lag Determination of Altimeter Systems," Journal of Aeronautical Sci- ences, Volume 12, No. 1, January 1945. 20. Johnson, Daniel P., "Calibration of Altimeters Under Pressure Conditions Simulating Dives and Climbs,” NACA Technical Note No. 1562, March, 1948. 21. Dommasch, D. O., Sherby, S, S., and Connolly, T. P., "Airplane Aerodynamics," Pitman, New York, 1951. 22. Smith, K. W., "Pressure Lag in the Piping of the MK V Trailing Static Head," RAE Technical Memo No. Aero 258, May, 1952. 23. Herrington, Russel M. and Schoemacher, Paul E., "Flight Test Engineering Manual," USAF Technical Report No. 6273, May, 1951. 24. Schaefer, Herbert, "Machmeters for High-Speed Flight Research,” Journal of the Aero- nautical Sciences, Vol. 15, No. 6, June, 1948. > 25. Swanson, W. E, and Gray, A, K., "Methods of Flight Test Performance Data Reduction for Turbojet Propelled Airplanes, "North American Aviation, Inc., Report No.NA-47-1033 of October 24, 1947. 26. Dommasch, D. O., et al, “Flight Test Manual," Part I, Revised Edition, Preliminary Copy, NATC, Patuxent River, August, 1953. 1:58 AGARD FLIGHT TEST MANUAL VOLUME I, CHAPTER 2 THRUST AND POWER DETERMINATION Ву Daniel O. Dommasch Princeton University Section 2:1 Walter J. Hesse Naval Air Test Center United States Navy Sections 2:2, 2:3, 2:4, 2:5, and 2:6 Jean Soisson Centre d'Essais en Vol, France Sections 2:7, 2:8, 2:9, and 2:10 John K. Moakes Aeroplane and Armament Experimental Establishment United Kingdom Sections 2:11, 2:12, 2:13, and 2:14 Rene Le Duc France Sections 2:15 and 2:16 Kenneth J. Lush Air Force Flight Test Center United States Air Force Section 2:17 David W. Bottle Aeroplane and Armament Experimental Establishment United Kingdom Sections 2:18, 2:19, and 2:20 VOLUME I, CHAPTER 2 CHAPTER CONTENTS Page TERMINOLOGY 2:1 INTRODUCTORY COMMENTS 2:1 2:2 POWER MEASUREMENT; INTERNAL COMBUSTION ENGINES 2:2 2:3 THEORY OF RECIPROCATING ENGINE POWER CORRECTIONS 2:6 2:4 CORRECTIONS APPLICABLE TO POWER CHARTS 2:7 2:5 TORQUEMETER POWER CORRECTIONS 2:8 2:6 POWER MEASUREMENT - COMPOUND ENGINES 2:9 2:7 JET THRUST MEASUREMENT - INTRODUCTORY COMMENTS 2:9 2:8 DISCUSSION AND REGION OF APPLICATION OF THE METHODS OF JET THRUST MEASUREMENT 2:10 2:9 THE JET FLOW MEASUREMENT METHOD 2:10 2:10 THE CLIMB PERFORMANCE METHOD 2:13 2:11 THE MEASUREMENT OF THE USEFUL THRUST OF TURBO-PROPELLER ENGINES 2:15 2:12 MEASUREMENT OF SHAFT HORSEPOWER 2:17 2:13 THE ESTIMATION OF JET THRUST 2:17 2:14 CONCLUDING REMARKS ON TURBO-PROPELLER THRUST DETERMINATION 2:18 2:15 RAMJET THRUST MEASUREMENT 2:18 CONCLUDING REMARKS ON RAMJET THRUST MEASUREMENT 2:16 2:21 2:17 MEASUREMENT OF ROCKET THRUST IN FLIGHT 2:22 2:18 GENERAL ANALYSIS OF JET THRUST MEASUREMENT 2:23 2:19 THE MEASUREMENT OF THE THRUST OF JET ENGINES 2:28 2:20 THRUST MEASUREMENTS WITH AFTERBURNING 2:32 REFERENCES 2:35 TERMINOLOGY N Engine RPM Q Torque Pressure K Torque Constant Pb BMEP: Brake Mean Effective Pressure L Length of Stroke A Piston Area Pm, MAP Absolute Manifold Pressure BHP Brake Horsepower P Air Density po Sea Level Standard Density Ga Weight or Mass Air Flow Rate а (f/a) Weight Fuel-Air Ratio H.V. Heating Value of Fuel Пт Thermal Efficiency e Volumetric Efficiency Pm Manifold Density F Thrust Fo Gross Thrust F Net Thrust N р Static Pressure А. Nozzle Area V True Airspeed Q Mass Flow y Specific Heat Ratio C с/сь Gas Constant R TERMINOLOGY (Continued) T Ambient Temperature 8 Pressure Ratio A Effective Nozzle Area F pre Pre-Entry Thrust F int Internal Thrust F post Post-Exit Thrust F SN Net Standard Thrust FSG Gross Standard Thrust F JN Net Jones Thrust FJG Gross Jones Thrust Fв Trunnion or Bearer Thrust u V cos • : Velocity Component in Free-Stream Direction V Resultant Velocity 6 Angle between Flow and Free-Stream Direction Local Swirl Angle Subscripts m Manifold Conditions e,ex Exit Conditions t Total Conditions, also Tailpipe 8 Free-Stream i Intake e Exit W Far Downstream Superscripts (') (prime) Test Conditions 6 在 ​@ 2:1 INTRODUCTORY COMMENTS (3) Effects of engine cooling on climb and range performance, In conducting flight tests, the power plant is used not only as a means of propelling the airplane, but also as a means for measur- ing the drag characteristics of the airframe. Thus, when we measure thrust and/or power, we generally have two separate purposes in mind: (4) Effects of piston engine exhaust stack thrust on apparent power required curve, (5) Effects of induction system on power available. (1) The determination of the capabil- ities of the propulsive system as installed in the airplane; (b) Jet Type Aircraft (2) The measurement of the drag characteristics of the airframe in the pres- ence of the operating propulsive system. (1) Effects of entrance air spillage on drag characteristics of airframe (sub- sonic). (2) Effects of external induced flow produced by internal engine flow. Because the airplane (considered here as composed of an airframe and a powerplant) functions as a unit, we cannot, in a general sense, completely divorce engine and air- frame characteristics from one another. This holds true for performance analyses and also for stability and control investigations where engine operation, particularly at low airspeed at full throttle may profoundly affect the air- plane's moment balance. (3) Effects of inlet design on flow characteristics in jet intake and compres- sor. (4) Effects of entrance shock waves on airframe drag (supersonic). (5) Effects of yawing and pitching moments produced by the changing direction of airflow passing through the engine(s). Although interaction effects may be small in many cases, they are a possible cause of failure to obtain generalized data using reduc- tion processes which ignore them. Thus, even though possible interaction effects are not considered in detail in all portions of this manual, they should be kept in mind and made the subject of investigation in cases where one suspects that poor performance is attrib- utable to such effects. A partial list of inter- actions is as follows: (a) Propellered Airplanes (1) Effects of airplane configurations on propeller efficiency (cowl flaps open or closed, etc.). (2) Effects of propeller on airplane drag characteristics, stalling characteris- tics, stability characteristics, airspeed sys- tem calibration, etc. In some cases, the method of conducting tests and of reducing the data tacitly takes into consideration interaction effects, at least in part. For instance, this is the case for the reduction procedures described in Chapter 6 for piston engine aircraft, where- in the brake horsepower required rather than the thrust horsepower required character- istics are determined so that direct meas- urement of airframe drag is avoided. Meth- ods such as these are generally satisfactory, provided attempts are not made to gener- alize data over too great a spread of al- titudes or other operating conditions. 2:1 power output should be for the engine as installed in the airplane and such charts should be used with care by the test en- gineer. The measurement of items such as shaft brake power of a piston engine or the gross thrust of a jet engine is a rather simple pro- position and direct measurements are possi- ble. On the other hand, measurement of thrust power of the piston-engine-airplane combination or of the net thrust of a jet en- gine is not so simple a matter because inter- action effects are likely to enter the picture. Indeed, for thrust power developed by a piston engine we must normally assume a propeller efficiency since we cannot readily measure the efficiency in flight. In the following sections of this chapter, the detailed analyses of thrust and power measurements for various types of engines are discussed. In editing this work, which has been contributed by several of the NATO nations, the principal modifications have been to standardize terminology. 2:2 POWER MEASUREMENT; INTERNAL COMBUSTION ENGINES Furthermore, propeller charts prepared either from computations or from wind tunnel tests very seldom represent accurately the propeller operation as it actually occurs in the presence of the airplane. In the case of a jet engine, the question arises as to whether or not the possible drag changes due to engine operation should be subtracted from the engine thrust or added to the air- frame drag. The question further arises as to the proper value of ram drag which is normally computed on the basis of all the engine intake air obeying the laws of one- dimensional flow and on the basis that there is no spillage of the entering air. The power of an aircraft internal combus- tion engine is determined either by a torque- meter or by the use of power curves (prepared by the engine manufacturer, or by a suitable engine testing laboratory). The preferred method is actually to measure the power with a torquemeter. For engines not equipped with torquemeters, reasonably accurate values of power can be obtained from the power charts. In a standard torquemeter installation, the torque force is balanced by an oil pressure which acts on small pistons located around the periphery of the reduction gear housing. The brake horsepower is then given by the following equation: BHP = KNQ 2:1 For practical purposes, it is generally sufficient to work on the basis that the ram drag of the engine is equal to the rate of change of momentum of the air entering the engine, and that the net thrust of the engine is equal to the difference of the gross thrust produced aft of the turbine and the ram drag computed as noted above. This may produce difficulties in drag data generalization if air spillage exists under some circumstances and not under others. However, since air spillage is an indicator of faulty design, the fact that the tests reveal its existence should not be construed as a deficiency in the test method or of the definition of terms used. Finally, a general word of caution: en- gine charts prepared on the basis of ground tests and/or computations are seldom a sufficient basis for extrapolating test data to determine what the standard thrust or where N = the engine RPM Q Q = torque pressure - K = torque constant. From Eq. 2:1 it is evident that it is a simple matter to obtain the brake horsepow - er of an engine when the torquemeter oil pressure and the engine RPM are known. This method of measuring power has been 2:2 used very successfully, and when the torque- meter equipment is properly installed and maintained, precise power measurements may be obtained. horsepower by replacing the indicated mean effective pressure with the brake mean effec- tive pressure, Pb; i.e., POLAN Brake horsepower can be expressed by an equation similar to the equation for indicated BHP : . РЫ .NK. 33,000 2:2 BRAKE HORSEPOWER FULL THROTTLE LINE 2,600 R.P.M. 2,400 R.P.M. 2,200 R.P.M. 2,000 R.P.M. 9 MAP- ABSOLUTE MANIFOLD PRESSURE Fig. 2:1 Sea Level Calibration Curve 2:3 Thus the torquemeter oil pressure is a direct function of the brake mean effective pressure. Because of this fact, many torque- meter installations are calibrated so that the torquemeter oil pressure indicates values of brake mean effective pressure instead of the actual oil pressure. This alternate presenta- tion is used to advise the pilot of the BMEP that his engine is developing so that he may know its relation to the operating limitations. It should be noted that when the instrument dial is calibrated in terms of BMEP, the constant K will differ from that defined by Eq. 2:1. MAXIMUM POWER LIMIT RESTRICTED REGION FULL THROTTLE, CONSTANT R.P. M. LINES BRAKE HORSEPOWER 2,600 R.P.M 2,400 R.P.M. 2,200 R.P.M. 42 i 2,000 R.P.M. 40 FULL THROTTLE, CONSTANT MAP LINES 38 MANIFOLD PRESSURE, INCHES OF HG. 32 30 28 26 60 10 20 30 40 50 ALTITUDE. THOUSANDS OF FEET Fig. 2:2 Altitude Calibration Curve 2:4 given engine can be determined from them for given values of RPM and manifold pres- sure. A full throttle line obtained by connect- ing the end points of the RPM curves is also shown on the chart. As with other instruments, care must be exercised to insure that the torquemeter func- tions properly. In addition to the normal in- strument problems of lag, hysteresis, needle fluctuation, etc., the torquemeter readings are subject to error at low ambient temper- atures because of the congealing of oil in the line from the engine to the instrument panel. This problem is usually overcome by incorporating an oil bypass line with a valve in the system which is left open except when readings are actually taken. The altitude calibration chart shown in Fig. 2:2 can be: (1) plotted from equations or (2) obtained from actual data determined by flight tests or simulated flight test con- ditions. The latter scheme is, of course, preferable; however, the equations which are used are based on data which have been collected over a period of years using many reciprocating engines, and the results ob- tained are reasonably accurate, As stated previously, when an engine is not equipped with a torquemeter, the power developed by the engine must be obtained from a power curve as prepared by either the engine manufacturer or other engine test laboratory. The use of power charts, although not really an accurate procedure, does give a fair indication of the power being delivered (usually within 5%). The equation which relates horsepower at altitude to power at sea level is P BHP : ВнР. sl . [.3245 -0.1324] Po 2:3 The reciprocating engine performance charts consist of two plots, a sea level cal- ibration chart (or some other fixed altitude chart) and an altitude calibration chart. The former is always obtained by actual test; the latter may be calculated by certain equations or may be determined from actual or simu- lated flight test conditions. The altitude scale of Fig. 2:2 is usually based on a linear density ratio scale in order that the power lines may be straight. The solid lines of the figure represent full throttle operation for a given supercharger condition. The two charts when combined present the whole performance story on a specific engine for any type of operating conditions. Therefore, in addition to being used for power determination, these charts can be used to find the potentialities and the limitations of a given engine. When the data of the sea level calibration curves are combined with the data of the al- titude calibration curves, the combined charts are presented as in Fig.2:3. The reader is referred to Ref. 1 for a detailed discussion on the construction and use of such perform- ance charts, For the sea level calibration tests, the engine is connected to a dynamometer or some similar device whereby the power can be measured. Runs are made at constant values of RPM with varying manifold pressure to provide charts such as illustrated in Fig. 2:1. Although engine performance charts are not now used to the extent they were, they still serve a useful purpose in that they al- ways provide a check on torquemeter read- ings, and furthermore, they give the perform- ance characteristics of a specific engine for any combination of RPM, manifold pressure, , and altitude conditions. From these charts a pilot can determine what performance he may expect from his engine and therefore achieve the best settings for a specific flight. Such charts are applicable only for stand- ard sea level conditions. The power of a 2:5 2:3 THEORY OF RECIPROCATING ENGINE POWER CORRECTIONS If we assume that all terms are essen- tially constant for a given engine, with the exception of the air density and volumetric efficiency, we obtain the equation BHP : constant (pm)(e). The theory of corrections is based on certain simplifying assumptions, which in most cases are logical. Although some as- sumptions are made that are not exactly correct, the overall result is usually within allowable flight test limits. The theory is based on the fact that BHP is proportional to the amount of fuel burned; i.e., Since we do not normally measure the man- ifold density, we must express the BHP in terms of quantities we do measure. Ref. 2 shows that the BHP is closely approximated by the relation BHP : Ga GO (6-) . (H.V.)ny const. 2:4 BHP : const pm Vio Expressing the airflow through the engine in terms of displacement, RPM, manifold air density, and volumetric efficiency, where Ta = ambient air temperature = - Pm = manifold pressure. Go - (Displacement)(N)(e)lem). : 2:5 Therefore, the correction equation can be . 46 44 +42 B140 FULL THROTTLE, CONSTANT MAP LINES FULL THROTTLE, CONSTANT R.P.M. LINES 2,400 R.P.M. BRAKE HORSEPOWER BRAKE HORSE POWER | 2,200 R.P.M. 2,400 R.P.M. 12,600 R.P.M. 2,200 R.P.M. 2,000 R.P.M. 38. 36 1,800, R.P.MA 1 34. 320- 1,800 R.P.M. 30 2850 261 24 22 10 20 30 40 50 24 28 32 36 40 44 MANIFOLD PRESSURE, IN INCHES OF HG. ALTITUDE, THOUSANDS OF FEET Fig. 2:3 Combined Sea Level - Altitude Calibration Curves 2:6 written as (b) Outside Air Temperature CAT' MAP OAT' BHP BHP - MAP MAP' . CAT MAP ОАТ 2:6 when CAT - Carburetor Air Temperature OAT : Outside Air Temperature A correction must be applied if the outside air temperature is not standard for the par- ticular altitude. The brake horsepower var- ies about 1% for a change of 6 degrees cent- igrade (or 10 degrees Fahrenheit). A colder temperature means that the air is more dense. Consequently, more power is developed at lower ambient temperatures. The formula for correcting to standard conditions for a deviation of outside air temperature for a given manifold pressure is obtained indirectly from Eq. 2:6 as 8 and the primes represent test conditions and the terms without superscript represent standard conditions. Eq. 2:6 is the relation- ship normally used for power corrections. HP OAT' 2:4 CORRECTIONS APPLICABLE TO POWER CHARTS HP ОАТ 2:7 (c) Manifold Air Pressure Engine performance charts are always drawn for standard altitude conditions without ram. Therefore, if a given engine is opera- ted under nonstandard conditions or operated with the benefit of ram pressure, certain corrections must be made to adjust the engine performance to correspond to the perform- ance charts applicable to the engine, Sometimes a given problem requires the computation of the manifold pressure neces- sary to produce a given brake horsepower un- der standard conditions. In such cases, the observed brake horsepower reading is as- sumed invariant and we correct the manifold pressure by applying a correction on outside air temperature. Since manifold pressure is a direct function of horsepower, its cor- rection equation appears as (a) Outside Air Pressure or Altimeter Reading OAT MAP MAP . OAT' No correction need be made to the altitude reading (except for instrument and position errors) provided that the altimeter has been set to 29.92" Hg. and that pressure altitude is to be used as the reference for all correc- tions. 2:8 (d) Ram Pressure There is available a choice in procedure. For example, suppose we have data on alti- tude, OAT, MAP, and BHP. The question arises, "Should we correct altitude and MAP to give us the same BHP that we would obtain on a standard day or should we correct the MAP and BHP to values which we would have obtained at the observed altitude reading on a standard day?" The usual practice is to use the latter scheme; therefore, we start our corrections on the basis of the altimeter read- ing which is a measure of ambient pressure. The basic performance charts are com- puted for the condition of zero ram and it is, therefore, necessary to make allowances for these effects of ram pressure on engine oper - ation, During part throttle operation, the amount of ram pressure which may exist cannot be obtained from the performance chart. It must actually be measured. This . is true because for part throttle operation the performance charts present data on RPM, MAP, BHP, and altitude, but not throttle position. When an engine installation provides positive ram pressure, the same performance 2:7 condition may be obtained as for an engine without ram, but at a lower throttle setting. because all we do is specify some reference state at which we make our corrections. (b) Correction of BHP to the Observed MAP For full throttle operation, the story is quite different because the throttle position is fixed (full); therefore an engine which oper - ates with ram will be able to operate at altitudes and powers beyond those predicted by the performance charts. These differences can readily be evaluated from the flight data and the perforinance chart data. It is there- fore possible to obtain the amount of ram using a performance chart for full throttle opera- tion. Ram pressure is usually expressed in feet of altitude, This type of correction answers the fol- lowing question: "What power will I obtain on a standard day for the same MAP that I observe today?" Since in this method we are correcting to the observed manifold pres- sure, we automatically state the following: MAP: MAP'. * 2.5. TORQUEMETER POWER CORRECTIONS • Torquemeter power corrections are usu- ally divided into two categories, namely, part throttle and full throttle. Both methods are based however, on Eq. 2:6, the difference being that in part throttle corrections a simplification is made by making the ratio MAP/MAP or BHP/BHP' equal to unity. This simplification is realistic in part throttle corrections because the engine can be op- erated at the same MAP or BUP on a test day and a standard day. When this relation is substituted into our basic correction equation 2:6, we obtain the following expression for the corrected BHP: OAT' CAT' BHP = BHP' BHP' ОАТ CAT 2:9 If it is desired to correct the carburetor air temperature to the standard day condition, this can be accomplished by the following relationship: CAT = CAT' + OAT -OAT ' 2:10 Correction of MAP to the Observed (c) BHP (a) Part Throttle Corrections (Any Type Supercharger) This type of correction answers the follow- ing question: "What MAP would I need on a standard day to obtain the observed BHP that I read today?" For this particular type of correction we state that the standard BHP is equal to the observed BHP, and write Whenever performance data are obtained at part throttle operation, we do one of two things, namely: (1) correct the BHP to stand- ard conditions at the observed value of mani- fold pressure; or (2) correct the manifold pressure to standard conditions at the ob- served value of BHP. Thus, we either spec - ify the manifold pressure or the BHP, to be the standard value. BHP : BHP'. - We may then correct the manifold pressure by inserting the above into the basic correc- tion Eq. 2:6 to obtain ОАТ CAT In addition to this, the normal practice is to make all corrections at the observed pressure altitude. Consequently, we also specify the altitude to be a standard value. No errors are introduced when doing this MAP = MAP ✓ = MAP OAT' CAT' 2:11 2:8 2:6 POWER MEASUREMENT - COMPOUND ENGINES Again, the carburetor air temperature can be corrected to standard condition as shown before by Eq. 2:10. It should be noted that using the above procedure, we always make our correction at the observed value of the pressure altitude. This is the simplest way of performing power corrections. (d) Full Throttle Corrections The full throttle corrections are more difficult to make because it is necessary to correct both MAP and BHP for a given set of flight conditions, We saw earlier that for part throttle operation we needed to correct only one variable, but for full throttle opera- tion it is necessary to correct both MAP and BHP. We must do this because, when opera- ting at full throttle, the only thing that we can duplicate from day to day in actual flight testing is the throttle position, namely full. For example, if one flies at full throttle and desires to know what MAP and what BHP would be obtained on a standard day, it would be necessary to correct both quantities to standard conditions. Then, theoretically, if one flew these settings on a standard day, all data would coincide exactly. The compound engine is merely a modi- fication of the basic reciprocating engine in that it is equipped with exhaust gas turbines designed to extract power from the exhaust gases and feed this power back to the crank - shaft. The compound engines utilized by the U.S.Navy incorporate three exhaust blowdown turbines which are connected to the engine crankshaft by means of a fluid coupling. Since there is no definite speed ratio be- tween the turbines and engine crankshaft, the use of engine power charts is even less re- liable for the compound engine; therefore, practice thus far has dictated the use of a torquemeter to determine compound-engine power. The application of the torquemeter to power measurement of the compound en- gine is the same as illustrated in previous sections; thus, the procedures will not be repeated here. 2:7 JET THRUST MEASUREMENT - INTRODUCTORY COMMENTS If one corrected only BHP, however, and assumed the MAP to be the standard value, conceivably part throttle conditions would be represented for a standard day. Thus, it is evident that since operationally we can only duplicate the throttle position for full throttle operation, we must make corrections to both MAP and BHP. The acceptance test of a jet engine should include the measurement of thrust, fuel con- sumption and tailpipe temperature which all are functions of the airplane speed, the alti- tude and the engine RPM. Of this group of characteristics, only the thrust cannot be measured in flight using the same techniques employed on the ground test stand, It is possible to measure the inflight thrust in several different ways. Two methods presently used in France which give close agreement between their results are de- scribed here: the first is based on a study of the flow in the engine aft of the turbine, and the second on a study of the airplane perform- ance in a climb. We shall consider first the principles on which the methods are based and the cases to which they may be applied. Subsequently, the calculations required to reduce the flight data In addition to correcting these for full throttle operation, we must also take into account the ram pressure which is available to the engine and the number of supercharger stages. Thus, the full throttle power correc- tions become much more complex than the part throttle corrections. The procedure for accomplishing full throttle corrections is contained in Chapter 6. 2:9 to terms of thrust are considered. 2:9 THE JET FLOW MEASUREMENT METHOD 2:8 If FG = Gross thrust DISCUSSION AND REGION OF APPLICATION OF THE METHODS OF JET THRUST MEASUREMENT Qe Il Mass flow through tailpipe (a) The Method of Jet Flow Measurement Vex = Exit velocity Ae Exit area This method is based on the fact that the thrust delivered by a jet engine depends on the nature of the entering and exhaust flow through Pe = Exit static pressure the engine. p - - Ambient static pressure Q. Entering mass flow The net thrust FN of a jet engine is deter- mined by the tailpipe nozzle pressure ratio, the total tailpipe temperature Ttt , the amb- ient static pressure p the nozzle area Ae and the true flight speed V. V True flight path speed and FN = = Net thrust, The method of measuring engine thrust in terms of the aforementioned parameters is rapid and simple to use; however, initial ground thrust stand calibrations are required to correlate the theoretical equations with the actual operating conditions. It is desirable to verify the calibration coefficients de- termined on the ground by an alternate in flight measuring technique, we have the customary gross and net thrust definitions given below: FG = Qe Vex + Ae (pe -p) ( 2:12 FN = FG - QV. 2:13 For subcritical or unchoked flow, Pe = p. (b) The Climb Performance Method For sonic (choked) conditions at the exit of the tail cone, the exit pressure Pe is re- lated to the total tailpipe pressure Ptt , by the equation, Pe 2 Y/17-1) +62) . This method is particularly applicable to interceptor-type aircraft having high rates of climb at maximum or climbing RPM. It is only necessary to measure the classic air- plane parameters: time t, static pressure p, ambient air temperature T, true airspeed V, and the fuel consumption. From these we may determine the true and energy heights, the actual and energy climb rates, and the airplane weight at any instant. Ptt +1 2:14 where Ptt, the total tailpipe pressure, may be measured at any point in the tailpipe, if we assume isentropic flow between the tur- bine outlet and the exit section, Similarly for choked flow, we have the temperature relation The climb is made under the conditions specified in Chapter 7 for the maximum ener- gy climb. As we shall see later, this proce- dure almost entirely eliminates the influenc of engine flow on drag, which later inter- action factor may cause errors in perform- ance determinations as noted in section 2:1. 2 Te T +7 Y+1 2:15 2:10 where Te = exit static temperature Tu = tailpipe total temperature. Editor's insert: In the United States a direct reading gross thrustmeter has been developed based on Eq. 2:19, expressed in a modified form. To obtain the gross thrust - meter relation for unchoked flow, we note that Eq. 2:19 may be written FG 2y Ptolly-lly -1)/y If the exit conditions are subsonic (un- choked), the exit velocity is given theoret- ically by the isentropic relation ( -. p 2 YR (- 1)/Y Ae Yol ༢༣༨ . 2:20 Vex - [ (697) vor] Y-1 = and introducing 8 = p/po; where po = stand- ard sea level static pressure, we obtain 2:16 FG 2y For sonic exit conditions ( 0-2) - 8 AE Y-1 2:21 Vex : ✓YRTE . 2:17 from which we see that the quantity FG/Ae is a function only of the pressure ratio Pu/p and does not depend directly on either the ambient pressure or the total tailpipe pres- sure, but only on their ratio. The exit mass flow rate differs from the entering flow rate because of the addition of fuel. Normally, the fuel weight is ignored, since it is a small quantity in comparison to the total mass of air passed through the engine in a given time. Under these circum- stances, Qe and Q may be assumed equal, both being given by the relation Pe Q = pe Alvex - Ae Vex 2:18 Moreover, Eq. 2:21 is identical in form to the Machmeter calibration formula except for a radical. Therefore, a Machmeter connected on one side to the total tailpipe pressure pick- up and on the other to the static source may be calibrated to read in terms of the para- meter FG/Aed. RTe 1 For the unchoked nozzle 1 Pro \y-1)/y ly .)y T: (+1 :) Pe :p and Te Actually, it has been found that a simple differential pressure gage measuring the dif- ference between Plt and p may be directly calibrated to read gross thrust in pounds with the calibration being precise at sea level and only slightly in error at altitude. For precise work at altitude, simple corrections may be applied. and the gross thrust equation becomes FG QVE ZYR . -TT х x Ае Ае Y-1 For sonic flow at the exit, we have FG Pe (y-1)/y : V V? P + (De-o) lex RTE 11/1-G2-20%* A)? 10K (0)" Ae RT17 2:22 or FG 2y р 외 ​Pit and, using Eq. 2:17, FG + De p = (y + 1)pe-p; Ае Y-1 ۲۹۱) = y Pe - 2:23 2:19 Ae 2:11 then from Eq. 2:14 and FG DIE 2 FG Y/17-1) -D = ly+1) P17 lyt - КРО 1.26 for choked flow Ae Y+1 Ae8 D 2:28 2 ylly-1) р PAT ( y +1) where K = f(Pt/p) in both cases. f For gross thrustmeter purposes, Eq.2:24 may be written ylly-1). F G Рft of( +1) 2 Y +1 -1 Ae р For choked flow, K is essentially constant with a value of between 0.95 and 1.00. For subcritical flow this coefficient decreases gradually as Plt/p decreases. We note that the pressure Pt, should be determined by some sort of averaging total head tube which ad- equately samples the exhaust gases. Tocal- culate the net thrust it is necessary to deter- mine the mass flow rate Q. The equation for mass flow rate depends on the nature of the flow (subcritical or choked). 2:25 Since the tailpipe flow is at an elevated temperature, y~1.33 and FG L. 1.26 SA For subcritical flow, Eq. 2:18 gives 2:26 pe . Ае IRT e 2 YR р For normal nozzles, the flow becomes choked at Plt/p~1.85; therefore, Eq. 2:21 applies to lower ratios and Eq. 2:26 to higher ratios. -1/ T 1,16) =117. Y-! Pit and p = Pe, 27 Q Since Te = Ttt (p/Pr)(y-1lly PRE /т, = Х For practical purposes it must be realized that isentropic, one-dimensional flow con- ditions (as assumed in the derivation) do not actually exist except on the average, and therefore, an efficiency factor should be ap- plied to all equations. This factor must be determined by calibration procedures. Ае (Y-IR р 217 (y+1)/y Il ( 4 ) 67.82*** ۲۹۱ Dat 2:29 Because the real nature of the flow is a function of the pressure ratio, the K factor is also a function of the pressure ratio and should be determined accordingly. or, alternately, Q р 27 Х AD VTT ly-IR Thus, we may write F G 2y Pylly-illy = Ae8 * (34). [- .)(x1 Pt 12ly-llly Prilly 9-17 -1/y. 2:27 for subcritical flow 2:30 2:12 Eq. 2:30 may be written and for choked flow, Q TT 2Y Х outte SA e (-1)R - SA PTH2ly-1)/y Ptt Pfly-1)/y 0:2243 6) -Y КQPo () VGA Y 2 \6y+1)/(y-1) /1 R V+ -)(x+ p 2:31 2:35 For choked flow, Eq. 2:18 gives Pe y ola YRTE - Pe Ае RTe RTe and using Eqs. 2:14 and 2:15, +1)/y-1) ola Sen +264) where KQ = f(Pty/p). We might compare this method of deter- mining mass flow in terms of exit conditions to the possibility of measuring conditions at the intake duct where the temperature dis- tribution is far more uniform than in the tailpipe.* In this latter case, the nature of the entrance flow and the pressure distribu- tion therein can be markedly influenced by changing flight conditions and possible flow separation, particularly when split ducts are employed. Moreover, the calibration of the entrance cone is a laborious proposition, whereas the calibration of the tail cone is quite conveniently accomplished. (67)671 R yt 2:32 or, in another form, ($)/,(). ܙܙ*/ܘ Q art PIA 2 2 \r+1)/(x-1) R+ 1(+1(. SAe 2:33 We note finally that the term (V in level flight (at subsonic speeds) has a value about 30% as great as the gross thrust, so that greater errors are permissible in the deter- mination of the ram drag than in the gross thrust. 2:10 THE CLIMB PERFORMANCE METHOD Since the mass flow rate depends on the temperature as well as on the nozzle pres- sure, an additional source of error is intro- duced here since one must properly measure the total tailpipe temperature. To account for the fact that the flow is not actually of the isentropic one-dimensional type and for the difficulty of accurately sampling pres- sures and temperatures, a nozzle discharge factor should be added to Eqs. 2:29, 2:30, 2:32 and 2:33, whence for unchoked flow, The most severe criticism which can be made of the preceding method is the fact that its accuracy depends on the proper deter- mination of the coefficient K and KQ, which must be established by ground calibrations. Even using nozzles more convergent than the flight tailpipes, it is not possible on the ground to obtain the pressure ratios, Plt/p, encountered at altitude. Also in the region where the curves of K and KQ as functions 27 KOPO х Ae ly-I)R PT *+y 2ly-1)/y Pilly-1)/Y - 0) * See also Eq. 2:46 for determination of mass flow rate from entrance conditions. 2:34 2:13 of the ratio Puc/p do not require extrapolation, there is no guarantee that the ground caiibra- tion will hold al all altitudes. the same values at other altitudes, it is suf- ficient to determine the relationship between FN, and FN at some higher altitude, first using the "jet method" and then using the “performance method”, which directly gives the ratio FN/FN. Attempts have been made to correlate the "jet method" of thrust determination with data obtained from accelerated and decel- erated level flights (with corrections imposed for slight inadvertent climb or descent rates). These tests were based on the fact that the net thrust equals the drag plus the inertia force. Thus, this latter method compares the thrust obtained at the same engine RPM un- der the same conditions of climb, but at dif- ferent altitudes. It essentially eliminates the drag variations which might be produced by changes of entrance duct flow. We shall sce also that the method allows us to deter- mine the increase in thrust due to after- burning. Two flight conditions were considered: the first one at high RPM (for which we wish to measure the thrust), and the second at a reduced RPM (nearly closed throttle) and at the same true speed as the first. The difference between the two values of accel- eration at the same airplane speed gives the variation between the thrust at the reduced RPM and at high RPM. The reduced RPM thrust can be considered as a corrective term, and may be determined, for example, from the "jet method" because even a larger error in this term introduces only a small error in the thrust measured at high RPM. In the following work, the subscript o rep- resents low altitude and symbols without subscripts refer to high altitudes. Drep- resents the aerodynamic drag; w, the rate of change of energy height; m, the airplane mass, and g, the gravitational acceleration. From the flight equations, we have (FN-DV = mgw. It follows that FN-D < 313 313 $TS FNO- vo 2:36 The procedure described above has not provided satisfactory results and it is be- lieved that the major source of error is the “a priori' assumption that at a given alti- tude and airplane speed, the airplane drag is independent of engine RPM. Wind tunnel model tests have shown that variations of CD as great as 25% may be produced by variations of flow rate through the intake, and for this reason the outlined procedure was abandoned. from which FN W FNO mo Wo la D W Do + D。 mo wo vo Wo FNO 2:37 We now seek a simple method which per- mits us to establish that the coefficients K and KQ are functions only of pressure ratio and not of altitude itself. We may assume that the values of K and KQ, established by ground tests, are correct when the airplane is at its best climbing speed at low altitude, and use the "jet method" to compute the net thrust FN, under these conditions. To establish that K and KQ have During a climb, the first term of Eq. 2:37 decreases. As measured during recent tests on an interceptor at the Flight Test Center (France), the first term had an initial value of 1 at sea level; 0.6 at near 20,000 ft.; 0.4 at 30,000 ft.; and 0.2 at 40,000 ft. The second term, on the other hand, varied from 0 at the ground to 0.04 at 20,000 ft.; 0.09 at 2:14 30,000 ft.; and 0.1 at 40,000 ft. Thus, the second term may be considered as a correc- tion term at altitudes less than 30,000 feet for the case considered. In practice, at altitudes where the ratio w/wo exceeds 0.5, one may consider the second term as of second order in comparison to the first; therefore, this term can be com- puted from wind tunnel data for the particular airplane without introducing serious errors, Knowing the airplane weight and the airspeed, we thus compute D and Do. and some prior knowledge of propeller ef- ficiencies under the encountered range of operating conditions. Methods of measuring jet thrust, and useful propeller thrust are discussed in other sections of this chapter. In this section these two measurements are considered in relation to the turbo-propeller engined aircraft, by investigating the ac- curacy to which net jet thrust and shaft power should be measured, and thence discussing current and possible methods available for making these measurements. Accuracy Requirements As for the thrust FNo, this may be deter- mined by the "jet method", since at low al- titudes the sea level calibrations cannot be far from correct. Aircraft performance characteristics must be defined according to some standard of accuracy. In the United Kingdom the standard sought after is that the level speed performance should be defined within 1%. * In applying this method through altitude intervals during which w/wo remains greater than 0.5, the results have been found to be in excellent agreement with the "jet method”, less than 3% variation having been found among all tests. This procedure also allows us to deter- mine the increase in thrust due to afterburner operation. In this case, we may still apply Eq. 2:37 but with subscript o referring to the engine alone and the absence of a subscript to conditions with the afterburner operating. If similar climb schedules are chosen, the first term of Eq. 2:37 remains the important one over a large altitude range, so that the use of wind tunnel data for determination of the second term is permissible, notwithstand- ing the possible variations in drag produced by afterburner operation. At first glance, it would appear possible to achieve this without considering the ac- curacy of determination of engine power and thrust, merely relying upon the accuracy of the airspeed instrumentation. It is, however, usually quite impracticable to measure the aircraft performance under standard condi- tions, hence some form of performance re- duction must be used (the exact form is im- material) and for this, engine powers and thrusts (or their non-dimensional equiva- lents) must be measured. Also, if airframe drag data are determined from power and thrust measurements, these must be meas- ured with an accuracy equivalent to that specified for the level speeds. 2:11 THE MEASUREMENT OF THE USE- FUL THRUST OF TURBOPROPELLER ENGINES Some distinction can be made here be- tween random and systematic errors; for instance, if the random errors are within the limits to be prescribed, considerably larger systematic errors in engine meas- urements can be tolerated if airframe drag is not required. However, inasmuch as sys- tematic errors are usually easier to detect and to make allowance for than random er- rors, this distinction is at best academic, As with other types of power plants, in- formation on airframe drag and aircraft performance under standard conditions is likely to be required for turbo-propeller engined aircraft. This necessitates meas- urement of net jet thrust, engine shaft power, * 95% probability of a single observation. 2:15 and systematic and random errors are, there- fore, usually grouped together. stallation to installation of a particular en- gine type. By differentiating the performance equa- tion in an identical manner to that adopted in the development of differential perform- ance reduction methods, equations can be developed similar to Eqs. 8 and 12 of Ref. 3, from which the accuracy requirement for engine measurements to meet the specified accuracy in performance definition can be determined. Using this method, and insert- ing experimentally determined values of the coefficients of Vi/Vi, AP/P, etc., in these equations, the results shown in the table below were obtained for a particular installation in 1953. The climb data have been included to il- lustrate their more stringent accuracy re- quirements which tend to zero error as the ceiling is approached. However, since climbs are equally dependent for consistency on pilot technique and atmospheric conditions, it is not usual to base instrumentation ac- curacy requirements on them. Basing the requirements on level speed, it would appear, therefore, that 1.5% for power and 12.5% for thrust could be safely adopted as individual contributions to errors in airspeed. Grouped together with other sources of error, such as air temperature and pressure, it is necessary to reduce these values still further. This can be done if it is assumed that all errors are random, and 0.5% error in speed is allowed for each of the seven possible variants, namely: air temperature, static pressure, dynamic pressure, aircraft weight, jet thrust, engine torque, and engine speed. This results in an overall random error of 1.32% in speed on a 95% probability basis (if this is the standard adopted for defining the accuracies of measurement), and enables final and fairly realistic determination of rounded-off limits of accuracy for engine speed, engine torque and jet thrust to be quoted, respectively 0.75% 0.75% and 6% The level speed values hold for an engine where the jet thrust power is about 10% of the total thrust power, which is fairly rep- resentative of current engines. Reduction in this ratio (which is the direction in which engine development should proceed) will re- sult in a slightly greater accuracy being re- quired for the engine power measurement, and a relatively greater reduction in the accuracy required for the thrust measure- ment. The climb values depend as much upon the ratio of total thrust power to drag power at the best climbing speed, as upon the ratio of jet thrust power to total thrust power. They may, therefore, vary from in- PERFORMANCE ITEM X ENGINE PARAMETER Y Accuracy Required of Instrumentation Percent Per 1% in X X AY Y AX Near Sea Level Near Useful Ceiling (登​·錢​) Level Shaft power 2.5 1.4 Speed Jet thrust 25.0 12.5 Climb Shaft power 0.55 0.28 Climb Jet thrust 5.0 2.6 2:16 2:12 MEASUREMENT OF SHAFT HORSE- POWER advisable to consider whether in fact the pressure gage records this, In practice, both absolute and differential pressure gages are used, and the latter may be vented to aircraft static or cabin pressure. This is achieved, as in the piston engine, by separate measurements of torque and en- gine speed. Engine speed indicators meet- ing the accuracy requirement of 0.75% can usually be selected from equipment which is generally available. Allowances may have to be made for the differences in back pressure which occur between ground calibration, and in flight, both for the instrument and the torquemeter piston. All ambiguity and unnecessary calculation can be obviated if a differential pressure gage is used to measure the direct differential across the torqucmeter pistons, both during the ground calibration and similarly in flight. The usual torquemeter on turbo-propeller engines is connected to the propeller epicyclic reduction gear train, in which the movement of an annular gear (which would be fixed in the absence of a torquemeter) is opposed by oil pressure acting on a piston and connecting rod assembly. By a system of ports in the cylinders, the annular gear is constrained to remain, within limits, in a predetermined position, so that the geometry of the linkage does not change, and the oil pressure nec- essary to maintain this position is taken as a measure of the torque. 2:13 THE ESTIMATION OF JET THRUST Because of the relatively low standard of accuracy required in the estimation of net jet thrust for a turbo-propeller engine, it is not necessary to adopt such precise meth- ods as when dealing with the turbo-jet engine, The usual approximation is to assume the effective area remains constant at the value for the cold nozzle. Shortcomings encountered in torque meas- urement have been: (1) Layshaft type. Change in cali - bration with oil temperature. (2) Epicyclic type. Excessive lag, and change in torquemeter constant with torque, in epicyclic versions where the con- necting rods make an appreciable angle with lines of action of the pistons. There are little data available to suggest what error is involved here, because to date little interest has been shown in establishing effective nozzle areas for these engines. The error will depend upon the jet pipe configura- tion and the location of the tailpipe pitot head, and in general any carefully placed pitot should sample the mean total head to within 10%. This has been shown to be true by rake measurements on at least one engine, which produced discharge coefficients ranging be- tween 1.04 and 0.92 respectively at the low- est and highest pressure ratios achieved in a ground run. (3) All oil pressure types. It is necessary to consider whether the indicator is a differential or absolute pressure gage. Items (1) and (2) can only be rectified in the design stage. It is advisable, however, to consider each installation on its own merits, and to arrange the program of ground calibration of the torquemeter so as to as- certain whether these effects are significant. The measurement of intake momentum should present no additional difficulty, be- cause similar assumptions with regard to the effective area can also be made here, with an even more generous latitude in general in the measurement of jet pipe total tem- perature. Item (3) involves the consideration that the torque is balanced by the differential pres- sure across the torquemeter piston, and it is 2:17 It will be observed that the final nozzle pressure ratios on turbo-propeller engines are much lower than those for turbo-jet en- gines, and differential pressure gages in- tended for measuring the total head should be of suitable range, a full scale value of 2 psi sufficing for all current engines, coefficients which are used to describe ram- jet performance. Later, we shall investigate how these coefficients may be determined by flight tests; however, we must postulate at the beginning that certain of the engine char- acteristics should first be determined by wind tunnel tests wherein the effects of com- bustion are artificially simulated. (Under these conditions the wind tunnel model pro- duces some thrust and the mass flow through the jet simulates the mass flow encountered in actual flight.) As a final word of warning, turbo-propeller engine jet pipes are frequently not parallel to the longitudinal axis of the aircraft, and allowances should be made for this where necessary. 2:14 CONCLUDING REMARKS ON TURBO- PROPELLER THRUST DETERMINA- TION The wind tunnel data may then be com- pared and correlated with flight test data. These data should establish whether or not the assumptions of the analysis (one-dimen- sional flow all passing through the engine without spillage) are appropriate. Consider any thrust-producing device as illustrated in Fig. 2:4, Assuming that Po and P4 are equal, we have EN : QIVA - Vol FN 2:38* where Q = mass flow per unit time Vo = free stream velocity In the measurement of useful propulsive thrust of turbo-propeller engines, torque and shaft speed should be estimated to 0.75% of the total speed and net jet thrust should be estimated to 6% of the total thrust, if it is required to define aircraft forward speed to within an accuracy of 1%. Care is required in the design of torque- meters and in the selection of engine speed indicators, if they are to meet this require- ment. The estimation of jet thrust within a tolerance of 6% implies that in all normal installations, a rake calibration of the single pitot installation is unnecessary, and that the effective area associated with the final nozzle pressure ratios deduced from the single tail- pipe pitot readings can be assumed to be the same as the cold nozzle area. It is advisable to examine the torquemeter installation and its pressure gage, if it be hydraulic, in order to ensure that due allow - ance is made for possible effects of altitude and cabin pressure on the torquemeter cali- bration. The torquemeter calibration should be arranged to investigate lag and tempera- ture effects. V4 = ultimate wake velocity Eq. 2:38, of course, is the classic second law of Newton that force is equal to rate of change of momentum. If p represents gas density and A rep- resents area at a given station, then Q = PAYPAYA 4 4 4 2:39 оо so that from Eq. 2:38 FN v2 Р A 444 -Po A 2:40 2:15 RAMJET THRUST MEASUREMENT We shall first examine the nature of the * Neglecting the mass of the fuel. 2:18 NA Fig. 2:4 A4 A3 Ао AI A2 ( Section Aft Of Combustion Area ) 2:19 or air passing through the engine has been sub- jected (compression, combustion, expansion). Povo FN 2(KA4 – A 2 Measurement of CT in Flight 2:41 where Since the areas A3 and A1 are known, it is only necessary to measure the pressures p3 and pi to determine the areas A4 and Ao. 2 PAVA 4 M2 4 K = (since pa po Pol -D 2 P va M ° 2:42 It is easy to show that if we know the pressures existing at two stations such as A2 and A3, the ultimate wake Mach number, M4, is determined.* If only compression and expansion were involved, K could theoreti- cally be one, but for any possible combustion process, K is less than 1. Then, knowing Mo, with M - Mach number. If S is the wing area of the airplane powered by the ramjet, then we may write ma 4 ov K = FN : CT S 2 2:43 may be found. where Cris the ramjet thrust coefficient. Comparing Eqs. 2:41 and 2:43, we have СТ . 2 S 를 ​(KA4 - Ao). ) 2:44 Eq. 2:44 is quite general, applying not only to ramjets but to any propulsive system. For a turbojet, K is greater than 1. Moreover, the quantities Ao and A4 vary with speed so that Eq. 2:44 is not too useful when applied to this type of engine. On the other hand, when applied to a ramjet with a given relative heating parameter T2-Ti A = TI Editor's insert: It was not clear from Mr. LeDuc's paper in just what fashion the areas A4 and Ag and the ultimate wake Mach number are to be determined. Indeed, there seem to be several possible approaches to this problem depending on whether the free stream conditions are subsonic or super- sonic and on what assumptions are made in the analysis. Presumably the laws of con- servation of momentum and mass are applied to give: + pove) AO = 10, + Piva; (momentum) IPO 2 PAYO A? V ve (mass) 2 2 PA from which where T = absolute total temperature, it is found that the parameters in Eq. 2:44 are essentially independent of speed. Thus, the thrust coefficient CT is of prime importance in the study of ramjets. PA?v pode A (Po + povola. (61 + AL PA 2:45 In an ideal ramjet without losses, Kis equal to 1. The actual value of K, which is of course less than 1, is a measure of the overall efficiency of the process to which the Assuming that combustion is completed within the engine. 2:20 from conventional aerodynamics For subsonic free stream conditions, we may assume isentropic compression between stations 0 and 1, in which case 2 M4 3)(y-1)/y 109) 03-17 Vy 2:47 roles Pi where Pt3 so that Eq. 2:+5 becomes = total pressure at station 3, and Y has a value somewhat less than 1.4 (approximately 1.33). I + y MS y PY Aě - 6 month aga, eta Ро P.1 (Y+1)/ Y The static pressure at station 3 depends on whether the flow is choked or unchoked. For unchoked flow the exit pressure is equal to atmospheric whereas for choked flow, the pressure exceeds atmospheric by a given amount. PO O 2 Y у м. 2:16 CONCLUDING REMARKS ON RAMJET THRUST MEASUREMENT and consequently ty MO Ao гума Pilly - A, C (a) It is pointed out that the determination of Ma is a relatively delicate process. This is so because the equations presented so far are written for one-dimensional or average flow through the engine, and it is well known that the actual flow may differ considerably from the assumed one-dimensional type, particularly for nozzles having a large value of the ratio A3/A max 2 2 Vy I +YMS 2YME AL A, ( Pilly+1)/y Po emo 2:46 Flight tests conducted on two French ram- jets indicate that for one of these, the equa- tions seem to give the correct results where- as for the other, discrepancies of the order of 5% are found between flight and tunnel tests based on a comparison of Ct and Cp in steady level flight with CD obtained from tun- nel data. so that if we know Mo, Al, Pi and Po, then A. may be determined. For supersonic flow, the design of the entrance section would have to be considered and the nature of the entry shocks evaluated for a precise analysis. Eq. 2:46 may be written for the exit flow by substituting sub 4 for subo and sub 3 for sub 1. The value of the tailpipe total pres- sure may be used to evaluate the theoretical ultimate wake velocity. Assuming an isen- tropic expansion from station 3 to 4, we have (b) It is appropriate at this point to com- ment on the validity of Eqs. 2:41 and 2:44. It should be understood that these relations give the overall net thrust which, as is well known, may be produced by pressures acting both on the internal and external surfaces of the ramjet vehicle (particularly the external surface in the vicinity of the entrance sec- tion). 2:21 where Q - rate of gas mass flow If we consider only the thrust on the in- ternal walls, we have Fi QI V3 - V.) +103-P) A3 -10,-pola, Af . area at end of nozzle 2:48 P . ambient air pressure and if we write Pf gas pressure at end of nozzle AF : FN -Fi Vf Ve = exhaust velocity (relative to mo- tor) 2:49 The mass flow is given by the equation: it should be recognized that in some cases AF may comprise 15% of the total thrust, FN. Q : CoPc At 2:51 where At = throat area of nozzle Finally, we point out that the design of the pressure pickups at station 3 is very im- portant since they must provide a proper av- erage value of the static pressure (and of the total pressure if Eq. 2:47 is used) while at the same time be resistant to high tempera- tures. * Cp : discharge coefficient Pc : chamber pressure Combining (1) and (2), we have F: (Co.Ag.Vp) Pc + Pf. Ap-P.Ar 2:52 (Editor's note: Mr. LeDuc's original paper also included a section on fuel con- sumption characteristics of ramjet engines, which has been deleted here.) 2:17 MEASUREMENT OF ROCKET THRUST IN FLIGHT With supersonic flow in the divergent part of the nozzle, the internal flow will be in- dependent of external pressure. Also, the chamber temperature is independent of cham- ber pressure being determined by the chem - istry of the combustion process. It follows that for a given rocket, the term Pf Af will be proportional to the chamber pressure Pc and hence that € We will consider only cases in which the exhaust gases are under-expanded, the gas pressure at the end of the nozzle being no less than the pressure of the ambient air. This will usually be so, as over-expansion is uneconomical. The thrust of the rocket motor is given by the equation F BPc - Pfaf 2:53 F : QVp + (Pp - Polat where B is a constant of the particular de- sign and is determined from test runs on a ground rig. . 2:50 * It is, of course, also important to obtain a satisfactory average value of the intake static pressure Pi. It is customary to define the in-flight thrust as that which the rocket would give if the ambient air pressure at the end of the nozzle were equal to that of the undisturbed air. Any departure from this condition which 2:22 results from the presence of the airplane is classified as an interference effect in drag. This classification is arbitrary, but it has the substantial advantage that airframe ef- fects are attributed to the airframe and that the thrusts of the same motor in different installations are immediately comparative. consider several possible methods of meas- uring jet thrust and shall also consider several definitions used to describe jet thrust of an engine as installed in an air- frame. In the following work we assume that the engine axis is aligned with the air- flow direction so that there are no effects of pitch or yaw on the thrust developed. It is convenient to define the thrust of a jet engine in terms of a simple ducted body whose axis is parallel to the direction of the undisturbed motion, i.e., the flight path. Fig. 2:5 shows the internal flow ahead of, through, and downstream of such a ducted body, though the flow downstream is idealized in that no mixing of internal and external flow is shown. With assumptions of ideal gases and so on, it is possible to analyze the factor B in terms of nozzle geometry and relate it to more fundamental quantities such as the discharge coefficient. This, however, is hardly worthwhile inasmuch as the empiri- cal approach suggested above avoids unnec- essary idealizing assumptions. However, a theoretical treatment of nozzle flow for an ideal gas may be found in Ref. 4. > From momentum considerations the thrust between any two planes* perpendicular to the 2:18 GENERAL ANALYSIS OF JET THRUST MEASUREMENT *Such planes will be referred to as a "sta- tion" with a reference to its position, e. g., station i is the plane at the engine inlet, In this portion of the chapter we shall Station At Infinity Upstream Entry Station i Exit Station Position of Effective Area f Station At Infinity Downstream W Free-stream Direction Boundaries Of Pre-entry Stream Tubo Tinclination Of Stream- line To Free-stream Direction) Boundaries Of Equivalent Post-exit Stream Tube FIG. 2:5 DIAGRAMMATIC REPRESENTATION OF THE FLOW THROUGH A DUCTED BODY Fig. 2:5 Diagrammatic Representation of the Flow through a Ducted Body 2:23 the internal thrust direction of motion is given by (Refs. 5 and 6) the change between the two planes in the quantity Fint - 'pe u + Pe -Poo I dag ) Jude+fie-Polo (p ) DA. SMU? +P; - POI DA, U² i 2:54 2:58 and the post exit thrust Fpost تا ۔ w W Applying this to the thrust of an installed jet engine, the integration is limited to the internal flow, and the difference in the quantity must be taken between stations and w, where station is sufficiently far up- stream for the pressure of the internal flow to be that of the undisturbed stream, and the internal stream tube is therefore parallel to the direction of the undisturbed flow. Station w similarly is far enough downstream for the pressure to be uniformly that of the undisturbed stream. In en Pw up y daw -fil revé + Pet + Pe - Pool dhe 2:59 so that FN = Fpre + Fint + Fpost · 2:60 The thrust is then defined as The pre-entry thrust represents the force exerted in the upstream direction on the external flow by the pre-entry internal stream tube and this force acts on the engine instal- lation in the form of an external pressure mainly near the lips of the entry. F FN-Sen u dane U W vô 2 U А, 2:55 where FN is the net thrust and represents the net force of the internal flow on the body in the upstream direction. Also the gross thrust FG is defined by The internal thrust is the force exerted in the upstream direction by the internal flow on the internal surfaces of the air ducts and engine. The post-exit thrust is the force exerted in the upstream direction by the post-exit internal stream tube on the external flow, and this force acts on the engine installation in the form of an external pressure mainly near the jet exit. Fou Semua DA few dAw 2:56 (a) Practical Definitions of Thrust from which it follows that as a special case under static conditions, the gross thrust represents the force of the internal flow on the body in the upstream direction. The net thrust can be considered as the sum of three parts as follows: The definitions of net and gross thrust (Eqs. 2:55 and 2:56), while strictly correct, depend on integration over the internal flow at station w far downstream of the body. In a practical case such an integration cannot be made because the distance is inconven- iently far from the exit, and more important, due to mixing in the wakes it is not practical to separate the engine thrust and external drag at station w. We must therefore The pre-entry thrust Fpre = S, 12. ups + P PO IDA, - PO MO AO" (P. P. u² 1 2:57 2:24 Similarly, for choked flow attempt to relate the conditions at station w to those at or near the jet exit where meas- urements are possible. Practical definitions of thrust depend on the accuracy of such relationships in determination of the post- exit thrust. FSG Ро Pe Sac[y Lorena She hoe ] -1) June Y + cos Pa Ρω 2:63 The standard thrust is obtained by choos- ing the exit station e or "effective area" station f as the downstream reference sta- tion and by assuming that there is no contri- bution to the thrust from the flow between that station and station w far downstream (i.e., no post-exit thrust). The standard thrust is thus directly ap- plicable if the jet discharges into a region where the pressure is equal to Po and the post-exit thrust is therefore zero. It is the thrust normally specified in engine manu- facturers' brochures and may be quoted either as a gross or net standard thrust. If we make the customary assumption that for unchoked flow, the exit pressure is equal to the free-stream pressure, we find that for $=0°. Y-1 2Y Y dA Y-1 2:64 (Pe FSG Poo 5 [ {) $ -1} dae Ae which coincides with Eq. 2:20 except that average conditions have not been assumed. For $ = 0º, Eq. 2:63 for choked flow be- comes - FSN Sleeve +PC - Pe) dAQ - muco Sipe Fusce fe. [v+1 -1 Dome Pe FSG Doo So (Y+) Pe POO DA 2:61 and Y Pe An equivalent expression for this net thrust is obtained from thermodynamic con- siderations, in the following fashion: The differential standard gross thrust Pet ܪ ( 2 Y- (from Eq. 2:14) Fsg is . 2ں so that IDAE o FSG = l pe + Po-Polda Pe =lpe ve cos ide + Pe - Polda 2 Pet FSG - SA Sell ( 1.26 --] I dae POO 2:65 for Following the derivation of Eq. 2:20, but assuming that exit and free-stream condi- tions are different, we have 2Y 2 Y y Pe ve ² Y y = 1.33. Y-1 noveerimine :) ( er Pe --} Subtracting the entering rate of change of momentum, QUO gives the standard net thrust; thus from Eq. 2:62, where Y-1 Y Y Pet 27 2Y Pe Y- POO Poo Pe FSG 1937, tot ) dae force FsN" S[ 4 Pe{ [ (738-1 cos't } + -) POO Pe Pe -1 COS dhe-muco ०२ POO 2:62 2:66 2:25 (b) The Jones Thrust constant y between w and c 2 De Y y Pe Y1 The Y Y Pw Y + Y- Pe Y-1 PW ใช้ Y-1 hence 2 2 Vå avě + 2 los 27 + Y-1 - Powe and 27 onlen For convergent nozzles (subsonic or sonic velocity at exit) which exhaust at a local pressure which differs from Po the assump- tion can be made that the jet expands or con- tracts adiabatically and isentropically to the undisturbed pressure Poat station w without any transfer of energy or momentum through mixing in the wake. This model of the flow enables the conditions at station w to be calculated from those measured at station e, the jet exit. The thrust so ob- tained is designated the Jones thrust since it represents an extension of Sir Melvill Jones' original method of obtaining the drag of a body (Ref. 7). For these assumptions the conditions at station w can be calculated from those at station e using the condition of continuity, of mass flow between stations e and w, and the energy equation assuming isentropic flow. The resulting expression for the jet Jones thrust in terms of the measured quantities at station e is derived as follows: By definition, the gross Jones thrust is = 1 + ) (7-1) ve Pe Pe - but 2Y ve Pe Pe - pe > so that Pe Pw Pe Pw Por Pe 1 PW De oulsen W = 1 + : 1 + Pet Pe Pet Pe Pe 2 Per Per Pe PwYw Fuo San daw. JAW Now, since we have assumed the existence of isentropic flow at constant y throughout If we assume that angle of swirl y is neg- ligible and that isentropic flow without mixing exists between stations w and e, then conti- nuity provides* Ý (Pe Y- Pe 会​”: Pe D i فوای . De w let Pele cos de dae PW Yw daw or and 2 2 pv dAw Pe ve Vw cos de dae Ve dA . (Pw H Y1 7 Vw 22 Assuming isentropic flow conditions and Ver + २९५) Pe -1 since *If the swirl angle is not small and the jet diameter changes appreciably between sta- tions e and w, cos de must be replaced by cos de cos yw, where yw is obtained from We by the law of conservation of angular momentum. Y- H Pe ve = 21 Pe 24 2** et Pe Y-1 2:26 and FUG Sam Povo 66 If we presume a constant value of re- covery factor for the thermocouple tempera- ture pickup, then from Eq. 1:63, Vw 3 cos de dae Ten • Te + al Team we have Y- (Pet Il FUG'S. 44 po{ le :) -} F18 2Y Y-1 Pe and from Eq. 1:62 Pe 1/2 (2 Te m² = Tex - Te | - Pe. . cos de: {! + dae so that 'Pet H Pe : 2:67 Ter ; Te + el Ter - Te) = c Ter + (1-6) Te and and FUN= (Eq. 2:67) - QUOD 2:68 Teti her Te € + (1-€) ose Tex for unchoked flow. but Y-1 1 Pe If we assume that for the choked flow, we have isentropic conditions between e and w (which assumes the absence of shock out- side the engine) then, Pet Ž 2 17-18 ( Pe for isentropic flow, hence and Eq. 2:67 becomes Y-1 Tex Pe ket + (1-6) pet Fuo = Son 7 Pe cos de [ [1+ ka {1- : * {1-63 By Ae 2:71 Ae 2:69 From Eq. 2:30 Pe on tform фе cos de VATRE From Eq. 2:70 the actual tailpipe total temperature may be obtained from a tailpipe temperature indicator and the nozzle pres- sure ratio, provided a properly calibrated pickup is utilized. Using Eqs. 2:70 and 2:61, we see that mass flow may be determined from measurements of exit static pressure, total exit pressure, and total exit tempera- ture, Tex 2Y 17-1R P. Y-1 fe] dae 2:70 If the exit pressure equals ambient static pressure, then the Jones thrust is identical to the standard thrust. If Pet Poo the two thrusts are essentially the same up to a ratio of Per/Pe > 2, but at higher pressure where Ter is the exit total temperature, obtained from a calibrated thermocouple. 1 2:27 ratios, the difference increases, and is no longer negligible particularly for net thrusts at high flight Mach numbers. will be obtained by measuring the momentum change of the internal flow. However, there are certain possibilities for direct measure- ment which will be discussed first. (a) Direct Force Measurements Since the assumption of isentropic flow between stations w and e is open to some question, the equation for Jones thrust should be modified if more information, either theoretical or from measurements is avail- able in particular cases. (c) Application of Thrust Definitions to Particular Engine Installations The definitions of the preceding para- graphs may not be immediately applicable to particular installations. For instance, if the jet axis is at an angle to the direction of motion then this must be taken into ac- count in the application of the momentum equations. In most normal installations this angle will be small and the corrections to the definitions as quoted will be unimportant. Again, the jet pipe exit may be cut off ob- liquely. In this case the plane e will have to be arbitrarily defined to suit the particular case and any jet deflection will have to be taken into account. Some engine installations have cooling flow ducted round the engine and exhausting round the final nozzle as indicated in Fig. 2:6. In such cases, it is arbitrary whether the cooling drag including ejector losses be debited to engine thrust or added to the airframe drag. (1) Static Gross Thrust of Bare En- gine on Test Bed. By definition the gross thrust of a simple jet engine under static conditions, FGTB , can be measured directly on an uncowled engine by a force balance on a ground level test bed (Fig. 2:6A). The test bed must be arranged so that entry air can be drawn from a source of uniform total pressure; jf windage over the engine is not negligible or if there are any obstructions such as silencers in the path of the hot jet, these effects must be allowed for by correc- tions. The gross thrust obtained by direct force measurements includes the post-exit thrust appropriate to static conditions. Com- parison of thrusts measured by balance, with the Jones thrust obtained from pitot static traverse at station e showed agreement in a series of tests on a centrifugal engine to within 2% which is of the order of accuracy of the tests. (2) Static Gross Thrust of Engine In- stalled in an Airframe. By suitably suspend- ing an aircraft on a thrust balance, the static gross thrust of an installed jet engine can be measured. Such direct force measurements may disagree with static test bed results due to: 2:19 THE MEASUREMENT OF THE THRUST OF JET ENGINES Drag of parts of aircraft in flow induced by the jet at exit. General b. Thrust or drag of cooling air (Fig. 2:6) and air bled from the compressor which is sometimes ducted to exit round the final nozzle, or to engine and aircraft equip- ment. Methods of measuring the in-flight thrust of jet engines installed in aircraft will now be considered. The thrust under static con- ditions of the bare engine and of the engine installed in the airframe will also be con- sidered briefly insofar as such thrusts are used as a basis to evaluate thrusts from flight measurements. From the definitions previously presented it follows that thrust Changes in ram ratio and entry total pressure distribution due to in- efficiency of flight entry under static con- ditions (Fig. 2:6B). 2:28 Slove Entry I Flow Induced ! By Jet Y 6 6 Jogo que P p. 8 A) BARE ENGINE ON TEST BED Pressure Around Engine, P P Entry Condition Stotic e PLENUM CHAMBER CENTRIFUGAL ENGINE OD 444 Entry Condition in Flight B) SIMPLE INSTALLATION OF ENGINE IN NACELLE Thrust Measuring Trunnions Pressure Around Engine, p Flexible Seal DIFFUSER BURNERS EXHAUST 537ZZON C) ENGINE IN NACELLE WITH COOLING FLOW AND AFTERBURNING Fig. 2:6 Diagrammatic Sketches of Jet Engine Conditions for Thrust Measurement 2:29 (e. g., force an entry air ducting) is given by Fan-F85010,-U00) + A, 10, + Pool Such force balance measurements made on installed engines must, therefore, be used with due precaution if absolute values of thrust are required. They can be usefully used to check the thrust of a particular engine relative to the standard for a given aircraft engine combination. 2:73 from which it can be seen that the net standard thrust can be obtained from the trunnion thrust if the conditions at station 1 (the entry to the compressor) are measured. * (3) Thrust of a Jet Engine on its Trunnions or Bearers. It might be conven- ient in some cases, for example, for engines using reheat, if the thrust in flight could be obtained from measurements of the thrust force of the engine on its trunnions or bearers. For centrifugal engines with plenum chambers the relation between net standard thrust and trunnion thrust is given to a first approximation by FSN - Fg is normally less than 35% of B FSN so that the error in FSN due to errors in the compressor entry conditions can be kept small, Trunnion thrust F : F B SG - Ae (p - Po) A е In addition, Fe must be corrected for engine incidence and aircraft acceleration. Measurement of thrust of axial engines in flight by the trunnion method is therefore possible, though it should be considered as a last resort in view of the difficulty of the installation, i.e., the engine with no restraints other than the trunnions, the force measuring cells in the trunnions, and the pitot static measuring gear at the compressor entry with its danger to the engine if any parts broke away. 2:72 (b) Momentum Measurements from which it is seen that FB depends on the force of the engine due to the local static pressure p around it (see Fig. 2:6B). This force would be very difficult to meas- ure to the required degree of accuracy and trunnion thrust measurements for the centri- fugal type of engine with plenum chambers therefore are of little value for aircraft performance purposes. Tests by N.G.T.E.* have shown that the force acting through the engine trunnions can be measured in flight though the installation is not easy, but as shown above, this force is not simply re- lated to the thrust defined for aircraft per- formance purposes. Thrust by Integration of Measurements at the Final Nozzle. The standard (or Jones) thrust which must be measured if thrust and drag are to be obtained from flight tests, can be determined from measurements at the final nozzle exit. From the expressions for thrust previously developed, it follows that the standard or Jones gross thrust can be obtained precisely from traverses over the jet exit area measuring the following: For axial type engines (Fig. 2:6) the force on the engine due to the local static pressure round it is likely to be smaller than in the case of centrifugal engines and may be neg- ligible. With this assumption the difference between trunnion thrust and net standard thrust, that is, the thrust taken on the air- frame other than through the engine bearers (1) The total pressure in the direc- tion of the local flow, PT. (2) The static pressure, p. * An approximation from wind tunnel or en- gine brochure data could not be relied on to give FSN to within 5% in all cases. *Results unpublished. 2:30 (3) The direction of the local flow to the flight path $, where will be deter- mined from the local angle of swirl and the radial component, Also for the net thrust therefore necessary, particularly for un- choked jets, to measure static pressure at least at the center and walls. Under static conditions on a test bed, the Jones thrust measured by pitot static inte- grations can be compared with direct force measurements. In such tests (Ref. 8) the thrust by integration was 2% higher than balance thrust, and a similar result was obtained on unpublished tests on an axial engine. This difference is often of the order of the experimental error and it has not been possible to trace its source. (4) The local total temperature. Such transverses must take account of the jet boundary layer and any alteration of jet size with temperature. From the results of Ref. 8, it may be concluded that absolute measurements of thrust can be made in flight using a pitot static comb to an accuracy of 2%, though at the cost of some complication of installation. To make such complete measurements would clearly be an arduous task in flight, and acceptable simplification will be con- sidered. The local angles of swirl and pitch effect both the measurement of total head and static pressure and also the component of momentum in the thrust direction. If Prandtl type pitots and statics are used set parallel to the nozzle axis, then for conver- gent nozzles an error of 0.5% in thrust would be caused by a value of $ of approxi- mately four degrees. Under normal circum- stances in a well-designed jet engine o will be less than four degrees so that it need not be measured. (c) The Single Pitot Method* A very much simpler application of the momentum method of measuring thrust is the well-known single pitot method (Refs. 8 and 9), which on account of its simplicity has been used in Great Britain up to the present time for almost all routine meas- urements of thrust in flight in connection. with performance measurements. From Eq. 2:62 it can be seen that FSG 8 sao bilo () dae Ае Poo Ao A The need to measure the static pressure distribution over the jet exit has been dis- cussed in Refs. 8 and 9. In Ref. 8 it is shown that on an engine with a conventional final nozzle, values of (Pe - Po)/(Pte - Po ) varying from 0 at the wall to 0.3 at the center with elliptic distribution were meas- ured on the ground and in flight, i.e., the exhaust gases do not fully expand in the final nozzle. If no measurements of Pe had been made and the thrust estimated assum- ing Pe - Poo below the choke and Pte/Pe - 1.85 above the choke, then an error in thrust of 2% would occur with the engine just choked. This error increased to some 5% at Ptę/Pe = 1.2 and fell to 0.5% at Pte/Pe above 2.0. 2:74 In the single pitot method the relationship between FSG and the single pitot pressure Pér is taken to be of the form of Eqs. 2:64 and 2:65, i.e., FSG : fe PO Aė Boo 2:75 Static tests made on an axial engine gave similar results (unpublished). If accurate absolute measurements of thrust by the momentum method are required, it is *Editor's note: An averaging total head tube can be used to improve accuracy. 2:31 where he is an effective jet exit area which is obtained as a function of pressure ratio pe/Pc from measurements of FSG and per/Po normally made on a test bed at ground level conditions. Comparison of Eqs. 2:74 and 2:75 shows that the accuracy of the single pitot method in measuring thrust in flight depends on the inherent assumptions that the final nozzle total pressure distribution, static pressure, and static pressure distribution at a given value of per / Po is the same on the test bed and in flight. methods, analagous to the single pitot method in that they depend on the measurement of jet effective area de but he is based on other measurements. First, it is possible to base thrust measurements on wall static pressure measurements for example, in the jet pipe, upstream of the final contraction, or at the nozzle guide vanes, Such wall static pressures are steadier than single pitot pressures and are therefore easier to measure accurately, but they are more liable to consistent errors if there is any buckling of the wall or local deviation from smooth- ness near the static hole. A more reliable result can be obtained from the mean of several wall statics at a given station. Further, it is often necessary to extrapo- late values of Ae obtained on the test bed to higher values of pressure ratio per / Pc only obtained in flight. This is normally done assuming that above the choke per/P00 – 1.85, Ae remains constant, although 'altitude test bed measurements on a centrifugal jet engine showed an increase of a'e with pe/poo above 1.85 (Fig. 2:7A). 2:20 THRUST MEASUREMENTS WITH AFTERBURNING (a) General The definitions of thrust and principles of measuring it previously outlined are identical when afterburners are installed in the jet pipe. The magnitude of errors in thrust meas- urement of the single pitot method due to the above assumptions has been measured in flight on a particular centrifugal engine by comparing single pitot thrusts with thrust from a pitot static rake which showed that errors in single pitot thrust up to 6% were encountered in some conditions in flight (Fig. 2:7A). The application of methods of measure- ment may have to be modified due to (1) the very high temperature at the final nozzle, (2) the advantage of changing the final nozzle area when the afterburner is switched on, (3) the less uniform conditions in a final nozzle with afterburning, and (4) the larger volume of cooling air with large resulting ejector losses at the final nozzle. The single pitot method has the great advantage of simplicity and ease of instal- lation as the thrust is obtained from only one additional pressure reading. Errors due to the inherent assumptions should how- ever be watched, especially where flight conditions as regards air entry conditions, Reynolds number, nozzle pressure ratio, and exit static pressure (Fig. 2:7B) differ appreciably from the conditions of the engine calibration tests. Insufficient experience has been obtained in Great Britain to establish the best method for measuring thrust with afterburning, so the methods available can only be enumer- ated. (b) Thrust by Integration of Momentum at the Final Nozzle (d) Static Pressure Method Brief mention may be made of thrust This method is the ideal if a very full 2:32 1 . + 1 1.0 1 의 ​EFFECTIVE AREA, A. GEOMETRICAL COLD AREA Slope Checks With Altitude Test Bed 0.5 1 1 ! . 1.2 1.5 2.0 2.5 pre 1 PRESSURE RATIO, Pe Ground Level Test Bed Ground Level Test Bed Extrapolated Calibrated Against Roke In Flight Fig. 2:7A Variation of Single Pitot Effective Jet Area with Pressure Ratio. Centrifugal Engine. (Ref. 8) 1 N 0.3 CENTER C 0.2 Pe-Poo - DOO 0.1 WALLY O 1.2 * 2.0 1.5 2.5 Pre Poo Simple Theory Convergent Nozzle Choked Centrifugal Engine Flight Tests х Axial Engine Ground Tests Fig. 2:7B Static Pressure in Jet Final Nozzle Measured in Flight. (Ref. 8) 2:33 investigation of the thrust is required, but the difficulties of finding a material capable of withstanding the temperatures alterna- tively, or of cooling the rake, are formidable and would probably only occasionally be justified. A more hopeful alternative would be a single pitot static traversed relatively quickly across the jet. A temperature transverse would again be necessary to establish net thrust unless the engine mass flow were measured at another station, or was based on test bed measurements of engine characteristics. the turbine and the afterburner flameholders. Because losses between this plane and the jet final nozzle are a function of attitude and pressure ratio, this method is subject to in- accuracies due to this cause, unless the losses can be measured. For net thrusts, engine mass flows would have to be derived from a temperature calibration or engine characteristics. (d) Thrust Measurement by Aircraft Performance Measurements (c) Methods Based on Engine Calibration Methods such as the single pitot in the final nozzle, or static head upstream of the final nozzle, identical in principle to those already discussed for the non-afterburning engine dependent on static engine thrust calibrations are attractively simple, but are likely to be even less accurate than for non-afterburning engines, due to the more uneven distribution. A possible method of estimating thrust with afterburning would be to establish the drag of the aircraft, without afterburning, by measuring the thrust by one of the methods given in section 2:19, measure the increase in performance with afterburner in operation and hence determine the thrust with afterburning. This method is not likely to be very ac- curate, and a correction would have to be applied for any change in drag due to change of jet final nozzle area. The thrust could be measured by this method up to the highest speed with afterburning by measuring the drag without reheat in decelerating level flight (Ref. 11). The problem of air cooling a single pitot at the jet exit is thought to be soluble, but wall static measurements upstream of the final nozzle would be more attractive and should be tried out for reliability. (e) Trunnion Thrust An alternative method (Ref. 10) is to calibrate the thrust in terms of pitot and static measurements in the diffuser between As a last resort, the trunnion thrust of axial flow engines might be measured. 1 2:34 REFERENCES 1. PWA 01.60 by Pratt & Whitney Aircraft, “The Use of Operating Curves," December, 1946. 2. U. S. Navy Dept., Bureau of Aeronautics Power Plant Memo #16, "Aircraft Engine Horsepower Correction Methods," November, 1945. 3. Lush, K. J., “Differential Performance Reduction Methods for Turbo-Propeller Air- craft," U. K. Ministry of Supply Report No. AAEE/Res/275. 4. Liepman, H. W., and Puckett, A, E., “Introduction to Aerodynamics of a Compressible Fluid," Chapters 1, 2, and 3, John Wiley & Sons, New York; Chapman & Hill Ltd., London. 5. Jakobsson, Bengt, "Definitions and Measurement of Jet Thrust,” British Journal of Royal Aeronautical Soc., April, 1951. 6. A report in the British A. R. C. Series on the definitions of the thrust of a jet engine and of the internal drag of a ducted body will be issued shortly. 7. The Cambridge Univ. Aeronautics Lab., "The Measurement of Profile Drag by the Pitot Traverse Method,” Br. R. and M. 1688, January, 1936. 8. ) Stephenson, J., Shields, R. T., and Bottle, D, W., "An Investigation into the Pitot Rake Method of Measuring Turbo Jet Engine Thrust in Flight,” Br. Report AAEE/Res/265, December, 1952. 9. Wright Field Eng. D. W. Tech Note No. TN-T SEAL-2-6, “Determination of Turbo Jet Engine Thrust in Flight." 10. Flight Test Engineering Manual, USAF Tech. No. 6273, 1953. 11. Lush, K. J., "Preliminary Performance Assessment from Brief Flight Tests, Consid- erations of a New Technique,” Br. Report AAEE/Res/267, July, 1952. 2:35 AGARD FLIGHT TEST MANUAL VOLUME I, CHAPTER 3 A SURVEY OF PERFORMANCE REDUCTION METHODS Ву Kenneth J. Lush, AFFTC United States Air Force and edited by John K. Moakes Aeroplane & Armament Experimental Establishment United Kingdom I CHAPTER CONTENTS Page SUMMARY TERMINOLOGY 3:1 INTRODUCTION 3:1 3:2 THE PROBLEMS 3:1 3:3 METHODS OF PERFORMANCE REDUCTION 3:2 3:4 EXPERIMENTAL METHODS 3:2 (a) General 3:2 (b) Application to Simple Jet Aircraft 3:3 (c) Application to More Complex Cases 3:4 (d) The Effect of Interconnection 3:4 (e) The Assumptions Required 3:5 3:5 ANALYTICAL METHODS 3:5 (a) General 3:5 (b) Application of the Methods 3:5 (c) Control Systems 3:6 3:6 SUMMARY OF CONCLUSIONS 3:7 REFERENCES 3:8 SUMMARY The purposes of performance reduction are dis- cussed and the available methods classified. Argu- ments for and against each type method are made with reference to their application to various types of reduction problems. The classification is broadly as follows: (a) Experimental methods, which require no ad- vance numerical data from other sources, but because of the stringent requirements for their validity and convenience in use may be applicable only in certain cases. (b) Analytical methods, which do require advance numerical data from other sources. These may be further subdivided into: (1) Differential methods, based on linearized relationships suitable for small corrections, which depend for their existence on generalized information on engine behavior and airframe drag. These can only be applied when the generalized information is both available and convenient to use; thus, their use is virtually prohibited when compressibility effects are present. These methods are convenient for the reduction of isolated performance points, provided the necessary requirements for their applicability are met. (2) Performance analyses, which only require advance Information on engine behavior. These are possibly the best substitute (at least for tests at high altitude or Mach number) if experimental methods are either impracticable or very inconvenient. They may also be more convenient than differential methods, when performance data are required over a wide range of the variables, or when drag data are of primary importance. TERMINOLOGY b Aircraft Span coe Profile Drag Coefficient с Engine Parameter L Length Dimension in L.M.T. System Engine Speed z z Ni Speed of Compressor N2 Speed of Power Turbine р Ambient Air Pressure Q Fuel Flow ун T Time Dimension in L.M.T. System V True Air Speed Ve Equivalent Air Speed - V.To 8 W Aircraft Weight T Ambient Air Temperature ) regarded as ) enthalpy Tj Jet Pipe Total Temperature Po Air Density at Sea Level P Air Density at a Particular Pressure Height o Relative Density : P/PO PIPO 8 D Typical Length Representative of the Problem 3:1 INTRODUCTION RPM, manifold pressure, air tenperature, air pressure, and aircraft weight. Performance reduction is the term given to methods for deducing aircraft performance under standard conditions from tests made on the aircraft under non-standard conditions. With regard to the necessity for these cor- rections, it suffices to state that due to uncontrollable variations in atmospheric con- ditions, and aircraft and engine settings, it is usually impossible to make aircraft per- formance tests under precisely standard con- ditions. In operational use the air pressure, engine RPM and manifold pressure are easily varied over the working range. These quantities are, therefore, retained as variables or pa- rameters in the presentation of results; the steady level speed being plotred against air pressure (height) at each of several repre- sentative combinations of engine speed and manifold pressure (e.g., at the maximum for continuous weak mixture running and for combat). It follows that reduction methods are neces- sary to rationalize the test results so that they may be used for valid prediction and comparison. Performance reduction methods take many forms and the selection of a given form de- pends upon the nature of the tests and the amount and range of information which is required, as much as upon the nature of the power plant concerned. As regards air temperature, it is not gen- erally possible to obtain a selected value on any given occasion because the air tem- perature associated with a given air pressure varies from day to day. The practice* is, therefore, to plot the performance against pressure (height) for a “standard atmos- phere"; in this standard atmosphere" the air temperature associated with any given air pressure (or “pressure height'') is approx- imately the mean value for the conditions in which the aircraft is to operate. Before choosing a method it is desirable, therefore, to examine the purposes which performance reduction methods serve and the types of methods available for solution of the various kinds of problems. The aircraft weight as a variable is slight- ly different again, in that it bears some re- lation to the weight at take-off which is con- trollable. Therefore, it is usual to quote steady speeds at some standard weight which is related to the take-off weight: One such standard weight is 95% of the take-off weight, which is roughly the weight after climbing to operational altitude. 3:2 THE PROBLEMS The performance of an aircraft is a func- tion of a considerable number of independent variables. The number of these variables is sufficient to cause difficulty both in the description of the capabilities of the aircraft and the establishment of these capabilities by flight tests. In the description of the performance capabilities of the aircraft the difficulties are reduced by taking average values of those variables which are not con- trollable in practice. In general, the practice is to quoce perform- ance over suitable ranges of those variables which can be controlled, and for specified "standard" values of those variables which cannot be controlled. If the test programn could be so arranged that the values of the uncontrolled variables Consider, for example, an aircraft with piston engine and propeller, for which the steady level speed will depend on the engine * Editor's note: This represents British practice and should not be construed as representative of all cases. 3:1 3:4 EXPERIMENTAL METHODS (a) General were randomly distributed about the standard values, their effect would appear as scatter and the required data could be obtained directly, given sufficient test results. Such a technique would, however, be very expen- sive in flying hours and would take many months. For general purposes, it is quite impracticable. It is, therefore, necessary to have methods of deducing the performance under standard conditions from tests made under non-stand- ard conditions. This process is known as performance reduction. We have observed (section 3:2) that the independent variables on which the perform- ance of an aircraft depends may be control - lable or uncontrollable. It is convenient to distinguish further between: (1) variables which can be adjusted in flight, but not exactly to any predetermined value; (2) variables which can be adjusted in flight, with sufficient precision, to any pre- determined values in their working ranges. For brevity, type (1) will be referred to as adjustable and type (2) as fully control- lable. The information required from a particu- lar series of tests may be anything from the steady level speed at one particular height, weight and engine rating to air speed and fuel flow data for the whole practical range of working conditions. The choice of a reduction method may depend on the type and extent of the data required and the flight test program may depend on both; therefore, it may be that no one method is superior in all circumstances. This distinction is of importance when re- sults are required only at particular values of controllable variables. It is of prime importance in an experimental method of reduction that the required numerical data be obtained directly from the performance tests, so that if results are required at a particular value of any variable, the tests must either be made at that value (as can be done with a fully controllable variable) or must cover a range about that value (as must be done with an adjustable variable). 3:3 METHODS OF PERFORMANCE REDUCTION Methods of performance reduction may conveniently be divided into two classes which we will refer to as the 'experimental" and the "analytical" classes of method. In the extreme case, for example, where the performance is required at one partic- ular value of each of n controllable varia- bles, a single test would give an answer if all the variables were fully controllable, whereas 2n tests would be required if all were merely adjustable. Experimental methods will be those in which the performance under standard con- ditions is deduced directly from the flight tests on the particular aircraft with no nu- merical data being called up from other sources. If results are required over a range of all variables, - the distinction between fully controllable and adjustable variables is not, of course, of much importance; it affects only ease of analysis and interpretation of results. Analytical methods will be those in which numerical rical data from other sources are called up in deducing the performance under standard conditions from the tests under non-standard conditions. Thus, if a particular variable is adjust- able but not fully controllable, it may cause 3:2 From this we may deduce by dimensional analysis that N W inconvenience but does not make experimen- tal methods impossible. The difficulty intro- duced by a completely uncontrollable variable such as air temperature is more fundamental. We have already (section 3:2) dismissed as quite impracticable, arranging the tests so that the values of all uncontrolled variables were randomly distributed about their stand- ard values. talut ) :O * р 3:2 In Eq. 3:1 we had four independent var- iables, two of which (T and W) could not be conveniently controlled. In Eq. 3:2, after the application of dimensional analysis, we have only two independent variables, both of which can be controlled during any test where N and p can be varied. A conceivable alternative would be to make enough tests to give results over a range of each uncontrollable variable and deduce the standard performance by inter- polation. This would avoid the requirement that the test values of the variables should be randomly distributed about the standard values, but a large test program would still be required and waiting for the weather to produce a suitable range of air tempera- tures would be very objectionable except in fortuitously favorable circumstances. In this example, dimensional analysis has reduced the number of independent variables by two and so made control of all inde- pendent variables possible. Dimensional a- nalysis cannot reduce the number of inde- pendent variables in this case by more than two. By recourse to dimensional analysis, however, it is often possible to reduce the first difficulty by decreasing the number of variables involved, and to remove the second difficulty by associating the uncontrolled var- iables with variables which can be controlled. If now these independent variables (N/T and W/p) are fully controllable we can make level speed runs at the N/T and W/p which correspond to the values of N, T, W and p at which the steady level speed is required and deduce V/T (and thence V) with very little trouble. This is the approach used in current experimental methods. Let us consider first its application to jet aircraft, which is familiar; subsequently, we shall consider more complex cases. In practice with jet aircraft it has been found that W/p can be fully controlled by adjustment of p, but full control of N/T has not usually been attempted because it is usually difficult to correct the thermometer reading accurately for the effects of forward speed while the tests are in progress. (b) Application to Simple Jet Aircraft It is usual to assume that the steady level speed (V) is a unique function of the engine speed (N), the aircraft weight (W), the air temperature (T) and pressure (p). The air temperature is treated as an en- thalpy (dimensions L /T2 in the L.M.T. system), other quantities associated with temperature (such as viscosity) being neg- lected. Then we may write * Upon inspection, some of these variables appear not to be nondimensional. Powers of a typical length (D) of the problem have been omitted, because they are constant to the problem. In the general case the var- iables would be V ND W TTi' Dap filV, N, W, T, p) : 0 5 3:1 3:3 It is usual, both for level speeds and climbs (for which a similar relation may be deduced), to make tests over a range of N sufficient to cover the required range of N/VT and deduce the answer by interpola- Independent control of N and p is, of course, essential to this method. As an illustration of this type case, let us consider a turbo-propeller aircraft for which the steady level speed V is a function of N, W, T and p and the jet pipe temper - ature Tj, so that we may write tion. f31V, N, W, T, P, Tj) = 0 . 3:3 (c) Application to More Complex Cases In the above case, we have the following features: After application of dimensional analysis, we may write ( TT (1) Dimensional analysis has reduc- ed the numbers of independent variables by two, thus making the test technique prac- ticable, as this was the number of the orig- inal independent variables which could not be controlled. (2) Independent control of these two remaining independent variables is essential. With these in mind let us consider more complex engines, i. e., engines with one or more independent variables over and above those which were present in the case of the simple turbojet. It will generally be found that dimensional analysis will still reduce the number of independent variables only by two, but this does not matter if the re- maining independent variables can be fully controlled. If they are "adjustable" only, they may be a source of embarrassment; if so, the embarrassment increases with the number of adjustable variables. To give a single level speed under standard conditions, for example, at least two, four or eight tests would be required if there were one, two or three adjustable variables respectively. If on the other hand, full curves of performance against all variables were required, no embarrass- ment would result from adjustable variables. V N W N f4 = 0 PT; 3:4 If now, N/WT; (or any other convenient non-dimensional group containing the addi - tional variable T;) is fully controllable, inde- pendently of N/VT etc., the advent of the additional variables does not cause trouble. However, tests must be made over a range of this variable if it is adjustable only. If it cannot be varied independently of N/T and W/P, experimental methods cannot be used unless its test value is fortuitously exactly the required standard value or be- cause the controls have been so intercon- nected as to give this effect (see par. d below). This is equivalent in our case to saying that Tj must be varied independently of N and p; the pilot must have two indepen- dent engine controls. (d) The Effect of Interconnection Some engines may, however, have control interconnection schemes such that the oper- ator will not have independent control of the interconnected variables, and will not, for normal operational purposes, need it. If these control interconnections impose non-dimensional relationships*, the indepen- d * An additional difficulty is that one cannot make a test on a hot day at the N/T corresponding to combat engine speed on a standard day without exceeding the engine limitations. This has caused surprisingly little trouble. * The description "non-dimensional" is here applied to terms whose dimensions can be referred to standards from within the prob- lem. For example, W/p can be compared with some area, such as the wing area, associated with the aircraft. 3:4 dent variables will be reduced by the number of such relationships and the installation can still be treated experimentally. sometimes at a high weight (and low alti - tude) and sometimes at a low weight (and high altitude), any weaknesses in these as - sumptions can be detected and precautions taken. Large corrections should be avoided if possible. 9 If, on the other hand, as seems to occur in many cases, a non-dimensional relationship is not introduced, the pilot has lost indepen- dent control of the variables, and precise values of these variables (corresponding to standard conditions) cannot be applied. It is then necessary to use analytical methods of performance reduction. 3:5 ANALYTICAL METHODS (a) General (e) The Assumptions Required Use of a purely experimental method of performance reduction is only practicable if the number of independent variables is small, Inasmuch as there are usually many indepen- dent variables which affect the performance to some extent, it is necessary in using such a method to be able to assume that only a few of these are of importance. We have classified as "analytical" all methods of performance reduction which re- quire numerical data from sources other than the flight tests on the particular aircraft. Within this class there is a wide range of methods and the choice of the type of method to be used will be influenced by the informa- tion required from the tests and by the data available about the characteristics of propul- sive unit and airframe. For example, it is usual to assume that the level speed and the associated mass fuel flow of a simple jet aircraft are a function of engine speed, aircraft weight, air pressure and air temperature (considered as an en- thalpy) only. This implies that the following quantities, among others, can be omitted from consideration as independent variables: It is conventient to divide the methods into two sub-classes and to examine the merits and demerits of each sub-class of method with its associated flight test technique; further advance into detail is inappropriate to the present discussion. (1) The viscosity of the air; Methods of the first sub-class are essen- tially methods involving fairly small correc- tions and are commonly referred to as differential methods. One makes perform- ance tests under conditions approximating standard as closely as practicable and uses a reduction method to adjust individual results for the relatively small discrepancy between test and standard conditions. (2) The temperature rise of working fluid per unit mass of fuel injected. This implication is probably correct when the test and standard conditions do not differ greatly, provided that the aircraft and its engine are not near a “critical Reynolds number" or in a region where combustion efficiency is sensitive to working conditions, and provided that the working fluid is not changed (e.g., by dilution with water or ammonia). Methods of the second sub-class involve a performance analysis, the test program being designed to provide data suitable for analysis rather than to give test results under near-standard conditions. (b) Application of the Methods If the test program is so arranged that tests at a given N/T and W/p are made To estimate the corrections required for 3:5 the first type method, one usually linearizes the relations between the variables and esti- mates the slopes from an analysis of data from other sources, preferably some fund of generalized data. In other words, one must have prior information about the perform- ance response to changes in air temperature and weight, but the information need not be very precise as it is only used to estimate relatively small corrections. For methods of this type, the test program should cover a fairly wide range of air speed at one altitude at least. This assists with the estimation of the effect on airspeed of changes in thrust at constant altitude. The test program should be designed primarily to provide the required drag data. Such methods may be more convenient than the first type if the drag data required for the differential method are not available, or if the drag and thrust data cannot be put into an easily used form. Also, if drag data are required anyway, it may be more economical to merge the performance reduc- tion with this and to design the test program accordingly. Again, if the first type method is not much easier to use than the second, it may be considered worth while to go to a little more trouble to get drag data en passant, even though these are not a primary requirement. If such information is available and easy to use,* this this type method can be very convenient. It need impose no restriction on the test program. It does not, however, provide any internal check of the accuracy of the assumed data, nor does it provide drag data as part of the reduction process. To use the second type method known as “performance analysis”, one requires prior information about the performance of the propulsive unit and its response to changes in air temperature but not about the drag characteristics of the airframe, as drag data are obtained during the analysis. Obser- ved or estimated jet thrusts and/or shaft powers combined with estimated propeller efficiencies where appropriate give estimated drag or power curves. Estimates of the thrust or thrust power which would be available under standard conditions then lead to the standard perform- ances. A comparison of the drag curves obtained at various altitudes gives some check of the assumed performance of the propeller, etc. (c) Control Systems ** It is necessary with either type of analyt- ical method to be able to estimate how the thrust (which the power plant under test gives) varies with air temperature*. This requires a knowledge of the behavior not only of the basic power plant but also of its control system, * In the method used in the United Kingdom for piston-engine aircraft (Ref. 1), one engine parameter () and two airframe parameters (CDe and ebé) suffice to adapt generalized formulae to the particular aircraft. The method, however, is unsuitable in the pre- sence of compressibility effects on airframe drag and propeller efficiency. For instance, curves of (thrust power) I o/w..5) may be plotted against Ve VW for various values of W/p. * The aim is to show what the performance of the aircraft under test would be under standard conditions. It is not, for instance, to show what the performance would be if the engine did what its designer intended it to do. One will therefore need either to de- duce the engine performance under standard conditions from its observed performance under test conditions, or to estimate its per- formance under test conditions from power or thrust charts drawn up for standard condi- tions. The actual approach chosen will de- pend on the reliance which can be placed on the power charts, and the availability of suit- able instrumentation. 3:6 . methods). For these methods to be prac- ticable and economical, some fund of gener- alized data on the response of performance to changes in air temperature and aircraft weight is required from other sources from which linearized corrections are deduced. If the control system is such that the fuel flow at a given engine speed and intake pressure is independent of air temperature, the jet temperature would probably increase with air temperature; the engine output might then be limited by jet temperature at high air temperatures and by engine speed at low air temperatures. On the other hand, if the control system kept the jet temperature associated with a given speed constant, this change of regime would not occur, as the same limitation would always be the deter- mining one. A similar change of regime can occur near the full throttle height of a piston engine. It need not be troublesome. This information need not be very precise as it is only used to estimate relatively small corrections. If such information is available and easy to use, as it is in the case of the conventional piston-engined aircraft (Ref. 1), the method can be very convenient, particularly when reduction is required for flight tests covering only a small part of the full performance envelope. 3:6 SUMMARY OF CONCLUSIONS 3 We have seen that performance reduction methods exist in two broad groups, exper- imental and analytical. Similar but more flexible methods have been developed for turbo-propeller-engined aircraft (Ref. 2), which at the moment are somewhat inconvenient to use, due to the impracticability of generalizing the perform- ance characteristics of these engines at this introductory stage. Neither of these small correction methods is applicable, how- ever, in the presence of compressibility effects on airframe drag, nor may either be used with confidence in the region of the speed for minimum drag or drag power. (a) Experimental methods which require no prior numerical data for engine or air- frame are: (1) Practicable only if any automatic control or control linkage imposes only di- mentionally correct relationships; (2) Convenient if data are required over a range of all variables or if enough of the non-dimensional groups used in the reduction process are fully controllable. (2) Performance Analyses. Where experimental methods are either impractic- able or inconvenient, or where differential methods cannot be applied because of com- pressibility effects, the approach will neces- sarily be from performance analyses. This method also shows advantage over the differ- ential method when data are required through- out a wide range of all variables or when drag data are required. If these conditions are met, as they are in the case of the simple turbojet-engined aircraft and possibly some types of more complex turbine engine, the potential advan- tage of the tolerance shown by experimental methods towards ignorance of the aircraft and engine is decisive, especially at high alti- tude or high Mach number. The essence of the method is the estab- lishment of drag or drag power curves covering the speed range of the aircraft from measurements of observed thrusts and/or shaft powers combined with estimated propeller efficiencies where appropriate. From prior information about the perform- ance of the propulsive unit and its responses to changes in temperature and forward speed, estimates of thrust or thrust power which (b) Analytical methods which require ad- vance numerical data from other sources: (1) Small correction analytical methods (also referred to as differential 3:7 would be available under standard conditions then lead to standard performance. which they are required, and the amount of generalized data already available on the type of aircraft and power plant, in addition to the nature of the power plant and the degree of complexity of its control system. In certain problems a combination of experi- mental and differential methods may possibly be the best solution. The selection of the most suitable method for a particular application may therefore depend upon the amount of data involved, the region (height, Mach number, etc.) at REFERENCES 1. Cameron, D., "British Performance Reduction Methods for Modern Aircraft," U. K. Ministry of Supply Report No, AAEE/Res/170. 2. Lush, K, J., “Differential Performance Reduction Methods for Turbo-propeller Aircraft," U, K. Ministry of Supply Report No. AAEE/Res/275. 3:8 AGARD FLIGHT TEST MANUAL VOLUME I, CHAPTER 4 PERFORMANCE OF TURBOJET AIRPLANES By Daniel O. Dommasch Princeton University Sections 4:1, 4:5 through 4:12, 4:14 and 4:16 W. J. Hesse Naval Air Test Center United States Navy Sections 4:2, 4:3, and 4:4 T. W. Davidson Naval Air Test Center United States Navy Section 4:13 Phillip Hufton Aeroplane & Armament Experimental Establishment United Kingdom Section 4:15 TERMINOLOGY VOLUME I, CHAPTER 4 CHAPTER CONTENTS 4:1 INTRODUCTORY COMMENTS 4:2 Page 4:1 INTRODUCTION TO LEVEL FLIGHT PERFORMANCE TESTING OF TURBOJET AIRPLANES 4:1 4:3 TURBOJET ENGINE PARAMETERS 4:1 4:4 AIRPLANE-ENGINE COMBINATION CHARACTERISTICS 4:6 4:5 REDUCTION OF THRUST REQUIRED DATA TO BASIC FORM 4:13 4:6 THRUST AVAILABLE 4:16 4:7 RANGE AND ENDURANCE IN LEVEL FLIGHT 4:19 4:8 THE CRUISE CLIMB 4:23 4:9 THE NOMINAL BEST RATE OF CLIMB AIRSPEED 4:26 4:10 TIME TO CLIMB FOR JET AIRCRAFT 4:32 4:11 DETERMINATION OF MAXIMUM ENERGY STORAGE SCHEDULE AND TIME TO CLIMB TO A GIVEN ENERGY HEIGHT 4:37 4:12 PROOF OF CONSTRUCTION USED TO DETERMINE OPTIMUM ENERGY CLIMB SPEED FROM ACCELERATION RUN DATA 4:40 4:13 DESCENT PERFORMANCE OF TURBOJET AIRCRAFT 4:42 4:14 MEASUREMENT AND CORRECTION OF TURBOJET TAKE-OFF AND LANDING CHARACTERISTICS 4:42 4:15 TURNING PERFORMANCE OF TURBOJET AIRCRAFT 4:43 4:16 CONCLUDING REMARKS 4:46 REFERENCES 4:47 TERMINOLOGY CD هی هم بن CDG D E e Drag Coefficient Lift Coefficient Gross Drag Coefficient Parasite Drag Coefficient Thrust Specific Fuel Consumption lbs./lb. thrust-hour Characteristic Engine Diameter Endurance Span Efficiency Factor FG Gross Thrust FN Net Thrust FR Ram Drag G Gf hp Weight Airflow Rate Fuel Flow Rate as an Energy Rate Pressure Altitude Energy Height he L Length M Mach Number Z 3 p Pn Q q R Mass Engine RPM Ambient Air Pressure Pressure at Any Engine Station Mass Airflow Rate Dynamic Pressure (pv² =ypM²/2) Range 1 TERMINOLOGY (Continued) RC T Tn t V Vcal Wf W μ պ ♡ P Y σ Δ Subscripts Rate of Climb Ambient Air Temperature Temperature at Any Engine Station (as enthalpy) Time True Airspeed Calibrated Airspeed Weight Fuel Flow Rate Time Rate of Change of Energy Height Absolute Coefficient of Viscosity Efficiency Pressure Ratio P/Po Temperature Ratio T/To Air Density slugs/ft.3 Cp/cv Specific Heat Ratio P/PO Density Ratio Increment of Sea Level t Total Conditions Superscripts '(prime) Test Observed Value 4:1 INTRODUCTORY COMMENTS Because it is not possible to presume that a fixed (independent of Mach number) rela- tion between C₁ and Cô exists throughout the speed range covered by jet-propelled air- craft, the reduction procedures convention- ally employed with reciprocating engine air- planes are not applied to jet aircraft. Rather, the procedures used depend either on dimen- sionless parametric relations or on analyses of particular conditions of flight. It is possible to present jet performance in terms of several different sets of para- meters, each set having advantages in a given case depending on such things as equip- ment available, time available, accuracy re- quired, and so forth. The various procedures are discussed in the body of the chapter and it is up to the test engineer to determine which procedure should be used in a given case. The material in this chapter is essentially the contribution of people presently or re- cently associated directly with naval aviation and, for the most part, is similar to material appearing in an existing preliminary edition of the Naval Air Test Center Flight Test Manual (Ref. 4). However, certain portions of the material have been rewritten, taking into account simplified techniques or im- provements which have made their appear- ance subsequent to the publication of Ref. 4. 4:2 INTRODUCTION TO LEVEL FLIGHT PERFORMANCE TESTING OF TURBO- JET AIRPLANES The theory of reducing the level flight performance data for any flight vehicle is based on the nature of its power required and available curves, where these curves are defined in terms of thrust, power, or some engine parameter such as RPM. The parameters used for the reduction of data from turbojet-engine-powered aircraft stem partially from an analysis of the engine characteristics and partially from an anal- ysis of airframe characteristics. The en- gine parameters will be considered first and then combined with the airplane parameters to produce the combined engine airplane parameters which are actually used for level flight performance data presentation. 4:3 TURBOJET ENGINE PARAMETERS The performance characteristics of a tur- bojet engine depend on many variables. These variables include the engine size, engine RPM, flight velocity or Mach number, am- bient pressure, ambient temperature, and the internal efficiencies of the component parts. To present the performance characteris- tics of a turbojet engine in a practical way, it is necessary to collect the aforementioned variables into a minimum number of signif- icant groups so that the number of test flights or wind tunnel tests and plots required to present the complete performance picture of the engine may be considerably reduced. This grouping is accomplished by a di- mensional analysis of the variables which affect engine performance. If there are n variables and r fundamental units (such as length L, Mass m, and time t) then, as is well known, the equation relating the variables can be expressed in terms of (n − r) dimensionless numbers. The working procedure is as follows: · (a) Select r variables which include a group of all the fundamental units from the list of n variables, then (b) Set up dimensionless equations com- bining the variables in (a) with each of the others in turn. It should be emphasized that the equation which results from this procedure is no better than the original assumptions made; i.e., if one variable is originally left out, it will also be left out in the final answer. The variables and performance quantities perti- nent to the turbojet engine with symbols, units, and dimensions are listed in Table 4:1. 1 4:1 TABLE 4:1 - VARIABLES AND PERFORMANCE QUANTITIES OF THE TURBOJET ENGINE CHARACTERISTICS SYMBOLS UNITS DIMENSIONS VARIABLE Flight Speed V ft./sec. Ambient Air Temperature T °R Ambient Air Pressure P lb./in.2 Engine RPM N Engine Size or Diameter D ft. Viscosity μ Component Efficiencies η η none rad./sec. slug/ft.-sec. L m/Lt none L/t L2/t2 m/Lt² 1/t PERFORMANCE QUANTITY Thrust (FG, FR, or FN) F lb. mL/t2 Air Flow* QG lb.m/sec. m/t Fuel Flow (energy input) Gf ft.-lb./sec. mL2/t3 Pressure at Any Engine Section Pn lb./in.2 m/Lt2 Temperature at Any Engine Section** In °R L²/t² 2 *Note that this is a mass flow rate expressed in pounds mass per unit time rather than the customary slugs mass per unit time. ** Temperature is here considered as a measure of enthalpy. Our purpose is now to express F, Ga, Gf, Pn, and Th as functions of the listed variables, the functional relationship to be expressed in an equation made up of a group of terms, each of which is a dimensionless number. We shall first consider the thrust. From an analysis of the variables which affect thrust, we can write F = f ( V, T, P, N, D, H, nr. 7c. 7b. nt. nn) Eq. 4:2 consists of six variables. Since we have three fundamental units, Bucking- ham's Pi Theorem states that we can express these variables in an equation made up of three dimensionless numbers. Performing the analysis, we have: V PD²= f( ND · √T). ND 4:1 Since the component efficiencies are pri- marily functions of V, T, p, N, and D, and μis primarily a function of T and p, we can sim- plify the above equation and write thrust as a function of its prime variables as F = f ( V, T, p, N, D ) 4:2 *For simplicity, viscosity has been omitted in this relation. However, it is pointed out that at altitude, the Reynolds number term which would result from a consideration of viscosity may produce important effects. 4:2 For convenience we may recast the above relation in a somewhat different form. Any of the ratios can be turned upside down (the ratio is still dimensionless) and ND in a number can be replaced by V or T because of similar dimensions. Changing the above relation by these procedures gives F f BD₂ = 1 (Y, ND.). PD2 4:3 It is conventional to replace the ratio V/✓T by Mach number*, to eliminate Dfrom the equation for an engine of fixed size (this elimination cannot be made on engines with variable exit nozzles or variable inlet guide vanes) and to refer the pressures and tem- peratures to standard sea level conditions by using the following technique: F PD2 = F 14.750** 14.780** (const) 8 14.7pD2 The above relation represents an implicit function of the variables which affect thrust. The purpose of flight testing, wind tunnel tests, or theoretical calculations is to solve this equation explicitly. The solution is best shown in graphical form as indicated in Fig. 4:1. This one plot (Fig. 4:1) represents the thrust performance of an engine at all alti- tudes and all Mach numbers over the range shown. If parameters were not used, it would require many curves and curve sheets to present this same information. Therefore, it is evident that the use of suitable para- meters saves much time and money when the performance characteristics of a turbo- jet engine are determined by flight tests, wind tunnel tests, or calculations. Table 4:2 on the following page presents a summary of the functional relations among the turbojet engine performance parameters, obtained by dimensional analysis. and ND √519 ND √519√T NO √519 √8 where = P/P。 and 8 - T/T。. Thus, Eq. 4:3 can be written 통 ​+ (M. 씀​). N (const)√8 * Illustrated in the following equation: 1 V √gy R V 4:4 √gy RT = (const) M 4:3 F/8 M = 1,0 M = 0.5 M = 0.0 N/√6 Fig. 4:1 Thrust Parameter Variation Thrust TABLE 4:2 - FUNCTIONAL RELATIONSHIP OF THE TURBOJET ENGINE PERFORMANCE PARAMETERS Engine Parameter General Form Obtained From Dimensional Analysis Relation Referred or Corrected to Standard Sea Level Conditions for Constant Diameter Engine* F PD2 + (ND, M) F f 8 (N,M 0 Air Flow Fuel Flow Ga PD2 Specific Fuel Consumption Gf FJT f(ND, M) T Gov. 1(MD.M) = ND DO² / T = f( ND, M) Gf PD2 T = N, M G₂+(M) 8 816 • 1 (A.M) Gf 8√8 0%· ·(N·M) Gf F√O Fuel-Air Ratio Gf Go T f + (ND, M) Gf Ga Glo··(MM) N Pressure** (at any section) Pn f P (ND, M) Pn 8 (NO,M) • Temperature** f 푸 ​Tp = + (ND, M) In 꼼 ​= + (M) (at any section) *When the engine diameter is not constant such as with an afterburner installation, it is necessary to retain the engine size, D, within the functional relation. Subscript n refers to any section. The parameters developed by dimensional analysis can also be used to correct or refer observed data to standard sea level conditions or to any specified altitude. This fact is, of course, extremely useful in the test work of engines and it has led to giving the names "corrected RPM to N/√@" and "corrected thrust to F/8", etc. 4:4 Table 4:3 presents a collection of the corrected or referred turbojet engine parameters. The table with its footnotes is self-explanatory; however, it should be noted that when it is desired to correct or refer data to some altitude other than sea level, the following changes should be made: Replace 8 with 8 where 8 = p observed/p at selected altitude and replace with a where = T observed/T at selected altitude. 8 TABLE 4:3 - TURBOJET ENGINE CORRECTED OR REFERRED PARAMETERS Corrected or Referred Quantities Quantity Observed Value General Form Corrected to Corrected to SSL Variable SSL Constant Diameter Engine** Diameter Engine * F Thrust F PD2 F 842 ико Air Flow Qa QO√T Q0√o Qo√8 PD2 842 8 Gf Gf Fuel Flow Gf PD2 √T 802√0 Gf 8√8 Specific Fuel Gf Gf Gf sfc Consumption F√T F√8 Gf Gf Gf Fuel-Air Ratio f/a Ga T Gae Gae Pressure*** Pn (at any section) Pn 8 Temperature*** Tn T (at any section) Engine RPM ND N T 2/3 0/500/50 Pn NA --- N 7/6 V V V Flight Velocity V Flight Mach Number M M M M * SSL - Standard Sea Level Conditions. **▲ in this column is defined as the ratio of the nozzle exit diameter at any operating condi- tion to nozzle exit diameter at standard sea level condition. We note that these parameters will not produce generalized data when changes in flow regime are produced by diameter changes. *** Subscript n refers to any section. 4:5 an aircraft from these plots, but the stipula- tion is for a given aircraft weight and stand- ard altitude conditions; To obtain the char- acteristics at different airplane weights, one would require another set of these four plots, or to determine the characteristics at some pressure altitude where the temperature is not standard, one would require still another set of plots.* One set of plots is required for every change in weight and/or altitude deviation from standard values. The plots are valuable within their limita- tions, namely, the performance story of an aircraft at one weight and under standard altitude conditions. The method of obtaining these plots depends, of course, upon the tech- nique of applying fundamental dimensionless parameters to the airplane-engine combina- tion as was done for the engine in the previous section. (a) Airplane Parameters The thrust required by an aircraft in steady level flight is given by F = CD ½ PV²S = (CD₂ + C²²) ½ PV's. TEAR Since the lift coefficient can be expressed in terms of aircraft weight and speed, the above expression may be written in para- metric form as 2 F POYS 2 4:4 AIRPLANE-ENGINE COMBINATION CHARACTERISTICS Our discussion thus far has been devoted to engine characteristics to show where cer- tain of the parameters used in the reduction of turbojet aircraft data originate. Our pur- pose now is, to examine the performance characteristics of the overall vehicle, the airplane-engine combination. When we con- sider such a combination, the airplane vari- ables such as aircraft weight, angle of attack, etc., enter into the overall functional relation- ships. Before developing the equations necessary to handle the airplane-engine combination, it is desirable to examine the real purpose and goal of the level flight performance eval- uation of a turbojet-powered airplane. The aircraft must first be flight tested over the complete range of speed and altitude, suffi - cient and accurate data must be obtained during these tests, and then the data must be reduced and assembled for logical presenta- tion, Fig. 4:2 presents, four plots which rep- resent a typical method of presenting the most important performance characteristics of an airplane equipped with a fixed geometry jet engine. The data for Fig. 4:2 are based upon a given airplane weight and for standard con- ditions at all altitudes. These plots show: (1) variation of standard altitude with flight Mach number for various engine RPM's; (2) variation of standard altitude with fuel flow for various engine RPM's; (3) specific range (miles/lb. fuel) variation with flight Mach number; and (4) endurance (hours) variation with flight Mach number. It is evident that one can determine the complete level flight performance picture of 4/00 CDe M² + 14.77 S(EAR) (~)* (M²) 4:5 * As discussed later, the use of dimensional parameters sharply reduces the number of required plots. 4:6 MILES PER POUND OF FUEL STANDARD ALTITUDE,FT. 80% R.P.M. 90% R.P.M. 95% R.P.M. 100% R.P.M. STANDARD ALTITUDE, FT. MACH NUMBER, M FUEL FLOW RATE 40,000 FT. 30,000 FT. 20,000 FT. 10,000 FT. ENDURANCE, HOURS MACH NUMBER, M MACH NUMBER, M Fig. 4:2 Four Plots Showing Important Level Flight Performance Characteristics of Turbojet 4:7 40,000 FT. 30.000 20.000 10.000 80% R.P.M. 90% R.P.M. 95% R.P.M. 100% R.P.M. 1 The above equation is in general explicit for flight below the drag divergence Mach number; however, for flight above this Mach number, CDe becomes a variable and is a function of M and W/8. Thus in general, Eq. 4:5 must be written as (ignoring as before the effects of viscosity) f t. 등​• + (M. 뿡​) W). (b) Combined Parameter Relations 4:6 From an analysis of the engine alone we have (from the last section for a fixed geo- metry engine) 4/00 · Gf } (M) 4:7 in the airplane dimensional analysis, Rey- nolds number will appear in the equations, just as it would appear in Eqs. 4:7 and 4:8 if viscosity were included in the engine anal- ysis. When viscosity is ignored, the airplane parameters, like the engine parameters, will not generalize completely. Thus, low altitude data cannot be used to predict ac- curately the high altitude characteristics, and it is necessary to obtain performance data over the complete range of altitudes. Inasmuch as in steady level flight, net thrust is equal to drag, Eq. 4:7 can be equated to Eq. 4:8 and, since the fuel flows in Eqs. 4:10 and 4:11 are equivalent, these equations can also be equated. Equating either or both sets of equations shows that M = ·(*. *) 4:11 4:8 A similar dimensional analysis of the variables which affect the drag of an airplane or the analysis shown above yields a drag parameter for the airplane which is 이쁨 ​M m). This equation merely states that the flight Mach number of a turbojet-powered aircraft with fixed geometry engine is a function only (for the assumptions made) of the parameters N/√ and W/8. From this equation we can also write = f (~, M) or I = f (W,M) * 4:12 and Gf N + (~~/ M) O Gf or (WM)* 4:9 8√Ā 4:13 A further analysis of the airplane, that of equating the heat energy input as a func- tion of the airplane drag energy, yields the following fuel flow parameter G₁ 4:10 We note that in the above equations, air- craft weight appears in one of the parameters. It is pointed out that if viscosity is included *It should be noted that the latter forms of Eqs. 4:12 and 4:13 apply to any type of engine geometry, fixed or variable, because they come directly from Eqs. 4:9 and 4:10, which were derived from airframe considerations only. 4:8 } W 8 Level flight performance testing of air- planes with fixed geometry engines has been accomplished most frequently in the past on the basis of solving experimentally the two following functional relationships and M = f (学​) Gf 8/10 (NM). A typical graphical solution of these equations is sketched in Fig. 4:3. It is readily apparent that the level flight performance characteristics curves (the 4 plots of Fig. 4:2) can be obtained from the two plots of Fig. 4:3. In fact, any given num- ber of these sets of four plots corresponding to different airplane weights or non-standard altitude conditions can be obtained from Fig. 4:3. Actually, all the level flight performance data are generalized and available on the two engineering plots of Fig. 4:3. The four plots of Fig. 4:2 are merely convenient plots and are special cases because they are applicable only to a certain specified airplane weight. We can thus state that the purpose of level flight performance testing is to obtain suf- ficient data so that engineering plots (such as Fig. 4:2) can be constructed. M = 0.5 M = 0.6 M = 0.7 M = 0.8 Gf 8/M Lines of Constant W/S 2/10 N e N Ꮎ Fig. 4:3 Graphical Solution of the Performance Relations of the Turbojet Powered Airplane with Fixed Geometry Engine 4:9 7 It should be noted from the above plots that thrust, which is a very basic parameter, does not appear. For the fixed geometry engine, the performance characteristics have not been measured in terms of thrust until recently for several reasons. First, thrust has been more difficult to measure than the parameters N/√, M, and W/8 and second, service aircraft are not equipped with thrust meters. Therefore, it is desirable to present performance characteristics in terms of variables which can be evaluated by the pilot in the cockpit. It is pointed out, however, that the per- formance characteristics of the aircraft can be presented in terms of the thrust para- meter. Employing this system, plots of the following parameters are made.* = f 풍​(뿔​M) W Gf = f 3/8·1 (2·M). • 4:14 FG/8 FG values at the minimum points are about the same. W/8 - 80,000 W/8=50,000 W/8=20,000 Mmox at 100% RP.M. MACH NUMBER, M 4:15 Plots of these functions will also yield the performance story of an aircraft. The above thrust parameter may be expressed either in terms of net thrust or gross thrust. At the Naval Air Test Center (NATC), gross thrust is preferred because it is easier to measure than net thrust and accurate results are obtained with a relatively simple gross thrust meter installation (see Ref. 1). *Editor's note: The use of RPM as a re- placement for airplane drag or thrust re- quired has not been entirely satisfactory for the reason that the thrust RPM relationship varies somewhat from engine to engine and with time due to engine wear. Similarly, fuel flow data are easier to generalize when pre- sented in terms of Eq. 4:15 rather than Eq. 4:8. Fig. 4:4 Typical Thrust Required Curves for Turbojet Airplane Engine Combination Fig. 4:4 presents the typical variation of the gross thrust parameter with flight Mach number for several values of W/8. These curves (Fig. 4:4) represent the gross thrust required curves of the basic airplane-engine combination. (See Ref. 2 or Chapter 2 for satisfactory methods of deter- mining the gross thrust available curves for an airplane-engine combination.) Thus, when the thrust available and thrust required curves are combined, the level flight per- formance characteristics are known. It should be recognized that when variable geometry engines are performance tested in conjunction with an aircraft, it is advanta- geous to use the thrust parameter function because the parameter Eq. 4:11 then contains an additional term.* * See Chapter 3 for a more detailed discus- sion of this type of problem. 4:10 : Eq. 4:11 for a variable geometry engine then becomes f W M-1 (M··A) = (NO 8 * and, for a dual rotor engine, becomes (***): 米 ​N1 N2 M = f 4:16 4:17 The thrust parameter relationship even for the variable geometry engine or the dual rotor engine is still given by Eq. 4:14, which contains only three terms. Thus far, two basic procedures have been used at the NATC for presenting level flight performance data of jet-powered aircraft, namely, the required curves in terms of the RPM parameter (Fig. 4:3) and the required curves in terms of the thrust parameter (Fig. 4:4). Emphasis has recently been placed on the thrust method of presenting the data because of its more direct approach, elimination of inherent engine tachometer errors in the data reduction procedures, its simplicity for variable geometry engines, and, in general its application to the solution of more varied types of level flight performance problems. A very simple gross thrust meter is de- scribed in detail in Ref. 1. Consider, for example, the gross thrust available and required plots of Fig. 4:5. The thrust required curves of Fig. 4:5 are obtained from readings of the thrust meter mentioned above (to eliminate the nec- essity of cross plotting, flights may be at- tempted at constant values of W/8), and the * In these equations, ▲ is the ratio of vari- able area to some reference area; N₁ and N2 are the two rotor speeds in a dual rotor engine. thrust available curves from the methods outlined in section 4:6. Once these curves are found, the level flight performance story is told, namely: (1) maximum Mach number; (2) minimum thrust; (3) stall speed; (4) minimum level flight speed (which is important for catapult launching of heavy jet aircraft because it has been found that in some aircraft the minimum catapult end speed is dictated by this point instead of the following point); (5) the aerodynamic stall. (c) External Store Drag As mentioned before, the (gross) thrust method of performance presentation is more versatile than the corrected RPM method. One of the newer uses of the gross thrust method is its application to the evaluation of drag of external stores such as mines, bombs, rockets, etc. FG/8 Gross Thrust Available Gross Thrust Required- Gross Thrust Required, Low W/8 High W/8 MACH NUMBER, M Fig. 4:5 Thrust Available and Required Curves 4:11 1 i The reduction theory for the evaluation of external store drag can be developed from an analysis of the two gross thrust required curves shown in Fig. 4:6. The curves shown in Fig. 4:6 correspond to the airplane clean configuration and the configuration with some external store, both at the same value of W/8. Consider now points (1) and (2) at the same Mach number and assume operation at sea level (to sim- plify the equations). For the two configurations, the gross thrust is given by FG₁ = D+FRI FG₂ = D+FR₂+DS The change in ram drag, FR, can be ex- pressed in terms of gross thrust and air flow by the following analysis. The rate of change of ram drag with respect to gross thrust is given by ΔΕR AFG = which for V or = (G2V2 - G|V₁) 9 AFG V2 (see Fig.4:6) becomes AFR=Y AFG AFG g • 4:18 where FR is the engine ram drag, D the basic airframe drag, and DS the external store drag including interference effects. The change in gross thrust is given by FG = FR+DS 4:19 AFR (AFG) (AFG) 4:20 The term AFG/AG can readily be obtained from engine specifications curves or approxi- mated by 2Vef/g where Vef is the effective exhaust velocity. Inserting the results of the above equation in Eq. 4:19 gives FG/8 With External Store Clean Some W/8 V 1 g AFGI AFG =DS AG Θ 10 4:21 MACH NUMBER, M Fig. 4:6 Gross Thrust Required Curves for the Clean and External Store Configuration (AFG and V from flight data and ▲ FG/AG from specification curves.) The above equation thus provides a method of determining the actual drag of the external store. Eq. 4:21 can also be extended in certain cases to calculate the performance of an air- plane with some store that has been previously 4:12 tested on another airplane. Suppose, for ex- ample, airplane x has been tested in the clean configuration and with stores, and the flight data are presented as shown in Fig. 4:6, and suppose further that airplane y has been tested only in the clean configuration. Now if the external store drag Ds is the same on both airplanes (Dsx = Dsy), we can write वै AFGy = (AFG) FGX | + AFG) AG X airframe drag and the ram drag in steady level flight divided by the dynamic pressure and wing area. This coefficient is useful if combined airplane-engine characteristics only are being considered and, as we shall show, the gross drag coefficient is in special cases a function of lift coefficient and Mach number as is the airplane drag coefficient. To prove this latter statement, we con- sider the relation already established that FG = f(WM) (for steady level flight) and consider the nature of the parameter W/8 at constant Mach number during one g flight. In steady level flight, we know that W = YPM2 DM² CLS = Po. YSM². CLS. 2 Therefore, W = (PO M²) · (CLS) 2 4:22 Since all terms in the above equation are known except ▲ FGy, we can calculate this quantity and apply it to the clean configura- tion curve to obtain the gross thrust required curve for the external store configuration. It is pointed out that careful judgment must be used before one can assume that the in- terference drag is the same for airplanes x and y. When such an assumption can be made, considerable testing time and money can be saved. 4:5 REDUCTION OF THRUST REQUIRED DATA TO BASIC FORM For the purpose of simply determining if an airplane meets its design specifications, it is not often necessary to reduce thrust required data to its most fundamental form; i.e., plots of drag coefficient as a function of lift coefficient and Mach number. However, when precise corrections of climb data and the like are required, a knowledge of drag variation characteristics becomes important and, of course, such information is also of great interest to design engineers. We shall, therefore, briefly consider here methods of reducing flight test data to terms of aerody- namic coefficients. First, we note that it is possible to define a new drag coefficient, the gross drag coef- ficient, which is simply the sum of the and since po, Y, and S are constants Therefore, W = (M²CL) times a constant FG = f (M² CL, M) = f(CL, M). = If we measure FG in steady level flight, then FG equals the sum of the airplane drag 4:13 and the ram drag, so that FG qS 8f(M, CLL. = CDG урма 2 S where = free stream density ρ V = true flight speed A = stream tube area far ahead of the engine, the value of C CDR becomes Thus, CDG f(M, CL) (M²) (constant) = f (M, CL) 4:23 CDR 00 DR PAV² q S PV2 S 2 الله where CDG = gross drag coefficient. 4:25 As discussed in Chapter 2, we may readily measure gross thrust and we may also de- termine net thrust from flight tests, although the necessity of measuring total tailpipe tem- perature somewhat complicates the net thrust determination.* In this connection, Eq. 4:23 allows us to establish that the area of the stream tube of entering air far ahead of the jet intake has a cross sectional area which is also a func- tion only of CL and M for the steady level flight case. For this case we have CDR where CD airplane drag coefficient and, since = ram drag coefficient = CDG - CD CDG = f(CL, M) and CD = f(CL, M) CDR = √2(CL, M). Now, inasmuch as 4:24 DR = ram drag = pAV² = FR in steady level flight * For many practical purposes, the standard tailpipe temperature gage (properly cali- brated) will provide a sufficiently accurate temperature reading. From Eqs. 4:25 and 4:24, A = f(CL, M). 4:26 Eq. 4:26 shows that the engine mass flow during steady level flight depends on C₁ and M and thus, in theory, we may relate required engine RPM to fundamental aerodynamic parameters. The foregoing work serves to illustrate the interdependence of airplane-engine char- acteristics under steady flight conditions and shows that steady level flight data may be presented in terms of various equivalent sets of parameters. Once we deviate from the special case of steady level flight, however, it is apparent that the equivalence of the parameters ceases to exist. Consequently, curves of airplane performance presented in terms of engine RPM or gross thrust or gross drag coeffi- cient cannot be considered sufficiently gen- eral for a thorough-going performance eval- uation because results expressed in terms of parameters such as these have meaning only under certain restricted circumstances. If we wish to evaluate airplane perform- ance during climbs, descents and under con- ditions of normal and longitudinal accelera- tion, and moreover, if we wish to correct such flight data to standard, it is essential that we be able to separate specifically 4:14 provide the following information for each test point (a) Drag, D (b) Gross weight, W (c) Ambient pressure, p (d) Mach number, M In turn, we may then compute for each point D со = YPM2 S 2 CL W урма S 2 where y = 1.4 S = wing area. This provides us with the plots of Fig.4:7 with each curve representing a series of runs at one nominal altitude. LIFT COEFFICIENT, CL MACH NUMBER, M airplane and engine characteristics. For- tunately, we have the means to do this, and level flight tests serve to provide part of this means for they allow us to determine the nature of the airplane lift-drag polar by actual measurements. As explained in Chapter 2, the net thrust may be measured in flight, and if it is meas- ured in steady level flight, it equals the air- plane drag. Accordingly, we may make a series of level flight runs at various altitudes and obtain data in the form of drag and weight versus Mach number for each set of runs. For each test point it is important that the aircraft be stabilized; however, succeed- ing points need not be at exactly the same pressure altitude. Tests should be conducted starting at Vmax and each succeeding run made at a lower speed until the minimum speed for stabilized level flight is reached. Data to be recorded consist of pressure altitude, ambient temperature, total pres- sure, total tailpipe pressure, and tempera- ture and fuel flow. These data are used to DRAG COEFFICIENT, CD MACH NUMBER, M Increasing Altitude Fig. 4:7 4:15 Increasing Altitude → DRAG COEFFICIENT, Co Increasing CL MACH NUMBER, M Fig. 4:8 These data may now be cross-plotted to give curves of Cp at constant C vs. M as shown in Fig. 4:8 (or cross-plot may be prepared of Co vs. CL at constant M). Since the data of Fig. 4:8 have been re- duced to a basic form, they may be used to compute the airplane drag under varying situations and are not restricted to the spe- cial case of level flight. We note that in special instances, the in- terference flow produced by engine operation may alter the relations of Fig. 4:8 depend- ing on the airplane configuration; however, such cases should make themselves apparent in the scatter of data in the figure, in which case further detailed tests will be required to determine the cause of this interference. We now turn to the topic of determining standard engine thrust available at a given altitude. 4:6 THRUST AVAILABLE Because the engine is used merely as a measuring instrument when thrust required data are obtained, engine corrections to standard conditions need not be made for thrust required runs. However, to deter- mine Vmax, we must know: (1) what thrust the engine should have given under test con- ditions; or (2) what thrust it would give under standard conditions. In Chapter 2 and in the appendix to Chapter 7, the problem of standard thrust available is briefly considered and it is pointed out that engine charts are not too reliable a source of standard information, inasmuch as they do not necessarily represent properly the operation of the engine as installed in the airplane, nor, in some instances, are the specified operational limits actually obtain- able in the air. Moreover, it is well recog- nized that not all engines of a given model provide the same value of F/8 at a fixed set of values of M and N/√8. Accordingly, for really precise work, the engine thrust characteristic should be tied down by a comprehensive set of flight runs, which procedure, if attempted, would be rather costly as regards both time and money. If one is satisfied with something "less than the best," the maximum speed at any altitude may be determined using several approxi- mate procedures, two of which are considered here. In the absence of thrust measuring equip- ment, thrust required data are presented using the parametric relation N/√ē = f(M, w/8) discussed earlier in this chapter. Curves of N/√ē vs. W/8 for various Mach number values are readily obtainable from tests similar to those described in section 4:5 for the determination of drag coefficient variations. To obtain the desired curves, we first prepare curves of N/√ vs. M and W/8 vs. M for each set of test runs and then cross-plot to obtain curves of N/√ vs. W/8 for constant Mach number, as shown in Fig. 4:9 4:16 N/√6 Now suppose we wish to know Vmax at some weight W at an altitude where the pres- sure ratio is 8, and the temperature is 8. We first compute W/S locating point (1) on Fig. 4:9. Then for 100% RPM, we compute N/ē, thus locating point (2). Erecting perpendic- ulars to points (1) and (2) locates point (3), and by interpolation on the figure, the Mach number corresponding to Vmax is at once established. As we have already pointed out, presenta- tion of test data in the form of Fig. 4:9, al- though frequently a satisfactory procedure, is far from the optimum. A somewhat better method is to utilize the thrust required re- lation Now we also know that the thrust avail- able equation may be written FN f(M, NO). Thus, for small corrections A(FN/8) = where FNN AIN√O) + F FNMAM 4:27 FN 8 f(M, W/8). FNN = J(FN/S) a(N/√ē and a (FN/S) FNM am On the basis of this relation, test data may be plotted in the form of Fig. 4:10 (using a cross-plotting technique to obtain the form shown). HIGH M LOW M 0 w/8 Since Fig. 4:10 must be obtained by cross Fig. 4:9 Determination of Vmax 4:17 03/27 Increasing W/8 Lines of Constant W/8] (W/S) (W/8)2 MACH NUMBER, M Fig. 4:10 · ΔΜ plotting test data*, a test data point at the full throttle conditions obtainable during test might be at position (a) when plotted on Fig. 4:10. Since in steady level flight, thrust available and thrust required are the same thing at full throttle, point (a) represents a test thrust available point which may be cor- rected to standard conditions using 4:27. To make the correction, we proceed as follows: For the test conditions (correcting at con- stant pressure altitude), we know the value of N actually obtained. Call this value (N/√). If we presume that 100% of rated RPM is obtainable under standard conditions (which we must normally assume unless test results indicate the contrary), then we may compute the standard N/√ value using the standard atmosphere value for at test pres- sure altitude. Consequently, ^(~~) - (~) - (~~)' 이승 ​We now refer to the engine charts for the value of FNN at the test value of FN/8 and determine the increment A(FN), FNNA (ME) = Consequently, we may determine the Mach number increment as illustrated in Fig.4:10. Using this increment and the value of FNM from the engine charts, we now compute (FN)₂ = (FNM)AM. A 2 ΔΜ. This moves points (c) to (d), and we know that points (b) and (d) are points on the thrust available curve. Thus, since we have already linearized the problem, we connect points (b) and (d) by a straight line which is extended until it in- tersects the (W/8) line through point (a). This intersection, point (e), gives us the standard thrust available and also the maxi- mum Mach number for the (W/8) value con- sidered. If we correct all full throttle test points by the procedure outlined above and connect them by a single curve, this curve deter- mines the maximum Mach number for any value of (W/8). This latter correction procedure is con- sidered more accurate than the first and can, of course, be employed just as well for gross thrust data as for net thrust data. The only drawback to the procedure is its dependence on engine charts. However, since these charts are more accurate as regards slopes of curves than they are as regards actual values of the engine parameters, the proce- dure gives satisfactory results, provided the corrections are kept small. We close this first part of Chapter 4 by noting that for level flight tests of jet air- craft (including fuel consumption data), we need to obtain test information defining the following physical quantities: (1) P Ambient pressure (2) V - True airspeed This moves points (a) to (b), for the case where test conditions are warmer than stand- ard. Now at the test value of W/ §, the incre- mental increase in thrust produced by the change in N/√ would cause the airplane to speed up to point (c) which is on the constant (W/8) line through point (a) drawn parallel to the (W/8), and (W/8)2 lines as shown in the figure. *By scheduling or programming the runs according to prior computations relating pressure altitude to fuel used, it is usually possible to conduct runs at constant W/8, the success of this technique depending on the skill and cooperation of the pilot. 4:18 . (3) M - Mach number - (4) OAT - Outside ambient air tempera- ture (5) N - Engine RPM (6) W - Weight fuel flow rate Since W/S and M determine a given value of C₁, the following expressions may also be written for the case of steady level flight: Wf 8√8 = f(CL, M) 4:30 (7) Ptt - Total tailpipe pressure for net thrust determination Wf f(CL, M) 8√ M 4:31 (8) Ttt - Total tailpipe temperature for net thrust determination Standard instrumentation may be used for measurement. W{ 8√8 M W (*,M). 4:32 As regards pilot technique, the most im- portant thing in level flight testing of jets is the achievement of stabilized flight conditions for each test point. This is a more difficult problem than when piston-engine airplanes are being tested because of the more rapid change in weight (due to burning of fuel) to which the jet is subject in comparison to the piston airplane. Other than this, pilot tech- nique is essentially that described in Chap- ter 6. 4:7 RANGE AND ENDURANCE IN LEVEL FLIGHT Using the techniques of dimensional anal- ysis, it is easy to demonstrate that if Wf is the weight fuel flow rate (weight per unit time) for a fuel of constant heating value, then we have the parametric relations N Although any of the foregoing parametric representations may be used to present fuel flow data, the disadvantages of using RPM where it is unnecessary have resulted in standardization of terms of Eqs. 4:29 and 4:32. The quantities W/8√ and W/8√M in these equations are called the specific endurance and specific range parameters, respectively. Ref. 3 presents explicit equations for jet range and endurance from which we may derive definite equations for specific range and endurance. These equations are given below. specific range dR d W C1/2 = 1.414 CD CI√PS W1/2 4:33 Wf ・ +(1/ M) 4:28 dE = specific endurance Wf dw 이쁨 ​3/1/0 (MM). Wstart Wf W finish 4:39 2 3 4 5 3' 4' 5' W/8 MS√A (3) 661 (3') W₁ 8 = S.R.= (2) Wf (1) (4) W2 8: 661 = S.R.= (1) (4) 4:24 (2) ہیں W3 SPECIFIC RANGE, NAUTICAL MILES PER POUND Lines of Constant (W/8) 2 (W/S) 3 Increasing (W / 8 ) CRUISE CLIMB CONSTANT ALTITUDE CRUISE Wfinish GROSS WEIGHT, W Fig. 4:18 The Cruise Climb 4:25 Wstart (a) Sawtooth Climbs as a Means of De- termining the Speed for Best Climb The oldest flight test procedure used to determine the speed for best rate of climb at any given altitude is called the sawtooth method. This procedure is discussed thor- oughly in Chapter 6 in connection with the performance of piston-engine aircraft and the reader is referred to that chapter for details of the method. (b) Acceleration Runs as a Means of De- termining the Speed for Best Climb Another means of determining the speed for the best rate of climb is to employ level flight acceleration run data. If we assume the angle of climb at the speed for best rate of climb to be under 15°, then the power required to overcome drag in level constant altitude flight is essentially the same as the power required to overcome drag in a climb at the same airspeed.* Accordingly, the speed for maximum sep- aration of the power available and power re- quired curves in level flight will be about the same as the speed for this maximum separation in a climb, and it follows that we can obtain this speed from level flight runs. It is emphasized that all this holds true only for climb angles of less than 15°. For greater angles the assumption of lift equal to weight ceases to be realistic. Now, on the basis of the above assumption (that the flight path angle is less than 15°) we may write that the excess horsepower avail- able over and above that required to main- tain level constant altitude flight may be used either to produce a rate of change of potential or kinetic energy, i.e., Pav - Preq Pxs d(KE) = = dt + d(PE) dt Before completing our discussion of level flight range and endurance testing, it is worth- while mentioning the use of Eq. 4:28 for the presentation of range and endurance infor- mation, Within the framework of the assumptions used to develop. Eqs. 4:28 and 4:29, both should in theory provide equivalent results. However, this has not proved to be the case in practice. Indeed if one prepares curves of test data in the form of W/dē vs. N/√ē for constant Mach number values, it is found that the high and low altitude data tend to form individual curves for any given Mach number. Accordingly, the airframe para- meter relations 4:29 and 4:32 should be used in preference to the engine parameter equation (Eq. 4:28). 4:9 THE NOMINAL BEST RATE OF CLIMB AIRSPEED Specification-wise, the climb performance of an airplane is generally described in terms of its maximum sea level rate of climb, its service or operating ceiling, and the time to climb to a given altitude. Recent investigations have shown the most realistic approach to climb performance testing is obtained when one considers the time to climb to a given energy level rather than merely to a given height, since it is apparent that if an airplane arrives at an altitude at a very low airspeed an additional amount of time is required to accelerate it to fighting speed for that altitude before the airplane's mis- sion can be performed. For this reason we shall consider the topic of maximum energy climbs as well as the simpler topic of nom- inal maximum rate climbs. In the following sections we will first review the aerodynamic theory of climbs and develop the data reduction procedures, after which the special pilot techniques re- quired in climb testing will be discussed. 4:40 * Because lift and weight may be assumed equal in both cases. 4:26 where Pav Preq KE PE Thus = power available = power required kinetic energy (W/g) times (V²/2) potential energy = Wh. Pxs = W d(V)2 2g dt + W dt or, noting that Pxs 19 d(V)2 v dv dt W 2g v dvo + W dt dt 4:41 4:42 4:43 Now, as shown by Eqs. 4:40, 4:41, and 4:42, excess power may be used to produce a change of either kinetic energy, potential energy, or both. In a climb at constant true airspeed the only change is in potential en- ergy, whereas in an acceleration run the only change is in kinetic energy. Since either rate of change will be a maximum at the speed for maximum excess power, this speed may be determined either by running constant speed climbs over short altitude bands (as in the case of sawtooths) or by making ac- celeration runs from which the speed for maximum rate of change of KE may be de- termined. ? In making acceleration runs, for purposes of obtaining climb data, we are not interested in learning about the acceleration character- istics in the immediate vicinity of the stall or Vmax but rather in the region of the best climb airspeed. The approximate value of this speed may be estimated by trial ac- celeration runs during preliminary evalua- tion, from contractor's test data, or by using the estimation procedure discussed below. (c) Estimation of Speed for Best Rate of Climb In the absence of other data, estimates of the speed for best rate of climb may be obtained from the relations developed below. First consider the case of altitudes close to the ceiling of the aircraft. In this case the power available and power required curves will appear as in Fig. 4:19. From Fig. 4:19 we see that the aerody- namic speed for best range (given by the point of tangency between a line through the origin and the power required curve) is only slightly less than the speed for best rate of climb. Thus, as a first estimate to the speeds for best rate of climb at high altitudes, we may use the aerodynamic best range speeds at these altitudes (aerodynamic best range speed is the speed at which a jet airplane must fly to attain the maximum value of (c/CD), and is the speed determined by the tangent to the power required curve which passes through the origin.) The conditions determining the aerody- namic best range speed are simply T = D (Thrust H Drag) d (TV) dv d(dv) dv 4:44 (slopes of power available and power re- quired curves are the same). Assuming that dT/dV = 0, and simplify- ing, we have dD v but D = ½/pV²CDS, CL = 2W/pV²s, and CD = CDe +CL/TEAR, 4:27 POWER YBR so that v.dD = pv²C Des - POWER MAXIMUM EXCESS 4W2 pV2STEAR Solving for V4, = 0. V₁ = 4W2 CDe TEAR POWER REQUIRE POWER AVAILABLE TANGENT THROUGH ORIGIN SPEED FOR BEST RATE OF CLIMB VELOCITY, V Fig. 4:19 Power Curves at High Altitude 4:28 or VBR: derived equation for V times dD/dV, we ob- tain ༢ 2W PS. свет CDTE AR v² Vimax V4- 4W2 =0. 3 3p² V² S² πe AR CDE 4:45 4:47 (approximation to best rate of climb speed at high altitudes). At low altitudes near sea level, Eq. 4:45 predicts a best climb airspeed much lower than the actual, and therefore, at lower alti- tudes a different approximation is required. This is obtained in the following manner. The speed for maximum rate of change of energy is given by the condition that the slopes of the power required and full throttle power available curves are the same; i.e., using the second of Eqs. 4:44, we have V dT dv + T = V dD dv Using Eq. 4:45, this becomes V4_ V² Vmax VÅR - 3 and on solving for V, = 0 3 V = Vmax 6 + Vmax 36 + VBR R 3 4:48 (approximation to best rate of climb airspeed at low altitudes) For typical jet fighter or attack aircraft with aspect ratios between 3 and 5, VBR=1 3W PS and using this in Eq. 4:48 we have 2 Vfps max 6 + Vmax! 36 + ㅎ ​(3W) 2 (for low altitudes). 4:49 4:50 Since the term 1/3(3W/pS)² in Eq. 4:50 is quite small at low altitudes compared to Vimax /36, Eq. 4:50 has the simplified form at low altitudes V = Vmax. 1.732 + D. Again presuming dT/dV = 0, we obtain dD V dv + D — T = 0· 4:46 At low altitudes, the drag at Vmax of a jet airplane is due essentially to the parasite drag coefficient CD and the induced drag is entirely of second order. (We note this is not true at high altitudes.) Thus, in Eq. 4:46 T = ½ pV max CDe S. Substituting this in Eq. 4:46, together with the relation for drag and the previously 4:51 4:29 This equation, in common with Eqs. 4:45 and 4:49, predicts a speed for best rate of climb which is lower than the actual. Summarizing, we have the following: (1) At high altitudes (near the air- craft ceiling) the speed for best rate of climb will be somewhat greater than Vfps 3W PS During the run, continuous recordings of the following items should be maintained at intervals of not more than ten seconds. h'p V' · - t - Observed pressure altitude (with altimeter set to 29.92" Hg.) Observed airspeed Time OAT' Outside air temperature - RPM'- Engine RPM 4:52 M' - (for AR between 3 and 5 for fighter or attack types) (2) At low altitudes the speed for best rate of climb will be somewhat greater than V = Vmax Fuel - Mach number (if Machmeter is available) Fuel remaining A photopanel or other automatic record- ing system is most desirable for recording the data; however, it is not absolutely essen- tial, and data may be transmitted verbally via radio by the pilot to a ground recording station. At this point a few words concerning the accuracy of the data seem appropriate. If the altitude is held constant during the ac- celeration run, a smooth variation of V' vs. time should be obtained; however, the smooth-. ness of the data by itself does not indicate the absence of error. At low altitudes, the maximum accelera- tions obtained are quite high, producing a rapid rate of change of dynamic pressure which results in flow in the pitot system, and therefore, a lag in the airspeed indicator reading. Except for a minor flow in the static system due to change of position error with changing speed, all the flow is in the pitot system, and therefore, balancing the pitot and static systems against one another can- not alleviate this lag, and may actually in- crease it. Normally, the rate of change of dynamic pressure obtained during low altitude accel- eration runs is considerably greater than 1.732 4:53 The foregoing formulae determine the approximate speeds for best rate of climb at various altitudes. To apply these equations for preliminary estimation of the climb schedule, we determine the approximate speeds for best rate of climb at sea level and at the operational ceiling, and assume a linear variation of speed with altitude be- tween these speeds. (d) Range of Speeds Covered in Accel- eration Runs The starting velocity for the acceleration runs should be selected (at any altitude) as somewhat less than (at least 50 knots) the speeds given by the preliminary estimate or contractors' data, as the case may be. Dur- ing the runs it is essential that altitude be maintained as constant as possible from a speed of about 50 knots less than the start- ing speed to a speed at least 50 knots, and preferably more than this, above the esti- mated speed for best rate of climb. 4:30 that obtained in stall tests where lag is con- sidered an important factor, and it follows that lag errors in acceleration runs are a large possible source of error. These lag errors may be minimized by using a low volume pitot system with large diameter tubing; however, they cannot be eliminated altogether. Corrections for these errors may, of course, be imposed provided the lag con- stant for the pitot system is first deter- mined by ground checks. If corrections are not imposed, errors of the order of magnitude of 10 knots are pos- sible with high performance aircraft unless special low lag design systems are used. Once the data have been obtained, the first step is to convert observed pressure alti- tude to true pressure altitude, observed air- speed to true airspeed, and observed OAT to true OAT. Thus far in tests at the Naval Air Test Center, Patuxent River, no corrections have been made to account for non-standard en- gine thrust, nor for the time rate of change in engine airflow rate resulting from air- plane acceleration. The magnitude of the errors produced by omitting these correc- tions is unknown; however, the overall climb results produced by the acceleration methods have been satisfactory, so present con- clusions are that these factors produce er- rors of small magnitude. Once we have the corrected true airspeed variation with time, the points of maximum energy storage rate at constant altitude may be obtained in two fashions. 2 First, we may plot a curve of VAS vs. t, and graphically differentiate this curve to give a plot of d(VTAS )/dt. The peak of this curve represents the conditions for maximum energy storage rate. Second, since d(Vias >/ dt = 2VTAS times d(VTAS)/dt we may ob- tain this same result by plotting a curve of V TAS vs. t, and by graphical differentiation 2 TAS obtain a curve of dVTAS /dt. Using these two curves, a third plot of VTAS times dVTAS/dt is obtained whose peak point rep- resents the condition for maximum energy storage rate. In both of these procedures we are pre- suming that the weight change experienced during a particular run is negligible, and that the test weight is close enough to the standard weight so that the effects of non- standard weights are negligible. Weight changes during the run actually may be accounted for rather easily and may be important when afterburner operation is being evaluated. To account for such weight changes, recall that we are solving for the speed at which we have d(KE)/dt a maximum so that d(W/g times V2/2)dt is a maximum, With g and 2 being constants, this means that d(Wv2) dt maximum . 4:54 VBEST CLIMB Fig. 4:20 Typical Speed-Altitude Schedule as Determined by Acceleration Runs. Case of TAS Variation Illustrated. 4:31 (1) Pressure altitude, h'p (2) Time, t (3) Observed airspeed, V' (4) OAT (5) Gross weight, W (6) Tailpipe temperature (7) Thrust meter reading (if in- stalled) During the climb it may be discovered that the tailpipe temperature limits RPM, If so, the fact should carefully be noted, and an investigation initiated to determine if this limitation can be removed. Theory of Time-to-Climb Corrections When a climb to a given altitude or energy level is considered, the major factor which we want to determine is the time required to achieve the height or energy height. It is well known that should we compute time to climb on the basis of rate data ob- tained from sawtooth or acceleration runs, the time obtained would be optimistic; i.e., too small. This is true because in general the true speed for best climb rate increases with altitude, producing an increase in air- craft kinetic energy not accounted for in in- stantaneous rate data obtained, for instance, by sawtooth runs. For this reason alone, regardless of the climb schedule followed, the time to climb should be determined by a continuous climb to ceiling. Now, the climb schedule which we present to the pilot and which he will follow as closely as possible during tests is normally given in terms of calibrated airspeed versus pres- sure altitude, since these are quantities which he can observe from his cockpit instruments. However, although we present the schedule Thus, if we prepare a plot of WV² vs. t and graphically differentiate this plot, the peak on the resulting curve of d(WV²)/dt represents the conditions for maximum en- ergy storage with variable weight considered. Acceleration run data obtained at various altitudes are reduced by the methods just described. Using the reduced data, the speed for best climb at each test pressure altitude is determined and a plot similar to Fig.4:20 prepared. By flying according to the schedule of Fig. 4:20, the actual time to climb and ac- tual rates of climb may be determined. Reduction of these data to standard is discussed below. 4:10 TIME TO CLIMB FOR JET AIRCRAFT Climbs to operational ceiling (that alti- tude where the maximum rate of climb is 500 fpm) or to other ceilings in jet aircraft may be made in accordance with the schedule obtained from the sawtooth procedure, the acceleration run procedure, or according to the maximum energy schedule discussed in section 4:11 of this chapter. Both the sawtooth method and the accel- eration method should produce the same curve of nominal best climb speed* vs. alti- tude; i.e., the curve shown in Fig. 4:20. During the climb, regardless of the sched- ule followed, the following data are recorded at short intervals with a photopanel or other means. * In considering the energy climb schedule in section 4:11, we shall see that this nom- inal best climb speed actually does not cor- respond to maximum rate of climb conditions because it does not account for flight path accelerations. 4:32 in these terms, what we really seek is a schedule of true airspeed versus tapeline altitude under standard atmospheric con- ditions. If the pilot were flying in the standard atmosphere, our Vcal vs. hp schedule would provide the proper "true speed-tapeline altitude" schedule, but, since the standard atmosphere is only a statistical mean, the actual flight must deviate from that which we desire. Accordingly, we must make cor- rections both on climb rate, flight path speed, and necessarily for non-standard engine thrust. First, let us consider the nature of the speed correction: Provided the pilot follows the prescribed schedule, he will be at the correct calibrated airspeed at a given pressure altitude. There- fore, we have these two items as constants during correction (i.e., Vcal, and hp). It is shown in text books (Ref. 3, for instance) that there exists only one value of equivalent airspeed (Ve) for a given set of values of Vcal and hp and it follows at once that the airplane's equivalent airspeed is also a con- stant in the correction process, since it is the standard value. However, the true speed is not standard and accordingly, the kinetic energy content of the airplane is also not standard. In Chapter 7 it is pointed out that time to climb to a given energy height is in- sensitive to small deviations from the pre- scribed schedule, and therefore, it is pos- sible to presume that the true airspeed of the climb may be held invariant during the correction process for energy climbs. On the other hand, if we are interested in time to climb to a given tapeline altitude only with a jet-propelled airplane, changes in true speed are important, for we are in- terested only in obtainable climb rates rather than the sum of the kinetic and potential en- ergies of the aircraft. Knowing the obtain- able climb rate for standard conditions, it is a simple matter to compute standard time to climb. Thus for time to climb determinations, we have that the following quantities should be held constant during correction: hp, pressure altitude Vcal, calibrated airspeed Ve, equivalent airspeed Also, since Mach number depends only on Vcal and pressure altitude, we have constant Mach number. Now, if we hold as constants the Mach number and the equivalent airspeed, then we also have that the airplane drag is constant during correction and the increment rate of climb correction* depends only on the change in power available between test and standard conditions. This change in power available is due to two factors: (a) The increment in true airspeed (b) The increment in thrust. Thus, if RC is the rate of climb correc- tion, ARC = f(AV, AF). 4:55 Normally no correction for weight changes is imposed for correction of climb to ceiling, because if we start at the proper weight at sea level, the change in weight during the climb should, for practical purposes, simu- late the standard variation. To evaluate the quantity ARC, we first determine AV and then AT. Consider the quantity AV: If V' is the test true airspeed and V is * Dimensionless methods of analysis are also available for climb data reduction, but gen- erally require obtaining considerably more flight data than the differential method pre- sented here. 4:33 the standard or desired true airspeed, then Now, v' + AV = V. during correction, we have where AF OF (1/8) N Δ √o so that V V = داد Ve To test Ve √ostd Ρ std To std Otest P test +△(7) () Δ N The quantity aF/¿(N/√ğ) is readily ob- tainable from engine charts for the test values of (N/√@)', p and M. (Note if the charts are in terms (F/S) = ƒ(N/√☎ and M), we may write AF = a(F/8) Ə(N√√ē) (승​).8 or since pressure is held constant, the equa- tion of state (perfect gas law)gives where & has its test value.) Knowing AV and ▲F, we may write Test Excess Power Available V' (F' -D) = Pxs V Trest Tstd 'ז T Standard Excess Power Available Pxs = V(F-D) = (V'+ AV) ((F'+ AF)-D where D = airplane drag. Hence, neglecting the term AV AF, we have APxs - P'xs = V'AF + F'AV - DAV = v'AF + Av [F'-D]. Now, it is well known that in a steady climb V(F-D) RC W 4:56 If F' is the test value of thrust and F the standard value, then F' + AF = F. If thrust measuring equipment is avail- able, F' can be accurately determined, and if standard conditions of nozzle pressure ratio and total temperature are specified, standard thrust can be determined as dis- cussed in the appendix to Chapter 7. If thrust measuring equipment is not available, we may compute the changeAN/ between test and standard conditions. Then, since the Mach number is held constant 4:34 Therefore, V'(F'-D) (RC)' = W but T 1+ יז and so that F'D (RC)'W P'xs v'AF V' RC = (RC)' + W 4:57 4:59 Thus, if we assume that the acceleration change along the flight path is negligible between standard and test conditions (which it normally will be), we have and AV(RC)' W APxs V'AF+ From Eq. 4:57 (RC)' P' XS W V' 4:58 = We next must determine the relation be- tween (RC)', the test tapeline rate of climb, and the pressure rate of climb, (RC)p. After correcting the raw altimeter data for posi- tion and lag error, we may determine the pressure rate of climb, (RC)p dhp/dt from a plot of hp vs. time. During the climb a continuous record of temperature vs. height should be taken and after reducing the ob- served temperature to ambient, we have a record of T' vs. pressure altitude. From our (RC)p and T' vs. hp data, we may compute (RC)', the test value of tape- line rate of climb, using the following con- siderations: P XS P'xs + APxs W is RC = W so that RC = 1 + (RC)' APxs P'xs From Eqs. 4:58 and 4:57 APxs VAF AV + P'xs W(RC)' V' Hence, RC = 1+ AV V'AF + The equation of balance in the atmosphere dp = -pg dh and the altimeter is calibrated using this equation, presuming that p has its standard atmosphere value. Therefore, for calibra- tion - dp = pg dhp (RC)' W(RC)' whereas actually, dp = -p'g dh. On equating the two expressions for dp, we have dh = ρ' // dhp 4:35 and at constant pressure altitude dh = 'ז T dhp. On dividing by dt, we have (RC)' = (RC)p() Using this in Eq. 4:59, we find that 4:60 (6) Tailpipe total pressure for net thrust determination (7) Tailpipe total temperature for net thrust determination (8) Engine RPM (9) Test configuration RC = (RC)p √E+ V'AF W 4:61 Normally tests are not conducted under gradient wind conditions because of the dif- ficulty of actually determining the wind gra- dient. However, wind gradient corrections may be made using energy principles as discussed in Chapter 7, or preferably runs made on reciprocal headings to cancel or average wind effects. Under no circum- stances may the effects of wind gradient be neglected. The pilot technique required in conducting timed climbs in jet aircraft is essentially the same as required with piston-engine aircraft and this topic is considered in Chap- ter 6. The minimum test data required for time to climb determination are: time (1) Time (2) Pressure altitude variation wit (3) Ambient temperature variation with time (4) Calibrated airspeed variation with time In addition to recording the foregoing data, the pilot should record the existence of operating limitations such as limited RPM due to excessive tailpipe temperature and the like. Corrections for instrument error, posi- tion error, etc., are made according to the standard procedures considered elsewhere in the manual and corrections to standard are made using Eq. 4:61. The actual time to climb is determined by plotting 1/RC vs. STANDARD ALTITUDE, h or H ABSOLUTE CEILING (BY EXTRAPOLATION) OPERATIONAL CEILING TIME TO CLIMB RATE OF CLIMB (5) Fuel flow data (to determine the variation of weight with time or pressure altitude during the climb) STANDARD TIME TO CLIMB, t STANDARD RATE OF CLIMB, (RC) Fig. 4:21 4:36 altitude and then graphically evaluating the integral. h dh t = time to climb to altitude h = (RC) 4:62 climbs is considered in detail in Chapter 7 and here we shall only briefly discuss the theory of the energy climb. As shown in Chapter 7, the problem to solve is the maxi- mization of the quantity w as a function of time where Final data are presented in the form of Fig. 4:21. dhe V² W = 1 and he (energy height) = h + 2g dt ? 4:11 DETERMINATION OF MAXIMUM ENERGY STORAGE SCHEDULE AND TIME TO CLIMB TO A GIVEN ENERGY HEIGHT As we have mentioned, the nominal best rate of climb schedule does not generally produce the minimum time to climb to a given geometric altitude, and certainly does not produce the minimum time to climb to a given energy height. This is so because two factors are not considered in the con- ventional climb analysis: which is the mechanical energy content of the airplane per unit weight. The analysis in Chapter 7 reveals that one way to express the condition for maxi- mum energy storage is the equation dw aw ah a (2) 2g (1) The effects of changing flight path speed and, hence, kinetic energy with height. (2) The possibility of storing more energy at lower altitudes where the separa- tion of the power required and power avail- able curves is a maximum. When conducting a timed climb to ceiling, the effects of changing kinetic energy are automatically accounted for by the tests and accordingly, the rate of climb actually achieved is less than predicted by either sawtooth climbs or acceleration runs. By a more precise analysis of either sawtooth or acceleration run data, it is possible to pre- dict actual rates of climb with fair accuracy by accounting for kinetic energy changes with altitude. However, the most satisfactory procedure is always to determine time to climb by conducting a climb to ceiling ac- cording to the desired schedule. The general topic of maximum energy 4:37 4:63 Accordingly,* if we prepare a plot of h vs. V²/2g for lines of constant w, we shall have aw/ǝh = dw/d (V²/2g) at points on the w = constant curves where the slopes of the tangents to the curves are 135°. This is the method of determining the maximum energy climb schedule on the basis of acceleration runs. The optimum schedule may also be de- termined from continued climb to ceiling made according to several different sched- ules. In this case, we plot he vs. time and from this set of curves prepare cross-plots of w vs. V for constant values of he. From this second set of plots, we may readily de- termine the value of V for maximum w, and knowing he, we may determine the schedule V = (h). The procedure is fully described in Chapter 7. At the NATC, the maximum energy sched- ule has, in the past, been obtained by con- ducting acceleration runs and then analyzing * See section 4:12. the data to determine the optimum energy storage speeds. The analysis procedure ac- tually used has been to solve for the climb schedule by an iterative method; however, since a direct solution is available, this will be considered here. To obtain the raw data, acceleration runs are made at a number of altitudes (at least four) and data are recorded as discussed in section 4:9. Although weight corrections are not normally imposed on acceleration run data, as noted in section 4:9, it is possible to account for weight variations during the runs. Where an excessive time rate of weight change is encountered (as is possible with afterburner-equipped aircraft), accel- eration runs should not be used since it is not possible to predict a priori what weight will exist at a given altitude. In cases such as this, the energy schedule should be ob- tained using the method of continued climbs discussed in Chapter 7. Acceleration runs should be conducted at constant altitude; however, the effects of minor changes in height during the run may readily be accounted for by correcting the observed kinetic energy variation with speed for any small potential energy changes en- countered. After the observed data have been reduced to the form of true airspeed vs. time, the derivative w = dhe/dt is readily obtained by graphical differentiation of the velocity-time curve (note that at constant altitude dhe/dt 1/2g times d(V²)/dt = V/g times dV/dt). = data to obtain the plot (Fig. 7:8) of h vs. V²/2g, and drawing the 135° tangents, deter- mine the variation of V²/2g vs. h. These data are reduced to the form of true speed variation with altitude using obvious pro- cedures. Our next problem is the determination of time to climb according to the schedule ob- tained above. The timed energy climb is conducted in the same fashion as any timed climb and the same data are recorded. This topic is discussed in section 4:10. However, the corrections to impose on the energy climb are not the same as those required for the simple climb to altitude discussed in section 4:10. Because of necessity, the pilot follows a schedule of Vcal vs. altitude and the only quantities which he can correctly maintain during the climb are: W - (weight) (we assume this to be standard) Vcal (calibrated airspeed) hp - Ve - M D (pressure altitude, and hence, am- bient pressure) (equivalent airspeed) - (Mach number) · (airplane drag) The non-standard quantities are: T V - F as discussed in section 4:10 - (ambient air temperature) - (true airspeed) (engine thrust) Atmospheric turbulence When determining time to climb to a given geometric altitude (see section 4:10), cor- rections were imposed to account for the Because tests are run at constant pres- sure altitude and because no corrections are normally made for non-standard temperature or non-standard thrust, the pressure alti- tude may be considered the standard altitude of the test. We then obtain a plot of w vs. V (or V²/2g) for various constant standard altitudes by plotting the w vs. V curves for the test pressure altitudes as shown in Fig. 7:6 of Chapter 7. Using Fig. 7:6, we next cross-plot the 4:38 variation between the true airspeed of the test and the standard atmosphere true air- speed called for by the climb schedule; i.e., corrections were made only to the climb rate. Here we shall impose our correc- tions on the rate of energy climb which will simultaneously account for changes in flight path speed and in rate of climb. The correction procedure will be developed pre- suming that the calibrated airspeed-pres- sure altitude schedule is followed by the pilot as accurately as is operationally pos- sible according to the stipulations of sec- tion 6:10 of Chapter 6. The corrections for the case where it is assumed that the pilot maintains the proper true airspeed-pressure altitude relation are considered in the appendix to Chapter 7. At this point, we shall proceed on the assumption that the proper calibrated airspeed is main- tained. Under these circumstances we have WW = V(F-D) w' W = V'(F'-D) where the notation is as in section 4:10. Be- cause the equivalent airspeed and Mach num- ber are standard, D is unaltered during correction. Hence, V = V' + Av F = F' + AF and neglecting the product AF AV, we have on dividing wW by w'W AV AF 1+ + F'-D also that so that w'W F'-D = V' W = 'w' T V'AF + W יז 4:64 As mentioned previously, it is undesirable to conduct tests under gradient wind con- ditions although corrections can be imposed if the gradients are known. As shown in the appendix to Chapter 7, the wind correction for a horizontal wind with vertical gradient (dV/dh)wind is dw = dh anwind dt To determine the quantity w', we must determine h', the true altitude obtained up to the time t. For this purpose we utilize the relation developed in section 4:10, dh = Idhp and prepare a plot of T'/T vs. hp as obtained from the test data. This plot is as shown in Fig. 4:22. Then hpi 푸 ​ni₁ = s = dhp. ! 4:65 We next reduce the test calibrated air- speed to true airspeed by conventional means (which is easy to do since V= f(Vcal, P, T) and all quantities on the right side are known). Then Now we know that AV 1+ 4:39 ne = n'+ 2g T dhp we have Via he = he + (hp-h') + 2g (-1). 4:66 Knowing w and he, from the relation we have dhe W dt dt = dhe W про or hp PRESSURE ALTITUDE Fig. 4:22 We next prepare a plot of h'e vs. time and by graphical differentiation or other means obtain w' dh'e. dt W and AF are obtained from fuel flow and engine data, and substituting in Eq. 4:64, we obtain w, the corrected rate of change of energy. The corrected energy height is ob- tained from the following considerations: The test energy height is he = h'+ (V² /2g), and since pressure altitude is used as a con- stant during correction he = and on subtracting V² hp + 2g · + 29 (V² - Vi². hehe hp-h' + Now, since √ t = ne dhe. W V = V' Hence, if we plot the quantity 1/w as a function of he, the standard time to climb to any given energy height is readily obtainable. 4:12 PROOF OF CONSTRUCTION USED TO DETERMINE OPTIMUM ENERGY CLIMB SPEED FROM ACCELERA- TION RUN DATA The condition for maximum "nominal" rate of climb at a given constant altitude is that the climb should occur at the speed for maximum rate of change of kinetic energy at that altitude. Consequently, if we plot a curve of d(V²/2g)/dt as obtained from accel- eration run data vs. V or V²/2g, the speed for "nominal" best rate of climb is deter- mined by drawing a horizontal tangent to our curve. As is quite apparent, this construction locates the speed for maximum separation of the power required and available curves and would actually give the speed for best rate of climb at the given altitude were it 4:40 not for the fact that in the actual climb the true flight path speed is changing along with the altitude. Since the time to climb to an energy height is given by t = s dhe W in comparison to the conventional time to climb formula s dh RC it is apparent that what we want is the maxi- mum value of w at any given he. Therefore, if we could run tests at con- stant energy height, a plot of w vs. V or v²/2g obtained at constant he could be used to obtain the speed for best energy climb. In this case, we would merely have to draw a horizontal tangent again to the new curve of w vs. V²/2g to determine the speed. Ac- tually a curve of w vs. V²/2g at constant he can be constructed from acceleration run data; it is simpler, however, to make use of the procedures of Chapter 7. Since w = f(h, v²/2g), we could find the maximum of w at any given fixed value of energy height by plotting w vs. h for fixed he and by then locating the height for the horizontal tangent. The conditions for maximizing w at any given value of he are that aw = O at constant he ah v² 2g = O at constant he⚫ Either operation yields the same result, namely 35 / 53 11 a aw 2 V² 2g The foregoing equation defines any line in a plane perpendicular to the h, V²/2g plane and including the line h = V²/2g (presuming an orthogonal set of coordinates consisting of w, h and V²/2g). Since we are seeking to maximize w at constant he, we also know that the other con- ditions required are that aw/a (V/2g) = 0 at constant he and aw/ah = 0 at constant he. These latter two conditions amount to the same thing, since in the first case we are considering the projection of a curve of w at constant he in the w, V²/2g plane and, in the second, the projection of this same curve on the w, h plane. Thus, regardless of how we project the curve of w at constant he, at the maximum point, we have a tangent parallel to planes of constant w. If we consider the surface defined by w = f(V²/2g, h), the planes of he = constant have the equations h+ (V²/2g) = constant and, therefore, the tangents to the curves of w at constant he lie in planes of constant w along lines of h + (V²/2g) con- = stant. Now, therefore, if we plot our acceleration run data as indicated in section 4:11, we shall come up with curves of constant w projected in the h, v²/2g plane and the tangents, which maximize w at constant he, will appear as lines having slopes of 135° as shown in Fig. 7:8 of Chapter 7. This completes our discussion of energy climb theory for this chapter. For a more complete analysis, the reader is referred to Chapter 7. 4:41 4:13 DESCENT PERFORMANCE OF TURBOJET AIRCRAFT The descent data reduction process theory for turbojets is exactly the same as that for the climb data; however, at high rates of descent, consideration can be given to omit- ting small corrections such as the thrust correction made for temperature variation from standard and fuel consumption weight corrections. Tactical employment of a particular air- craft usually defines the type of descent nec- essary. Such controlling items as engine RPM, Mach number, airspeed, fuel pressure, and configuration have to be scheduled with altitude before proper evaluation can take place. Thus, determination of a descent schedule is much more difficult (and im- portant) than the reduction of descent data once they are taken. In general, the maximum range flight for jet aircraft includes cruise or climb-cruise at high altitude and a high speed descent at minimum power and/or minimum fuel con- sumption. 4:14 MEASUREMENT AND CORRECTION OF TURBOJET TAKE-OFF AND LANDING CHARACTERISTICS Take-off and landing performance of tur- bojet aircraft is governed by the same fac- tors which govern this performance of pro- peller-driven aircraft. The equations used by the NATC to re- duce take-off data to standard are developed in Chapter 6 (section 6:11) and need not be given here. Rather, at this point we shall briefly review a few considerations peculiar to jet-propelled aircraft. When dealing with reciprocating-engine aircraft, accurate estimation of thrust is quite difficult; however, when thrust meas- uring equipment is available for jets, it is a simple matter to determine the test values of thrust developed. Thus, thrust corrections • can generally be made quite accurately when we deal with jet aircraft. Moreover, if we do not have thrust measuring apparatus aboard the airplane, the increment between measured static thrust and standard static thrust may be used for correction purposes without generally introducing serious errors. Since the power developed by a jet air- plane depends on its speed, it is sometimes inadvisable to take off at speeds close to the stall for the reason that insufficient power may be available for climb out immediately after the aircraft has moved out of ground effect. Several serious accidents have oc- curred due to take-offs at low airspeeds with marginally powered aircraft. The optimum take-off speed at a given gross weight should be determined by a series of trial runs start- ing at a speed considerably greater than the stall speed. In correcting data for take-offs over ob- stacles, it is of help to have a knowledge of the airplane's lift-drag relation. As ex- plained earlier in this chapter, provided thrust measuring equipment is available, this relation may be obtained from level flight tests of jet aircraft. The landing run of turbojet aircraft is generally lengthened by the effects of residual thrust produced by the engines at idle RPM. Since this residual thrust is frequently ap- preciable, the use of such devices as jet de- flectors, drogue chutes and the like must be resorted to when minimum landing distances are sought. Generally, the braking systems of modern aircraft are not designed to dis- sipate the heat produced by continued full use of the brakes during the landing run and this factor must be considered when tests for minimum landing run distance are con- ducted. For further information on the take-off and landing tests, the reader is referred to sections 6:11 and 6:12 of Chapter 6 and to the detailed analysis of Chapter 8. 4:42 - 4:15 TURNING PERFORMANCE OF TURBOJET AIRCRAFT The turning performance of any airplane is limited aerodynamically by several fac- tors, namely, the thrust or power available, the drag characteristics, the maximum lift coefficient available and the controllability. In this discussion we shall limit our thinking principally to the effects of these factors on level (constant altitude) turns. We shall consider first the combination of thrust and drag characteristics which collectively pro- vide what is called the thrust limitation. and for constant wing area we have D = 8M² f(M,a, Ni√ē) Sm² f¡(M, CL, FN/8) 4:69* Similarly, the quantity nW, being the sum of the thrust component normal to the flight path and the lift, may be written nW = 8M² f₂ (a,M) +8f3 (M, N/√) sin a (a) Thrust Limitation Consider an airplane in a level turn pro- ducing a normal load factor "n" and accel- erating along the flight path with an accel- eration "a". If we sum the flight path forces, we have ola and from Eq. 4:70 f4(nW/8,M,N/√@). 4:70** 4:71 Substituting Eq. 4:71 in Eq. 4:69, we have FN cos a W D D = pM² f5 (M, nW/d, N/√ē) W 4:67 4:72 where FN = net thrust also FN cos a = 8f6(M,nW/8,N/√). Using Eqs. 4:72 and 4:73 in Eq. 4:67 8 == f (M, nw/8, N/√8) g W 4:73 a = angle of attack of the thrust line D = airplane drag W gross weight If we wish to consider the effects of en- gine spillage flow on airplane drag, then in general, f(M, a N/√8) CD= f(M, CL, NAB) f(M, CL, FN/8) 4:68 4:43 & f5 (M, nw/8, N/√ē), 4:74 * Where the terms N/√ or FN/8 account for the effects of entrance airflow on drag. **The last term being the component of thrust normal to the flight path. hence on dividing by n and placing 8/nW in- side the function ℗(nW/8, M, N/√ē). ng 4:75 This equation, of course, applies when- ever there is an acceleration along the flight path (in a turn or otherwise). Thus, if we consider a steady climb (n = 1), we have (for small climb angles) y = = = (W/8, M, N/√Ƒ). g 4:76 In developing the foregoing relations, the following assumptions, which are generally accepted in estimating jet turning perform- ance, have been made, namely: (a) Reynolds number effects are negli- gible both for the engine and airplane. (b) Effects of absolute pressure on the engine are negligible (i.e., the effects of ab- solute pressure on combustion are disre- garded). (c) The drag coefficient in turning flight is identical with that obtained in level unac- celerated flight at the same angle of attack and Mach number (i.e., there are no effects due to varying load distribution, local ve- locity variations and the like). In cases where these assumptions hold, we have developed a simple relation for turning performance. In particular, for the case of the steady turn a = 0, we have 0 = $(nW/8, M, N/√ē ) or 8 f7 (M, N/√8). W From Eq. 4:77 we see that at constant W, M and N/√ (or FÑ/8 as a replacement for N/√) the normal load factor n is a di- rect function of the pressure ratio 8; thus, steady level turning performance at one al- titude may be simply related to the perform- ance at any other altitude. For instance, if we consider that steady level flight repre- sents a one g turn (infinite radius), then a plot of maximum level flight Mach number vs. altitude or pressure ratio may be used to determine the maximum turning accel- eration in level flight at various altitudes as shown in Fig. 4:23. Tests at the A. and A.E.E. have shown that the generalization of data predicted by the equations is not always obtained, presum- ably due either to the effects of Reynolds number or of absolute pressure on the air- plane and engine performance. Nonetheless, because of the great effort involved in mak- ing complete turning performance tests di- rectly at all altitudes, it is considered well worthwhile to use the process described in an attempt to generalize turning data. As previously demonstrated in this chap- ter, a function f(M, N/√) is equivalent to f(M, FN/8); consequently Eq. 4:75 may be written a nW 8 M, FN 8 ng 4:78 The use of thrust rather than N, may well avoid the difficulties mentioned above, since it is hard to imagine large and sudden varia- tions of aircraft drag being produced by Rey- nolds number effects and Eq. 4:78 is correct excepting for the omission of these effects. (b) Lift Limitation If we analyze the CLmax or lift boundary of an airplane, arguments similar to those presented above lead to the relations nLW ♡ = f(M, N/√8) 4:77 4:79 4:44 or ոլ W 8 f(M, FN/8) 4:79 according to whether we use N/√ or FN/8 to define the engine operating conditions. These equations allow for the upward com- ponent of the jet thrust and again permit a generalization of data obtained under a vari- ety of altitude conditions. Clearly, these equations should be equally applicable to buffet boundaries as well as the true CL, or lift boundary. max However, generalization cannot be ex- pected in cases where the lift limitation is 1 Maximum Level Flight Mach Number (n = 1) 8, (Test Curve) +th +18 Mmax for n=2 Obtained from n=1 Data 28, = 82 M Mmax at 8, for n=l Mmax at 8₂ for n=2 ni 8 n2 82 Fig. 4:23 For Contant M and N/√ ; 4:45 set, say by instability or lack of control. In such cases, additional factors including nat- urally the c.g. position would need to be in- troduced. (c) General Comments It may be noticed that we have presented the foregoing work in terms of the load fac- tor “n” rather than in terms of actual turn- ing radius. This was done because the simplest relations are obtained in this way. It can be shown, however, that the steady level flight radius of turn is given by are of course interchangeable and it is not necessary to go into further detail here. Correction methods for turning perform- ance can readily be developed on the basis of the foregoing relationships. One simple procedure is to conduct level acceleration tests at measured thrust values and then correct the measured acceleration for the increment of thrust between the measured and standard thrust using the simplest of expressions Aa = g • ΔΕ W 8 f(M, W/d, N/√ē) or f(M,W/8,FN/8) 4:80 which follows at once from the fact that in a steady level turn (constant) TM2 R = 9 n² - √n² -1. 4:16 CONCLUDING REMARKS In this chapter the general base of the analysis of jet aircraft performance has been considered and the undamentals of the data reduction procedures required for these air- craft have been presented. The methods described are essentially those in use at the NATC at the time of this writing; however, the procedures for the re- duction of energy climb data have been pre- sented in a modified form to take into account a simplified data analysis procedure de- scribed by Mr. Davy in Chapter 7. where T = absolute temperature. Thus, using Eq. 4:77 for n, Eq. 4:80 is readily obtained. It is similarly possible to obtain expres- sions for the climbing or descending turns which involve the climb or descent angle y. Flight path acceleration and flight path angle Special topics, such as carrier suitability testing and the like, have not been considered; however, it is hoped to cover these items in future volumes. REFERENCES 1. Hesse, W. J., "A Simple Gross Thrust Meter Installation Suitable for Indicating Turbo- jet Engine Gross Thrust in Flight," TPT, TR 2-52, NATC, Patuxent River, Maryland. 2. Hesse, W. J., "Determination of Level Flight Thrust Available Curves of Turbojet Powered Aircraft," TPT, TR 1-53, NATC, Patuxent River, Maryland. 3. Dommasch, D. O., Sherby, S. S.,and Connolly, T. F., "Airplane Aerodynamics," Pitman, 1951. 4. Dommasch, D. O., (Editor), "Flight Test Manual, Part I," Revised Edition, Preliminary Copy, Patuxent NATC, August, 1953. 4:46 AGARD FLIGHT TEST MANUAL VOLUME I, CHAPTER 5 PERFORMANCE REDUCTION METHODS FOR TURBO-PROPELLER AIRCRAFT By Kenneth J. Lush AFFTC, United States Air Force and John K. Moakes Aeroplane & Armament Experimental Establishment United Kingdom VOLUME I, CHAPTER 5 CHAPTER CONTENTS Page SUMMARY FOREWORD 5:1 INTRODUCTION 5:1 5:2 DIFFERENTIAL REDUCTION METHODS 5:1 5:3 THE BASIC PERFORMANCE EQUATIONS 5:2 Level Flight 5:5 5:4 FUEL FLOW IN LEVEL FLIGHT 5:7 5:5 EQUATIONS FOR CLIMB REDUCTION 5:9 (a) Reduction Process 5:10 (b) Pressure Height and Pressure Rate of Climb 5:10 (c) Longitudinal Acceleration 5:11 5:6 DATA REQUIRED IN ADDITION TO THE BASIC PERFORMANCE QUANTITIES 5:11 5:7 REQUIRED ACCURACY OF THE COEFFICIENTS 5:13 5:8 POSSIBLE SIMPLIFICATIONS OF THE METHOD 5:14 5:9 REDUCTION OF LEVEL SPEEDS AND FUEL CONSUMPTIONS BY PERFORMANCE ANALYSIS 5:14 (a) Introduction 5:14 (b) The Approach Considered 5:14 (c) Standard Power 5:15 5:10 CONCLUDING REMARKS 5:16 REFERENCES 5:16 TABLE I 5:17 TABLE II 5:18 APPENDIX - USE OF LOGARITHMIC DIFFERENTIALS IN PERFORMANCE REDUCTION $ TERMINOLOGY b Wing Span Cp Propeller Power Coefficient = P/pN3d5 CDe Effective Parasite Drag Coefficient D Drag d Propeller Diameter e NACA Efficiency Factor Vemp AB3 + 4 (the f(Ky) = Ky4 + 3 Ky 271 (Kv3 + 3/KV) 2.23./W(LD)max Dmin 2./3 Kv2 D KyA + 3 f1(Ky) 6 f2) = Ky 12(ky) = 1/2 (3 - par ) - Kvents - (3 FKV JKY Ky+ 3 h Rate of Increase of Height hd DV/W = Rate of Climb Equivalent of Drag Power hp Pressure Height We 'F Fuel Flow J Propeller Advance Ratio V/Nd Ky Veremp C L Lift N Propeller Rate of Rotation (RPM) P Shaft Power XVe r = Pvto+ XVe S Gross Wing Area T Ambient Air Temperature OK TA Jet Pipe Temperature OK V True Airspeed TERMINOLOGY (Continued) Ve Equivalent Airspeed = V po Vemp Equivalent Airspeed for Minimum Drag Power W Aircraft Weight X Net Exhaust Thrust Cp n δη аСр Jan B JIT 7 OJ Propeller Efficiency (Propulsive) s 6 Density Ratio (Relative Density) pipo P Air Density Air Density at Sea Level Standard Temperature and Pressure Po Air Pressure р Suffix S Standard Value T Test Value SUMMARY behavior. Thus, the methods are flexible at the expense of simplicity in application. Reduction methods for level speed, climb and fuel consumption tests are presented, which are, in the absence of compressibility effects, suitable for use with turbo-propeller aircraft at the present stage. The methods re- quire individual treatment of the particular engine installation, inasmuch as it is not prac- ticable at present to generalize about engine It is anticipated that it will eventually prove possible to simplify the methods when experience has been gained on these engines, and the design of their control systems has been stabilized. EDITOR'S FOREWORD accomplish fairly complex interacting con- trol functions. As far as airframe characteristics are concerned, turboprop airplanes are similar to other propeller-driven airplanes. More- over, at the present time, turboprops are generally designed to fly at speeds lower than drag divergence so that the assumption of a lift-drag relationship independent of Mach number may be made in analyzing their performance. As noted in the body of this chapter, un- til engine design has been more or less stabilized, it is not safe to assume that any of the existing variables may be neglected and, therefore, the analysis presented here is rather complex since all variables of pos- sible importance are considered. On the other hand, the testing of turbo- prop aircraft is complicated by the fact that only very limited experience has been gained in testing these machines and, moreover, by the fact that power is produced both by a propeller and by direct jet thrust. Thus, the determination of both power required and power available is not as simple a matter as with piston engine aircraft. Similarly, it is not easily possible to determine stand- ard fuel consumption characteristics because of the large number of variables which must theoretically be accounted for in data anal- ysis. Unlike other chapters of this volume, which deal with fairly well developed pro- cedures, this chapter describes what must be considered (as of this time) to be experi- mental techniques. This fact should be taken into account in test planning and it should not be anticipated that the turboprop test program can be conducted in a routine man- ner with much hope of collecting standardized data, The possibility of obtaining really repeat- able data from flight tests of turboprop air- planes depends in part on the design of the engine and propeller controls and governing systems. Simple mechani Simple mechanical linkages with easily predictable characteristics tend to provide better data than servo systems which The testing of airplanes with mixed-type power plant installations (combination of reciprocating engine with turbojets) presents problems quite similar to those encountered in the testing of turboprop airplanes. Thus, although the testing of mixed power plant aircraft is not considered in detail in this volume, it is noted that the differential an- alysis techniques presented in this chapter are adaptable to this work. 5:1 INTRODUCTION are convenient where level speed perform- ance data are required over a large range of the variables. It has been shown that for many turbo- propeller installations, the application of experimental methods of performance reduc- tion, such as those used for turbojet aircraft (which require no advance numerical data) will be either impracticable or impossible. Alternative analytical methods which are available are: (1) "Differential methods”, based on linearized relationships between the varia- bles, suitable for small corrections to ob- served data. Here the performance items per se, viz., level speed or climb points, are corrected for deviations from standard of such parameters as aircraft weight or air temperature. The common difficulty in the application of either of these methods to turbo-propeller aircraft is the lack of knowledge about the response of the engines to changes in air temperature and forward speed and the associated difficulty in the prediction of the power which would be available under stan- dard conditions. This will only be solved when considerable experience is gained with these engines and their performance charac- teristics can be generalized. Until then each installation must be considered on the basis of its own characteristics, which in most cases must be established prior to or during the measurement of the primary aircraft performance quantities. The method requires advance numerical data regarding the behavior of the airframe as well as the propulsive unit, perferably some fund of generalized data in a conven- ient form. The information need not be very precise inasmuch as it is only used to estimate relatively small corrections, In the paragraphs which follow, a differ- ential reduction process is developed which is sufficiently flexible to be adaptable to any turbo-propeller installation, provided that compressibility effects on airframe drag and propeller efficiency are not present. The methods are convenient for the cor- rection of isolated performance points where compressibility effects are not present, and at the present level of performance are thus quite appropriate to turbo-propeller aircraft. The developments are presented in detail to illustrate the principles involved (which have been current practice in the U.K, for the past ten years for piston-engined per- formance reduction) and to indicate the most likely direction in which simplification, based upon generalization of engine charac- teristics, will take place. (2) “Performance analyses'', where observed or estimated jet thrusts and shaft powers combined with estimated propeller efficiencies give estimated power curves. From prior information about the perform- ance of the propulsive unit, and its response to change in air temperature, estimates of thrust power which would be available under standard conditions then lead to standard performance. Also given is a brief outline of an approach based on the performance analysis of method (2), from which could be developed a routine similar to that proposed in this volume for piston engine level speed reduction. 5:2 DIFFERENTIAL PERFORMANCE REDUCTION METHODS Prior information about airframe drag is not required inasmuch as drag data are obtained during the analysis. Such methods In developing the present differential methods, it was decided that because devel- opment of turbo-propeller engines and their 5:1 control systems is in its early stages, it was not possible to provide the degree of generalization concerning engine behavior which is at present practiced in similar reduction methods applied to reciprocating engines (Refs. 1 and 2) and it would be preferable to provide for the insertion of more engine parameters, and if necessary propeller parameters particular to the type under test. on the power and fuel flow of turbo-propeller engines. Inclusion of this effect in the reduction process presents difficulty; it is, therefore, only included in the correction of level speeds at particular ratings, e.g., at maximum continuous power. The correction process for fuel consumption tests at lower powers avoids changes in ram by making corrections to fuel flow at constant equiva- lent airspeed. It is assumed throughout that compressi- bility effects on airframe drag and propeller efficiency may be ignored. To a certain extent the test program can be fitted to the reduction process. If the tests are carried out at the required air pressure, corrections need only be made for the effects of the differences in air temper- ature, engine speed and aircraft weight from their standard values. 5:3 THE BASIC PERFORMANCE EQUATIONS If we use consistent units and neglect longitudinal acceleration, we have from the law of energy conservation In determining the performance with the engines operating at some limiting output, e.g. “maximum level speed” or “climb power", some difficulty may arise due to a change of these operative limitations with a change in air temperature. For example, at low air temperatures the engine may operate at maximum possible engine speed but with the jet temperature below the maxi- mum and vice versa at high air temperatures. OPVO + XVe - Dve + Whio : 5:1 Editor's insert: If we suppose that the lift drag polar can be approximated by the relation CL CO : . Сое + TEAR To handle this difficulty, only the “external features of the engine performance (i.e., shaft power, thrust and engine (or propeller) speed) are inserted in the performance equa- tions, their variations with changes in air temperature being considered separately from their effect on performance. In the case of fuel consumption tests, the engine will be operating frequently within the limi- tations and this difficulty will not arise. where Coe Effective parasite drag coef- ficient e : Span efficiency factor AR : Aspect ratio, then we have The effects of weight changes are consid- ered in deriving the equations for level flight and fuel flow, but in the climb the rate of loss of weight with increase of height is insensitive to air temperature. Therefore, no correction for weight has been applied. If these are required, however, the methods of Ref. 1 or 2 may be adopted. DECOS:9CDES cias + TAR Changes in ram may have a marked effect 5:2 - and, since q= Ź pove (see Chapter 1) and ci = W/?S? , we have q 2 Po In terms of Vemp the equation for Dve may be written ? BW D - Ave + Ve BVemp : emp DVE - AV . + W?. vemo Ve Vemp 5:2 where A : CDEP.5/2 or B : 2/P. STEAR BW? 1 Dve . AKÝ.vemp + + so that KV Vemp BW Dve : AVO+ Ve + BV Kuvamo [axi. vemp + Bw') ] 5:3 We next introduce a dimensionless veloc- ity, Ky, defined as but BW? ку 2 Ve Vemp Vemp : ЗА 5:4 so that, after simplification where Vemp is the equivalent airspeed cor- responding to minimum airplane power re- quired. Vemp is readily found in terms of the parameters A and B from the follow- ing considerations: BW? Dve Kỳ +3 ку 3 Vemp The power required is DV : (avo + ) v. avě to + BW2 + Ve or 8 BW? To Ve BW' "2 Kỷ +3 For minimum power Dve W 1372 3 Vemp ку 5:6 (DV) a Ve 1:05 = 0 = 3AV2 vo BW? Jo ve Now substituting for Vemp from Eq. 5:5, we obtain whence BW Dve AB' Ky 3 (209)* ( ** + 3). vieme WVz 27 3A Ку 5:5 5:7 5:3 and Eq. 5:1 may be written Another useful alternate form of Eq. 5:6 or 5:7 is obtained by writing Pro + x Ve = W3/2 f(KV) + who, BW1/2 / BWV2 Vemp 3 vemp 5:10 3 Vemp 2 8 BW2 where D = drag - Ave + va i so that from Eq. 5:5 2 BWV2 CDEPOS A : 2 ; B A12B1/2 Veme 3 W : Po STAR 3 Vemp P: shaft power 5:8 X : exhaust net thrust Now for maximum C/Co we have ń : rate of climb Kv: Ve /emp d 0 O CL . CL CL CDe+ Te AR ( Vemp equivalent air speed for minimum power required (22"). (+3) : whence at maximum C/Co AB f(KV) = 4 27 KV (Kỳ+3 23 W (L/D)max ку Vemp Ľ CDe . TTAR and If the air pressure is maintained constant during the reduction process, we also have CDeTTAR ( CL co Imax : 2 CDe do dT . T 글 ​e 1 NEAR 2 JCDe 2 AV281/2 Consequently, Eq. 5:6 becomes 5:11 where T is the ambient air temperature in °K or °R. Kº +39 Dve - Vemp CL 23 W W 3/2 Menomax KV If there are no compressibility effects on propeller efficiency, this efficiency is a function of CP and J only; hence, using 5:9 5:4 logarithmic differentials* where N is the propeller rotational speed. du don η . CP on dCp ПдСР СР . + J.on. η δυ zlo Also, as Vemp is, in the absence of compressibility effects on drag, proportional to VW (see Eq. 5:5), from the definition of Ky, we have dP 30N Cp on тәСР { dT + T } P N dky | dw - dve Ve +g 2 W j.on nou { dve I dTdN) + Ve 2 T N ку 5:12 5:13 Level Flight An expression for the change in equivalent airspeed which will result from changes in weight, air temperature, shaft power and thrust is obtained from Eq. 5:10 with h = 0; thus, dn dP | DT ηΡσ novo +X Ve + 액 ​XVe + nPro +xve { dx dve + х Ve } 3 0W Ky Of(ky) dky + f(ку дку KV η p 2 T 2 W 1 1 1 Using Eq. 5:1, this may be written more neatly 1 1 3 dW (-1){432 * am ) } ку dP + P I OT 2 T - x sax X d Ve + + DUX of(kv) dk v ку Ve 2 W f(KV) дку From Eq. 5:13 the right hand side becomes | W { Ку f(KV) afiky) ку (} + + aky f(кудку ) ! 3 of(KV) dve 2 Ve 5:14 Substituting for the propeller efficiency from Eq. 5:12 we have (denoting Cp/n times ondo Cp and J/n times on/DJ byaand Brespectively) X SOP :ll X 10X + + D х 6Ne) | - Ve Ve {la ce + a)+ 4(a+B/2-) °F (2+8/2- ) - (32+B+B over 뿌 ​IN (3a+81+8 om du er en av ! OW d 1 2 Kv. Of(KV) f(ky) aky 3 . over ove story of (y) Ку fKV) W Ve дку 5:15 * The use of logarithmic differentials is discussed in the appendix to this chapter. 5:5 In the above equation, dP and dx are the total changes in power and thrust, inclu- ding ram effects. Application of the equation is simplified if the ram effects are separated from the effects of changes in air temperature or engine speed. Assuming as before that ci CD : CDe + TEAR 8 we have that : BW2 D = Ave + ve and for minimum drag Let us rewrite Eq. 5:15 in the approximate form for finite changes and write AP & AX for the changes in P and X which would result at constant equivalent airspeed Ve from the excesses AT and AN of the test air temperature and engine speed over stand- ard. Following Ref. 2, we take AN, AT, and AW as "errors" in the test conditions, that is, the amounts by which the test values of N, T and W exceed the standard or desired values Ns, Ts and Ws and Ave as the corresponding correction to the observed equivalent airspeed Ve. We replace dP/P and dX/X by 이 ​aD ду : 0:2 Ave - 2 BW? vel hence vemd Equivalent speed for mini- mum drag = Bw / A and Dmin : 242 B" W. Also from Eq. 5:6 dP AP - Ve OP Ave pove Ve P Kit + 3 D: BW? 3 Vemp Ve BW KV+3 ] KO 5:16 ку 3vémp dx AX - Ve ox ax Ave Δve X dve ve Х and using Eq. 5:5 5:17 BV2AVZW Kỳ + 3 D We will also write 3 K? V fi (ky) : Dmin D 2./3K (Kỳ +3) Consequently, Dmin 2/3KÝ f2(ky) : ਨੂੰ (3 - ку f(KV) of(KV) OKV 2 D Kỳ + 3 6 (Kỳ 13) ) To obtain fz(Kv), we have from the defini- tion of f(Ky) that where Dmin is the minimum value of D (at standard W). afiky return ABVa 3k -3 ску 27 Kv Editor's insert: These relations are ob- tained as follows: 5:6 and ку f(KV) large -) 27 AB? ** + 3 Note that X/D can also be written as XV/(XV+ 7 P) and as fi(Kv) (X/Dmin). The latter forms may sometimes be the more convenient. Also, in the expression AP/P, P may be the observed or standard power whichever is the more convenient. hence Kv. Of(kw) f(ку оку 3 Kº-3 Kỳ + 3 5:4 FUEL FLOW IN LEVEL FLIGHT and 6 falkvi 2 3K-3 3- KV + 3 kỳ +3 It is convenient here to distinguish between "maximum level speed" fuel flows, such as those at the maximum steady level speed attainable without exceeding the engine limi- tations for continuous operation, and fuel flows under conditions not requiring operation at maximum continuous power. Fuel flows under the latter set of conditions must be precisely determined in order that the opti- mum conditions for cruising flight may be defined at any altitude other than the cruising ceiling, and a suitable reduction method is developed under (a) below. Then we have, after substituting Ave for dve * AT ΔΝ + AT Ts Ave : AP AV 용 ​- AN - AND Ve + AX AW fi(ky) - -f2(kv) Dmin WS 5:18 where Op Pave Ve ox + X ove +2f2(KV) - 3 Generally, it is not necessary to determine "maximum level speed” fuel flows with the same degree of precision as those at part power; moreover, the reduction process at full throttle is complicated by ram effects. The problem is considered in general terms in (b) below but no correction is derived because the main part of the correction is likely to be purely an engine matter. The remaining part of the correction involves the airframe, but as the correction is likely to be small even in extreme cases, it is concluded that in most applications it can be legitimately neglected. A method of correction is, however, indicated based on certain assumptions about the engine performance. Av = (-1){B+(1+0) We coup a } + $ * x :) (-7)+1+a) -5 AT = (- )la +B12 – 12) AN(-)13a+B) Ap : (a) : +B) : 5:19 (a) Fuel Flows at Less Than Maximum Power * By definition AP TOP POT N OP AN +7 P'ON' N P- 4T+ ox. At ox. AN It is proposed to correct the fuel flows for the direct effect of the departure of the test temperature and weight from standard at constant equivalent air speed and pressure ax AX : . ΔΤ ON X ΔΝ . 5:7 altitude. becomes From Eq. 5:18, at constant Ve and p we obtain AP Ap P AT -AT TS AN NS 312 ㄹ ​AN AX + Dmin fiſkv) AN Дх AP AT Ар + AT р TS 쁨 ​씀 ​AN filky) NS Dmin Δw + wsz/KV) = 0. - AW .f2(ky) WS = 0 5:21 Consider now how changes in P, X and N are associated with changes in fuel flow. We may write: 5:20 & P: (ve, p, WF, T) x = 2(Verp, WF, T) where the terms may be regarded as either all “corrections'' or all "errors''. If we now regard AP and AX as the "corrections to the engine shaft power and thrust required to give the observed equiva- lent air speed at standard temperature, and AN the associated correction (if any) to the engine speed, and also regard AT and AW as "errors" in the test conditions, Eq. 5:20 N: 03 (Ve, P, WF, T) where $3 is determined by the engine control system and will in certain cases be a con- stant. Then at constant Ve and p AP WE OP AWE TOP AT P T TS . POWF WE ax AWF AX : WF OWF WE WF ax AT T: OT TS Any change of N with temperature at constant Ve, P and WF is small and has been neglec- > ted. AWF AN WE ON WF δN N N OWF . WF 5:22 9 where AP, AX and AN are "corrections" and AT is an "error" as in Eq. 5:21; AWF is the correction to be applied to the observed fuel flows. Substituting in Eq. 5:21 we obtain AWE WE OP WE ON AN: NOWF WF. ox WF p OWF + fi (ky): Dmin OWE {ap. { AT } AW fs (KV) WS To op Ta ox AT +AP + fi(KV). TS Рота Dmin ata OTO AWE AW i.e., RF f2 (ky) WE TS WS AT RT 씀 ​5:23 5:8 where WE WE OP (1 + a). ax (-)" + RE: WE ON -(3a+B). NOWF - POWF + fi (ky): Dmin OWE TOP T O x RT: (-[(a+p12-3)(1+21 0) 1 + a) + f(ky): Р OT Dmin OT 5:24 The associated correction to engine speed (if any) follows from Eq. 5:22, It will often be possible to determine WF/P times do/a WF, WF/N times oN/OWF, WF times ox/ oWF experimentally. (b) Fuel Flow at Maximum Speed The reader will have noted that the method of reduction of fuel flows at powers below maximum continuous avoids the influence of ram on power, thrust and fuel flow by making corrections at constant equivalent airspeed, the fuel flow correction being deduced from the change in engine shaft power and thrust necessary to maintain the test equivalent airspeed under conditions of standard weight and temperature, (2) A correction for the indirect effect on fuel flow of the change in equivalent airspeed associated mainly with the power and thrust alterations caused by (1). This of course involves the airframe but inves- tigation suggests that the correction is un- likely (even in extreme cases) to exceed 2%. This latter correction will be needed for under optimum range conditions; i.e., flying at the cruising ceiling appropriate to maximum continuous power. However, for "maximum level speed" fuel flows at other heights it would seem legitimate to ignore this part of the correction for the errors introduced by its omission will usually be within the limits of engine to engine charac- teristic variations. It may even be possible, in view of the reduced accuracy requirements for rated fuel consumptions, to concede that the effects of realistic changes in ram and air tempera- ture on the power-fuel flow relationship are negligible. If so, the specific range figures can be obtained by correcting the level speeds to standard conditions by means of Eq. 5:18 and adjusting the fuel flows in proportion to the AP derived for the reduc- rion process. When dealing with “maximum level speed" fuel flows, where the engine is operating under conditions subject to an arbitrary limitation (which may itself change with air temperature), it is no longer possible to correct at constant equivalent airspeed be- cause the required engine shaft power and jet thrust are now fixed. The correction to the fuel flow, therefore, consists of two parts: (1) A correction at constant equiva- lent airspeed for the direct effect on fuel flow of the departure of the test temperature from standard with the engines operating at a constant limitation. This is purely an engine matter. 5:5 EQUATIONS FOR CLIMB REDUCTION It is proposed to correct the climb per- formance for changes in air temperature, shaft power and thrust at constant air pres- Bure and equivalent airspeed, Weight correc- tions to rate of climb are not normally required; if they are, the methods of Ref. 1 or 2 may be used. 5:9 (a) Reduction Process From Eq. 5:1, we may deduce that at constant Ve . ηΡ n Pro + xve dP + P | dTV Xve dx + 2 T no vo +XVe X w dlho) DVE + Who din o) D Ve thivo W . 5:25 If we now write XVe : rmpro +xve) and hd • DV/W= rate of climb equivalent to the drag power, we have 1 dT (1-7). in dP + η Р T) dx tr х dh hath 1 | DT h 2 T hd the 2 T 5:26 - where we note that 1-r = TP Tomp Vo + X Ve and that at constant pressure altitude Eq. 5:11 gives dolo : -dT/T. Substituting for dn/n from Eq. 5:12 we have rdx dI h «-1{ "1+a) + $? (a+B12 - ) - ( 3a+8} + ) + dh hid thi - х 2 т. nath 5:27 and the approximate equation for small finite difference is where Ah is the correction to be applied to the observed rate of climb and AN, AP, AT and AX are "errors" in the test condi- tions. Δx Δή rigth AN BN NS ΔΡ Bp Р - AT Вт Ts х (b) Pressure Height and Pressure Rate of Climb 5:28 where BN : (1-1) (3a+B) Вр (1-r) (1+0) BT = (1-1) (a+B/2 - + ) h 1 2 A complication results from the fact that heights and rates of climbs are commonly deduced from pressure measurements, the th 5:29 5:10 following: 2w\V2 pressure gage being calibrated in height above sea level sea level in the ICAN or NACA standard atmospheres. Thus, when the local temperature at a particular pressure height differs from standard, the pressure rate of climb will differ from the true rate of climb, their ratio being equal to the ratio of the corresponding air densities. Vemp - Company "13C0q7d'esi-vo Po for the evaluation of Ky and 12 SCDe Dmin : 2W The most straight forward way of dealing with this difficulty is to correct the observed pressure rate of climb if desired. ( пь°e (c) Longitudinal Acceleration for Eqs. 5:18, 5:19 and 5:24. In the foregoing, longitudinal accelera- tions during the climb have been ignored. This is legitimate for performance reduction work on turbo-propeller aircraft. If desired, however, a more rigorous treatment may be deduced by substituting for h and Ah in Eqs. 5:1, 5:10 and 5:28, etc. he and Ane These can usually be estimated with sufficient accuracy from estimates of CDe and the assumption of an average value for e (0.77 to 0.80), but if suitable flight data are available, these data should be used. Engine he = ŕ it h + 19 dh . + Gento + 5 Until further experience is gained, the engine derivatives listed below should be measured at each pressure altitude for which standard performance is required. ht V 9 5:30 (a) With respect to forward speed, In this case it is most desirable to transform the observed pressure rate of climb into a true rate of climb as a prelim- inary to the reduction process. Inclusion of the acceleration term is considered un- necessary for turbo-propeller aircraft, but it is routine for turbojet aircraft which climb at high true airspeeds with the result that the term V/8 times dv/dh is appreciable. Ve /p times oP / O Ve , Ve /X times oX/a Ve are required at constant throttle setting hp and T for use in Eq. 5:19. These quantities are most conveniently measured on single- engined aircraft by conducting accelerated and decelerated level flight runs. On multi- engined aircraft they can also be determined more reliably (if less conveniently) by holding a constant throttle setting on one engine while varying the settings on the remaining engine. 5:6 DATA REQUIRED IN ADDITION TO THE BASIC PERFORMANCE QUANTITIES (b) With respect to air temperature, Airframe The only airframe prerequisites are the T/P times OP/OT, T/X times ox/aT are required 5:11 (d) Fuel flow RPM linkage (1) at constant hp, Ve and We for evaluating A WF/WF in Éq. 5:23 (powers below rated') (2) at constant hp, Ve and throttle setting for evaluating AP/P, X/X in Eq8. 5:18 and 5:28 for "rated" powers WE/N times on/ oWe represents the way in which engine speed and fuel flow are linked; it is thus peculiar to the type of engine and will be established during other measurements without particular investiga- tion. It is required in the fuel consumption reduction Eq. 5:24 and in conjunction with WF/P times oP / oWF, WF/X times oX/OWE in the establishment of N/P times DP/ON and N/X times OXON for AP/P and AX/X N // of Eqs. 5:18 and 5:28. (3) at constant hp, Ve and tailpipe temperature for powers below rated, for evaluating AP/P, AX/X in Eq. 5:18 if it is required to correct the level speeds of the fuel consumption runs to standard condi- tions 80 that their consistency with the "rated" level speeds may be checked (see (e) below). (e) General C The most critical measurement is (2), the determination of the rated powers over a representative range of air temperatures, During these measurements the effects of operative limitations should be investigated. This will reveal whether any change in oper- ative limitation such as is described in section 5:2 is present and the air temperature at which such changes in regime occur. Any such changes are likely to have a marked effect on the temperature derivatives (see section 5:8). It may often be either impracticable or impossible to obtain precise control over those parameters which are intended to be held constant during the determination of a particular derivative. Should this occur, corrections can usually be made quite safely for small errors in Ve , from foreknowledge of Ve /P times oP / Ove , particularly if the latter is small. Similar adjustments may be made for small errors in setting up other parameters. (1) and (3) can be obtained by covering a range of jet tailpipe temperatures and fuel flows at constant Ve and hp at each air temperature and cross plotting as required. Limited experience has shown that the functioning of the automatic engine control system may not be entirely consistent. Fur- ther, the measurement of jet tailpipe temper- ature may not be very accurate. These, and the normal causes of scatter in perform- ance tests, make it desirable to compare all the level speed data on a common basis, to check the consistency of performance at rated and below rated conditions, (c) With respect to fuel flow, WF/P times DP / oWF, WF/X times o X/OWE are required at constant hp, Ve and T. These will generally be a by-product from the runs required at each temperature level to estab- lish the temperature derivatives of (b), and thus no special provision is required. They are needed at powers below "rated" for application in the fuel consumption reduction Eq. 5:24, In a constant speed engine, for example, where the operative limitation is jet tailpipe temperature, a convenient basis of compari- son is to plot aircraft speed against tailpipe temperature at standard air temperature and aircraft weight. To achieve this, the differ- ential method of section 5:3 may be used for the correction of the level speeds of the fuel consumption runs of section 5:4 as well as for the rated conditions for which it is specifically intended. 5:12 In correcting "non-rated" level speeds it is convenient to assume that the jet pipe temperature (T4) remains constant and an appropriate power derivative at constant T4 is used. This can be deduced from plots of observed power against T4 at two or more fixed air temperatures (T) if such data be available. sary until sufficient experience is gained to enable generalization to be made with relia- bility. It is desirable therefore that experi- mental results on derivative measurements on turbo-propeller engines of all types should be forwarded with details of the engine control system for digestion by a central agency, so that simplification may be expedited. 5:7 REQUIRED ACCURACY OF THE COEFFICIENTS If level speed data are available at air temperatures more or less randomly dis- posed about two or more distinct tempera- tures, it is advisable to correct these to the nearest standard air temperatures, and to present the results in a carpet plot of Ve against T and 14. Such a carpet is valuable for showing whether the limited information at the rated conditions fits into the general picture. This problem was considered in detail in Ref. 3. The broad conclusions that were reached therein were that the accuracies of estimation of the various parameters of the equations should be as shown in the accompanying table in order to meet the accuracies of correction shown in parenthe- sis. Derivative measurements will be neces- Level Speeds Rate of Climb Performance Item (Ave Fuel Flow AWF WF to 0.01 ( to 0.01 ) Ve (An to 0.01 hor 1/2 ft. per sec.) Reduction Parameter Ay +0.01 RT/RF +0.1 Вр +0.05 a Cp on ПОСР +0.1 +0.1 +0.05 B j.on . +0.1 +0.1 +0.1 η δυ ӘР ove +0.2 محمد داد cel OX +0.2 X ove 5:13 These limits are fairly wide and should enable single values or a small number of values to be used for the coefficients Ap, AN, Bp, BN, etc. if the present evidence that power is propor- tional to fuel flow, and independent of air temperature at a given fuel flow and pressure height, is substantiated by tests on other engines. This asumption is probably already valid in the case of “maximum level speed” fuel flows. 5:8 POSSIBLE SIMPLIFICATIONS OF THE METHOD If these features are repeated on other engines, there would appear to be a reason- able possibility of generalizing on engine performance and simplifying the method considerably. Since its development, the method has been used intensively on one aircraft only. In case the engine control system was such that the engine speed was held constant, and it was found that fixed values of the propeller terms could be taken on the basis of the accuracies quoted in section 5:7 and, of course, terms in ON/aWF in Eq. 5:24 and in OP/ON in establishing AP could be ignored. 5:9 REDUCTION OF LEVEL SPEEDS AND FUEL CONSUMPTIONS BY PERFORM- ANCE ANALYSIS (a) Introduction Also, although not actually used, single values of the power and thrust derivatives with respect to aircraft speed could have been adopted without objectionable errors, and perhaps a single value of X/D. Thus Av (Eqs. 5:18 and 5:19) was dominated by Ky, and could be satisfactorily defined by Ky independently of altitude and engine set- ting. It is possible that the need may arise for adopting the alternative method for level speed and range reduction. An outline of a possible approach is presented here, the approach being similar to that of methods developed elsewhere for the reduction of level speeds for reciprocating engine air- planes. It is emphasized that the method has not yet been applied to a turbo-propeller aircraft and thus unforeseen problems may be encountered. The derivatives of rated engine power with respect to air temperature fell into two classes: (1) where the jet tailpipe tempera- ture varied with air temperature (T/P times OP/aT effectively zero ) (2) where the jet tailpipe tempera- ture was held constant by adjusting fuel flow. In this case, T/P times DP/T had an appreciable value, but could be assumed constant (2-3) independent of engine setting and altitude without leading to objectionable errors in the particular case examined, provided the temperature range through which reduction was made was not excessive ( 5 10°C). (b) The Approach Considered It will be assumed that Reynolds number and Mach number effects on airframe drag and propeller efficiency are negligible and that, at a given pressure altitude, ambient temperature effects on specific fuel consump- tion can likewise be ignored. The first requirement is a universal power required curve independent of weight and al- titude. With the usual notation, in stabilized level flight W2 Dož pové scde+ PV-be 2 1 In fuel flow reduction the range reduction equations may later be materially simplified 5:31 5:14 Then drag power DVelo and flow. Thus, if the shaft power-fuel flow relationship is established for each pressure altitude under test conditions, such a relation- ship can be applied to standard conditions and range performance deduced.* Unfortu- nately, the jet thrust fuel flow relationship is more sensitive, and allowances may have to be made for its variation with air tempera- ture, if its effect on performance is not sufficiently small. W2 1 drag power to = 'po SCD Vo+ Ź - + 2 Ve & Boote v 3 drog power:vo Ve Roscoe SCDe W 3/2 (c) Standard Power 1 w + 121b²e Ve or To implement the equations previously developed, it is necessary to know the power available under standard conditions, at the test equivalent air speed and pressure alti- tude. Thus 3 mp Jo + xve.gov tuz ve XV ki +k2 w Ve W 3/2 w novo + x ve W 3/2 S 5:32 needs to be evaluated for the required engine settings. Whence all flight results can be represen- ted as a single functional plot against Ve W. This plot should provide a check a on the assumptions made, including the values Now if AP, AX are the required correc- tions to PT, XT of n. OP ON ОР Ps: PT + AP=PT+ AT + ОТ ON XS = XT + Ax = XT+ OX AT + = ax OT ox AN ON 5:33 During the development of this power required curve, measurements of installed powers, thrusts and fuel consumptions, will be obtained for a range of forward speeds at selected pressure altitude, at tempera- tures not very different from standard, and speeds differing from those obtainable under standard conditions by virtue of the differ- ences in test air temperature and aircraft weight from standard. From these test data, standard installed power available curves will be developed, in conjunction with test measurements or estimates of OP/T, DP / Ove, etc., and with due allowance for propeller efficiency, this leads to standard performance. at the test Ve. If the magnitudes of oP/a Ve , ax/a Ve be known, then (mp Vo + XVels w 1.5 S the power available parameter, may be Evidence to date suggests that for these engines, at a given pressure altitude shaft power is proportional to fuel flow, and in- dependent of air temperature at a given fuel * It has already been remarked that such a state of affairs will also bring about a simpli- fication in the reduction of fuel consumption data by the differential method. 5:15 where T - test air temperature plotted over the necessary small range of Ve Ws, for each rating and the rated performance arrived at by comparing these power available curves at each setting with the power required. TR = air temperature at which change of regime occurs Ts . standard air temperature T4 jet pipe temperature. . In applying these temperature corrections to power, it must be appreciated that oP/OT, etc., may change significantly with air tem- perature, as a result of an operative limita- tion being encountered. Thus part of the correction may be, for instance, at constant jet tailpipe temperature where oP / OT will be appreciable, and part where the tailpipe temperature is dependent upon T and OP/OT is small. It is necessary, therefore, to know the air temperature at which such changes of regime occur. Just as much prior information about engine behavior is therefore needed to utilize this method as is necessary for the differ- ential method, and thus the auxiliary program on derivative measurements outlined in section 5:6 still has to be undertaken. 5:10 CONCLUDING REMARKS Thus AP may be equal to (Tp-to)(87) Tg = 117 + f[T] TRT (TS-TR) (OM) The methods given should provide a foun- dation on which reduction routines can be built as and when the need arises. The flexibility necessary at the present stage makes them rather more tedious to use than the methods used for piston engined aircraft, but there are grounds for hoping that experi- ence will permit simplifications. OP + AN ON (T4=const.) REFERENCES 1. Herrington, Russell M., Shoemacher, Paul E., “Flight Test Engineering Manual,” Edwards, California, USAF Technical Report No. 6273, 1953. 2. Cameron, D., "British Performance Reduction Methods for Modern Aircraft," U, K., Ministry of Supply Report No. AaEE/Res/170, 1942. 3. Lush, Kenneth J., “Differential Performance Reduction Methods for Turbopropeller Air- craft," U, K., Ministry of Supply Report No. AAEE/Res/275, 1952. € 5:16 TABLE I fi(Kv) = = 2 V3 Kv2/(Kv 4 + 3) for values of Ky .02 .07.08.09 Ky .00 .01 .03.04 .05 .06 1.3 1.00 8 1.00 1.00 1.00 1.00 1.00 1.00 .99 .99 .99 1.4 .99 .99 .99 .99 .98 .98 .98 .98 .97 .97 1.5 .97 .96 .96 .96 .95 95 .94 .94 .94 .93 1.6 .93 .93 .92 .92 .91 .91 .90 .90 .89 .89 1.7 .88 .88 .87 .87 .86 .86 .85 .85 .84 .84 1.8 .83 .83 .82 .82 .81 .81 .80 .80 .79 .79 1.9 .78 .78 .77 .77 .76 .76 .75 .74 .74 .73 2.0 .73 .72 .72 .71 .71 .70 .70 .69 .69 .68 2.1 .68 .68 .67 .67 .66 .66 .65 .65 .64 .64 2.2 .63 .63 .62 .62 .62 .61 .61 .61 .60 .60 2.3 .59 .59 .58 .58 .58 .57 .57 .56 .56 .56 2.4 .55 .55 .55 .54 .54 .53 .53 .53 .52 .52 2.5 .52 .51 .51 .50 .50 .50 .49 .49 .49 ,48 2.6 .48 .48 .47 .47 .47 .46 .46 .46 .46 .45 2.7 .45 .45 .44 .44 .44 .44 .43 .43 43 .42 2.8 .42 .42 .42 .41 .41 .41 .41 .40 .40 .40 2.9 .40 .39 .39 .39 .39 .38 .38 .38 .38 .37 Ку .0 .1 .2 .3 .4 .5 .6 .7 .8 .9 3 .37 .35 .33 .31 .29 .28 .26 ,25 .24 .22 5:17 TABLE II f2(Ky) = 6/(Ky4 + 3) for values of Kv Ky .00 0 1 .01 .02 .03 .04 .05 .06 .07 .08 .09 1.3 1.02 | 1.01 .99 .98 96 .95 .93 .92 .91 .89 1.4 .88 .86 .85 .84 .82 .81 .80 .78 .77 .76 1.5 .74 .73 .72 .71 .70 .68 .67 .66 .65 .64 1.6 .63 .62 .61 .60 .59 .58 .57 .56 .55 .54 1.7 .53 .52 .51 .50 .49 .49 48 .47 .46 .45 1.8 .44 .44 .43 .42 .42 .41 .40 .39 .39 .38 1.9 .37 .37 .36 .36 .35 .34 .34 .33 .33 .32 2.0 .32 .31 .31 .30 .30 29 29 28 .28 .27 2.1 .27 .26 26 .26 25 .25 .24 .24 .24 .23 2.2 .23 .22 .22 22 .21 21 .21 20 .20 .20 Ky ,0 .1 2 .3 .4 .5 .6 .7 .8 .9 2 .32 .27 .23 .19 .17 .14 .12 .11 .09 .08 3 .07 .06 .06 .05 .04 .04 .04 .03 .03 .03 4 .02 5:18 APPENDIX USE OF LOGARITHMIC DIFFERENTIALS IN PERFORMANCE REDUCTION δY a2 X1 - AQ,X, X2 Αα дXI al = Q, Y 3 In performance reduction equations it is often simpler and neater to use what are sometimes called "logarithmic differ- entials." (a) General Form: A:6 If y: f (X1, X2, ... ) Y A:1 XI OY YOX1 : al logarithmic differentials take the general form A:7 dY XI DY XI X2 OY axz Y OXI XI YOX2 X2 + +... and similarly, for X2, X3, and so on. Hence A:2 DY dXI dX2 t.. ta2 al X1 This relation is fairly easily derived, since by definition X2 A:8 OY dY : OY ax dX + dX2 ... In particular, since axedxe+ A:3 1/2 -1/2 n Pro:nP 2 ()) TSL and if we divide both sides by Y and re- arrange the terms a little, the logarithmic form is obtained. A:9 (b) Products of Powers: where PSL and TSL are constants, we may write The logarithmic form is neat when the function is a product of powers of the independent variables. For example, if dimpto) d, din dP + 7 Р + | OP. 2 P. | dTo 2 Ta ηΡσ @ Y : AXI a 2 X2 A:10 A:4 where A is a constant, we have This form is not only neat, but it enables one to work with the proportional changes of n and P without needing to know their absolute values, which is often very conven- ient. Proportional changes in power and also drag, can often be generalized, enabling the same relations to be used for all engines or airframes of a class irrespective of size or weight. a 2 δY ox 2 Aa, x,ºr-1 AQ X2 A:5 5:19 (c) Sums of Functions: In particular, if 이 ​Y = x;"' + 02 +X2 +:: A:14 The logarithmic form is less conven- ient for sums, but may sometimes be desir- able, when dealing with a mixture of sums and products or powers. Suppose that then dy Y t, Y= f,(X1) + f2(x2)+ .. 22+ Xl.of. dx/ X1 +. xiol + x fi oxi A:11 A:15 then x, a, ofl.dXI Qi' dX + X1 62 x, al + x2 + dY : axl A:16 A:12 For example, the drag equation can often be written and we can write W2 dY Y X1 of dxi XI DE AVē+B Y OXI V2 A:13 A:17 60 W2 B dD dve 2 ve dve + dW 2 W 2 AV? W2 AVS-B ve D Ve W2 AVS+B Ve - В A:18 With the aid of Eqs. A:8 and A:16 as "standard forms”, many performance equations can be differentiated “logarithmically” by inspection. 5:20 AGARD FLIGHT TEST MANUAL VOLUME I, CHAPTER DATA REDUCTION AND PERFORMANCE TEST METHODS FOR RECIPROCATING ENGINE AIRCRAFT By Daniel O. Dommasch Associate Professor of Aeronautical Engineering Princeton University 3 VOLUME I, CHAPTER 6 CHAPTER CONTENTS Page TERMINOLOGY 6:1 INTRODUCTORY COMMENTS 6:1 6:2 LEVEL FLIGHT - POWER REQUIRED 6:1 6:3 POWER AVAILABLE IN LEVEL FLIGHT 6:5 6:4 DETERMINATION OF Vmax 6:16 6:5 RANGE AND ENDURANCE TESTING (CRUISING FLIGHT) 6:17 6:6 PILOT TECHNIQUE IN LEVEL FLIGHT PERFORMANCE TESTING 6:25 6:7 THE INSTRUMENTATION REQUIRED FOR LEVEL FLIGHT TESTING 6:26 6:8 CLIMB AND DESCENT TESTING - SAWTOOTH CLIMBS 6:26 6:9 CLIMBS TO CEILING 6:29 6:10 PILOT TECHNIQUE FOR CLIMB TESTS 6:35 6:11 MEASUREMENT AND CORRECTION OF TAKE-OFF CHARACTERISTICS 6:37 6:12 MEASUREMENT AND CORRECTION OF LANDING CHARACTERISTICS 6:45 6:13 CONCLUDING REMARKS 6:46 REFERENCES 6:47 TERMINOLOGY Units CD Drag Coefficient Numeric Сре Effective Parasite Drag Coefficient Numeric CL Lift Coefficient Numeric e Span Efficiency Factor Numeric AR Aspect Ratio Numeric S Wing Area Square Feet P Air Density Slugs/Cu.Ft. Density Ratio P/ Po Numeric H, h Altitude Feet Hp Pressure Altitude Hpc Pressure Altitude Corrected for Position Error Feet P, Pa Ambient Pressure In. Hg. abs. PCO Carburetor Deck Static Pressure In. Hg. abs. Pmo, MAP Absolute Intake Manifold Pressure 1 In, Hg, abs, 9 Dynamic Pressure tev? Lbs. per Sq. Ft. W Weight Pounds WF Weight Fuel Flow Rate Lbs./Hr. t Time Seconds V Velocity Feet/Second Vo Observed Indicated Airspeed Knots or mph Vc , CAS Calibrated Airspeed (Vo corrected for instrument, position and lag errors) Knots or mph Ve Equivalent Airspeed Knots or mph 1.A.S. Indicated Airspeed Knots, mph TERMINOLOGY (Continued) Units Voc Best Climb Airspeed Knots, mph Vmax Maximum Level Flight Airspeed Knots, mph Vstall Minimum Level Flight Airspeed Knots, mph Vg Take-off Speed (true) Knots, mph VW Wind Velocity (true) Knots, mph Vtas True Airspeed Knots Vo Touch Down Speed (true) Knots Q Torque Foot-pounds n, RPM Revolutions Per Minute RPMT Turbine RPM 7p Propeller Efficiency Numeric Tco, CAT Carburetor Air Temperature OR Tas Temperature at Standard Altitude OR Tao, OAT Observed Outside Ambient Air Temperature R THP Thrust Horsepower Horsepower BHP Brake Horsepower Horsepower BHPao Observed Brake Horsepower Available Horsepower BHPro Observed Brake Horsepower Required Horsepower BMEP Brake Mean Effective Pressure Pressure units EBP Exhaust Back Pressure In. Hg. bsfc Brake Specific Fuel Consumption Lbs./(BHP) (Hr.) MSD Metering Suction Differential In. Hg. R Range Miles (Feet) TERMINOLOGY (Continued) Units E Endurance Hours (Seconds) dR/W Specific Range Miles/Pound dE/W Specific Endurance Hours/Pound RC Rate of Climb (usually fpm) RCT True Rate of Climb (usually fpm) RCE Equivalent Rate of Climb (usually fpm) 8 Take-off Distance Feet SW Take-off Distance Through Air Feet sg Ground Distance Feet so Distance Travelled by Air Feet Fm Mean Accelerating Force Lbs. Y Angle of Flight Path with Horizontal Degrees K1, K2, kl, k2 Constants fpm Feet Per Minute Subscripts e Equivalent S Standard O Sea Level (also used for observed quantity) req Required av Available ew "ew" System Value T Test Value (also turbine) obs Observed 6:1 INTRODUCTORY COMMENTS Our basic assumption is that CL co Сре + Te AR Consequently, the thrust horsepower re- quired to maintain level flight at a given airspeed may be written Because of the nature of the reciprocat- ing engine-propeller combination, airplanes equipped with this means of propulsion are, for the most part, designed to fly at speeds less than drag divergence. Although Mach number effects are present and do modify the lift slope at velocities less than that for drag divergence, there are no major modi- fications in the lift-drag relationship until drag divergence is encountered. Thus, for all the important phases of level flight per- formance testing of reciprocating engine aircraft, we have the following fundamental relation valid 9 Co. Sv W?V.as THP rea + 550 q?S?FEAR(550) 6:2 : where a 1/2p Vrepresents dynamic pres- sure and CL has been replaced by its equal W/qs. CL co Со: Сре Coe + PCDE S.V3 THPrea : TeAR W? + PV STEAR(275) 6:1 1100 where Ср 8 Airplane Drag Coefficient Сре Effective Parasite Drag Co- 'efficient Since we cannot conveniently measure thrust horsepower, it is desirable to express power required in terms of brake horsepower, and for this purpose we know that CL 8 : Lift Coefficient THPrea BH Prea тр e 8 Airplane Efficiency Factor where тр = propeller efficiency. AR : Aspect Ratio Thus pcdes.vo BHPreg 1100p In our further discussions, we shall assume that Eq. 6:1 is valid in all cases for air- planes with reciprocating engines, and that normally coe and e remain constant at all times during the data correction process. Where it is important, the effect of variations of CDe and e will be noted. W? + PVS TAR( 275)7p 6:3 6:2 LEVEL FLIGHT - POWER REQUIRED Now letting Coes 2 KI 1100 mp 6:4 The most convenient method of presentation of power required data for piston engine airplanes known to the author is called the "ew" system which presents data from all altitudes on a common sea level-standard weight plot. The basis for this method and a procedure for fairing the data and at the same time finding the generalized plot are given below. and TEARS(275)p K2 6:5 6:1 and multiplying through by V, we obtain Taking ratios of horsepowers and velocities, THPreglew 3 K2W (BH Preq)(V) = p KiVa + = THP rea VIN Ws W Р 6:6 Ws THPreq vo )。 6:7 and . Examining Eq. 6:6, we see that if we plotour power required data at a given weight and density altitude in the form (BHPreq) (V) vs. V“, we should obtain a straight line with slope pK and intercept K2W/p: However, we shall not always obtain this straight line because of minor variation in e and CDe and somewhat more important variations in np during tests. Vew : Vo vo WS W 6:8 Now let us return to Eq. 6:3. This may be written Ka W} : BHProq = 8,0 KAVO+ Cm () pov Ws Eqs. 6:3 and 6:6 can be generalized so that data taken at various altitudes fall on a single curve. To show how this comes about, consider flight at the same value of CL at sea level under standard conditions, at standard weight, and at some altitude where the density is p and the weight differs from standard. At sea level, the density is Po and the weight is Ws , so that letting V : sea level standard velocity, ew or, on multiplying by Molo po W Nola BHP reqvo ( m ) . ( ) .K, 9.(oyť vo ( ) Cn WS W Ws W 2ws Vew S K2 W} W + Povov Ws Copvos 1 2w3co THPrenew 1100 550 Posc However, from Eqs. 6:7 and 6:8 this is (con- verting from THP to BHP) BHPregew - Kip. Vow + K₂ W} Povew 6:9 At altitude (flying at the same CL) we must have the same value of co as at sea level according to the basic assumption of this analysis; therefore, at altitude and on multiplying by Vew . Kzw. New = Kip Vew + Ро 6:10 2W (BHP regew V: ✓ PCLS Since Po and We are constante, Copy's THPreg 1 550 2wc (BHP reqew Vew = kivew + kz 8 1100 PSCE 6:11 6:2 7 where k, poki P. Coes 11001p where stabilized flight had not been achieved during tests or where instrument malfunction produced erroneous readings. Test data plotted in accord with Eq. 6:10 will appear as in Fig. 6:1. Points such as "a" in Fig. 6:1 are immediately seen to represent un- satisfactory data. 6:12 K2W k2 WS TEARS(275)7 P. PO 6:13 Since altitude is not a factor in Eq. 6:10, test data of power required versus velocity from all altitudes converted to terms of BHPregew and Vew, should plot into a single curve which is approximately a straight line. This curve may be used to conveniently fair the test data and to eliminate points As we have previously noted, CDe and e may not actually be constant throughout the range of lift coefficients and Mach numbers encountered in the tests, and, therefore, the "straight line” of Fig. 6:1 may actually exhibit curvature, particularly at its ends. At the high speed end (low CL), curvature may be produced by Mach number effects, whereas at the low speed end (high CL), Сое (BH Preq.! Vew slope Pocoes - Central Portion :K, T100 10 k Extrapolated From Central Portion Minimum Flight Speed vaw 2 Intercept . WS TEARS-2757pp. =k2 Fig. 6:1 General Plotting Form of Equation 6:10 6:3 curvature is the result of tire relation CZ Co-CDe+ : TAR MO BHP not fitting the actual CD - CL curve at the high lift coefficients. This low speed end curvature is quite common, and is duc es- sentially to the fact that CDe itself becomes a function of CL (or a) at lift coefficient values of about one. Curvature is also in- troduced by variations in propeller efficiency which were disregarded in the analysis. Not- withstanding the fact that our "straight line" may be a curve, the use of the procedure described is highly recommended because it provides such an excellent means of fair- ing data, and moreover provides a means of estimating CDe and e, provided the central portion of the curve is used as a basis for such computation. Eqs. 6:12 and 6:13 pro- vide the required relations between e, and the slope and intercept (obtained by extrapolation) as illustrated in Fig. 6:1. Vow , Fig. 6:2 Gencralized Power Required Curve If Ve is expressed in units of feet per second, Eqs. 6:12 and 6:13 provide the fol- lowing 1.53 WS From Fig. 6:2, the actual power required curve at any altitude may readily be ob- tained, for from Eqs. 6:7 and 6:8 we have Vew W пkzь* пр T V = 6:14 Ws and and 6:16 BHP ew : coe Pos 1100np ki BHP ✓o W. 6:15 where b wing span in feet пр propeller efficiency (85% on the average). From examination of Fig. 6:2 the reader will note that no data are shown below the speed for minimum power. This does not represent an omission, but rather reflects the fact that it is almost impossible to ob- tain stabilized level flight test data at speeds below the minimum power speed in what is known as the reverse command region. Ac- tually it is desirable to obtain at least one data point in this region to provide a clear definition of the speed for minimum power; however, due to the difficulty of achieving stabilized level flight, data are generally Once the test data have been reduced and faired in accord with Fig. 6:1, Fig. 6:1, a generalized horsepower required plot may be prepared using the faired curve of Fig. 6:1. This curve appears as in Fig. 6:2. 6:4 taken while the airplane is either slightly climbing or descending, and a correction is then imposed to account for rate of change of altitude (during the test run it is necessary, of course, that the speed, at least, be stabi- lized). testing and reduction of data have many ad- vantages for they bring about savings in both time and money. Our next topic is the measurement of power output and power available. 6:3 POWER AVAILABLE IN LEVEL FLIGHT Because power is rate of change of energy, we know that the descending airplane requires less power from the engine than when in level flight, and the converse holds true for the climbing airplane. The correction to be applied is from fundamental considerations On military aircraft, brake power devel- oped is generally measured using a torque- meter, which measures the torque developed at the propeller shaft, and a tachometer which permits determination of shaft RPM. The brake power is given by the relation W dH A BHPrea ( ) 33,000 dt np 6:17 2nQ BHP : 33,000 where dh/dt is positive for a climb and is assumed (in Eq. 6:17) to be expressed in feet per minute. 6:19 where Q : torque in foot pounds : RPM. n The increment of Eq. 6:17 is subtractive from the observed power in a climb and is additive in a descent. Thus (BHP reqlevel - (BHPreglobserved W Although it would be desirable to measure thrust and then determine thrust horsepower from the relation THP : TV/550, (where T is in pounds and V in feet per second), there is unfortunately no simple way to accomplish this for the well-known eason that the thrust in the propeller shaft is not the actual overall thrust produced by rotation of the propeller in the presence of the airframe. + () dH dt 33,000 6:18 In obtaining data in the reverse command region, rate of climb or sink should be re- stricted to about 50 fpm, in which case the rate of change of pressure altitude may be assumed equal to the actual rate of change of true altitude without introducing serious error. On smaller engines no provision may have been made to permit the installation of a torquemeter. In these cases it is necessary to refer to engine characteristic charts for the estimation of power output delivered by the engine. Engine charts are usually pre- pared in two parts, one being a sea level test calibration in terms of BHP under stand- ard sea level conditions as a function of MAP for various values of RPM, and the other, a computed set of curves for altitude conditions (although these altitude curves may also be obtained from tests). The instrumentation required, flight tech- niques, and tabular data reduction forms for reducing level flight data will be discussed later in the chapter. However, before we . become involved with these we should first consider the topics of power available deter- mination, and range and endurance testing. As we shall show, the actual flight testing for all of these may be done simultaneously, as may the data reduction. Such simultaneous The use of power charts is not entirely satisfactory inasmuch as in their prepara- tion it is generally presumed that such factors 6:5 as oil pressure, oil temperature, cylinder head temperature, etc., either remain con- stant or vary in some predetermined manner. Similarly it is normally assumed that the fuel-air ratio is properly maintained by the carburetor and that the ignition is perfect. Since the assumed conditions are seldom satisfied in actual test work, the power charts obviously only provide an approximation to the truth and should be used to determine the power delivered only when torquemeters are not available, it is a good idea to define precisely what we mean by the term. By definition, critical altitude is the altitude at which the throttle must be fully opened to develop rated power at rated RPM. An unsupercharged engine (or a ground boosted engine) has a critical altitude of zero feet (i.e., at sea level), whereas a supercharged engine, which is capable of exceeding engine limits at sea level, will have a critical altitude consider- ably above sea level. Aircraft with multi- speed, multi-stage superchargers will have several different critical altitudes depending on the possible blower combinations. The use of power charts to determine ac- tual power output is discussed in several text books and need not be considered in further detail here. However, these charts have another use, namely, the determination of standard power available and of critical altitudes and blower shift points for aircraft equipped with gear-driven superchargers. These topics we must consider in some detail. Because we do not directly measure power output of an engine, critical altitude is fre- quently defined either in terms of limiting BMEP or MAP at rated RPM, and tests are conducted using either BMEP or MAP as the determining factor. If a BMEP gage (torquemeter calibrated in terms of BMEP) is available, the procedure for determining critical altitude with a single-speed, single- stage supercharger is as follows: Constant BMEP Tests for Critical Altitude No corrections to standard are required for points determining the power required curves, since here we are merely using an engine as a means of measuring the drag requirements of the airframe. Thus, assum- ing a torquemeter is available, the meter reading together with the tachometer reading directly gives us the required information in determining power required data. Above the highest critical altitude, full throttle will be used to determine the test Vmax points at the various test altitudes; these points, therefore, may also be considered test points for the power available vs. altitude curves for the airplane in level flight. We note that due to differences in ram at the air intake, and possibly due to differences in engine cooling characteristics, a different power available curve will probably exist for the climbing airplane than for the craft in level flight. Consequently separate tests are nor- mally required for level flight and climb power available characteristics. The deter- mination of critical altitudes, shift points and standard power available as functions of altitude is discussed below. Depending on whether military or normal rated power data are desired, the propeller controls are set to maintain the proper limiting RPM. Starting out at low altitude, level flight runs are made holding BMEP at its limiting value by throttle control. Runs are made at part throttle until an altitude is reached where full throttle is required to give rated BMEP at rated RPM. Above this altitude a series of full throttle runs is made, and developed BMEP and MAP are recorded along with carburetor deck static pressure, carburetor air temperature, pressure alti- tude, and outside (ambient) air temperature. If plotted, the test data will appear as in Fig. 6:3. To determine the standard critical alti- tude, no corrections need be made to the part throttle curve; however, the full throttle curve must be corrected by means of the procedure outlined below or some simi- lar method. Assuming that the tests are Since we must test for critical altitude(s), 6:6 Full Throttle Curve ( Constant RPM) dH Observed Critical Altitude lo PRESSURE ALTITUDE , Part Throttle Curve ( Constant RPM ) BMEP (Torquemeter Reading) Fig. 6:3 General Plotting of Pressure Altitude vs. BMEP 6:7 To: OAT, °R conducted on a day warmer than the standard day, the observed and corrected data will appear as in Fig. 6:4. In. H20 RAM: (Pco-Pal, In. Hg. or 13.6 BHPMeasured torquemeter horse- power Computations Corrected Full Throttle Curve (1) Determine observed supercharger pressure ratio Test Dota For Worm Day Pmo H. ( )。 Pm PC Standord Critical Altitude (pen + RAM) . (2) Determine carburetor air tempera- ture corrected to standard conditions LTITUDE RE ALT PRESSUR . Tes = Too Tao + Tas (3) Determine the corrected pressure ratio for standard conditions, (Pm/Pc's, from Fig. 6:5 by entering the curve at observed pressure ratio and carburetor air tempera- ture, Tco, and following the slope of the nearest slanting line to the carburetor air temperature for a standard day, Tcs. BRAKE MEAN EFFECTIVE PRESSURE Fig. 6:4 (4) Determine corrected manifold pres- sure by the relation Pms + RAM) (PM/Pc)s : Pco (Pm/Pc's (POO (5) Determine corrected BHP A most convenient procedure to follow to correct full throttle power available data to standard is the semi-empirical Wright Case II Method. The procedure for the case of a single-stage supercharger is as follows: BHP BHPO (Pms/Pmo) (Tco /Tcs JE Data Required When making these single-stage correc- tions it is usually more convenient and less time-consuming to use the special charts as given in Figs. 6:6 and 6:7. PCO • Carburetor deck static pressure, In.Hg.abs. - Pmo • Manifold pressure, In. Hg. abs. Тco • Carburetor air temperature, °R If an engine is equipped with a single- stage two-speed supercharger (gear shift type), there will be two critical altitudes. The procedure for determining the lower altitude in this case is the same as described above with the engine operating in low blower. To determine the second (higher) critical altitude, the procedure is as follows: Ро • Ambient pressure, In. Hg. Tas : Standard altitude temperature corresponding to Po (standard atmosphere tables), OR Starting at an altitude a little higher than 6:8 MAIN AND AUXILIARY STAGE PRESSURE RATIO CORRECTION CHART EXAMPLE : a.) PM/Pc = 1.840 at 90° F. CAT (c) = 1.865 of 60°F. CAT ° b.) PM/Pc m/Pc or Pcr po 3.2 3.0 2.8 2.6 2.4 2.2 SUPERCHARGER COMPRESSION RATIO, 2.0 D C! B 1.8 1.6 1.4 1.2 -60 -40 -20 O 20 40 60 80 A 100 120 DEGREES F. AIR INLET TEMPERATURE, To or To Fig. 6:5 Main and Auxiliary Stage Pressure Ratio Correction Chart 6:9 BRAKE HORSEPOWER CORRECTION TO STANDARD CONDITIONS SHEET 1 OF 2 MANIFOLD ABSOLUTE PRESSURE INS. HG. 20 24 28 32 36 40 44 48 WRIGHT CASE DEVIATION OF OAT FROM STD-'c 25 20_15_10_5 5 10 15 20 25 COLDER THAN WARMER THAN STD. STD SUPERCHARGER EFFICIENCY CORRECTION 1.03 1.01 1.00 0.99 0.98 0.97 0.96 32 1.04 102 3.1 30 -3.0 8 01. 12 -2.9 2.9 14 91 18 2.8 2.8 20 2.7 2.7 2.6 2.6 -2.5 2.5 OBSERVED SUPERCHARGER 2,4 14 CARBURETOR DECK PRESSURE – INS. HG. COMPRESSION RATIO C Fig. 6:6 Brake Horsepower Correction to Standard Conditions OBSERVED SUPERCHARGER COMPRESSION RATIO 2 2.3. 2.2 2.2- _01 12 2.1 97 2.0 2.0 20 22 24 26 28 30 30.000 10.000 poo 10,000 Log 52 56 1.08 1.071 1.04 - 1.03 DUE TO DENSITY * 1.06 21.03 1.04 1.03 1.02 21.01 1.06 FINAL BHP 'CORRECTION 10.99 -0.98 70.97 0.96 / 0.95 0.94 20 24 28 32 36 40 44 48 MANIFOLD ABSOLUTE PRESSURE INS. HG. EXAMPLE OBSERVED DATA ALTITUDE 15,000 FT. OAT +1 MAP 40.5 IN. HG. CARD. DECK PRESS. 17.0 IN. HG. CALCULATED DATA OAT - STD. TEMP. 14--148). 16.3°C (AT) DERIVED DATA FROM CHART CORRECTED MAP 41.4 IN. MG. POINT (F) BHP CORRECTION FACTOR (MULTIPLYING) POINT (J) 1.088 bec 20 15 10 5 COL DER THAN STD 10 15 20 1.00 WARMER THAN 0.99 STD. SL-8,000 0.98 BHP CORRECTA 0.93/ . 000 or 0.97 0.94 10,000 30,000 18,000 6:10 0.91 SHEET 2 OF 2 BRAKE HORSEPOWER CORRECTION TO STANDARD CONDITIONS WRIGHT CASE I MANIFOLD ABSOLUTE PRESSURE – INS. HG. 20 24 28 32 36 40 DEVIATION OF OAT FROM STD.- 'c COLDER THAN WARMER THAN STD. STD. 25 20 15 10 5 0 5 10 15 20 25 48 44 SUPERCHARGER EFFICIENCY CORRECTION 1.04 103 102 101 100 0.99 0.98 0.97 0.96 52 1.9. 9 1.8 97 1.7 18 20 24 SUPERCHARGER COMPRESSION RATIO 26 OBSERVED SUPERCHARGER 28 , CARBURETOR DECK PRESSURE - INS. HG. 1.6 COMPRESSION RATIO X30 1.5 32. Fig. 6:7 Brake Horsepower Correction to Standard Conditions 1.3 1.2 OBSERVED 1.2 30,000 20,000 18,000 10,000 36 52 20 24 28 32 40 44 48 MANIFOLD ABSOLUTE PRESSURE INS. HG. 1.04 1.09 1.08 1.07! - 1.06 / 1.05 1.04 / 103 1.02 /LOI 1.00 FINAL BHP CORRECTION 10.91 0.98 0.97 -996 / 0.95 -0.94 JL - 6,000 DUE TO DENSITY 11.02 koo WILO 29 18 10 3 10 15 20 COLDER THAN WARMER THAN STD. 0.99 STD. 1 0.98 1 0.97 1 0.96 20,000 1,000 BHP CORRECTION 000'or | 0.93 0.92 10.000 - 10.000 6:11 Wright Case II Method outlined previously may be extended to provide the following cor- rection. The procedure is as follows: Obtain the low blower critical, the shift to high blower is made and the throttle retarded to hold the engine to its limiting BMEP at full RPM. We note here that the limiting BMEP in high blower will usually be less than the low blower value, and that the engine speci- fications should be consulted to determine the proper value. Tests are now conducted in the same manner as for determination of the low blower critical altitude, the final corrected plot of test data appearing as in Fig. 6:8. Ро Тоо Tas : Ambient pressure, In. Hg. Ambient air temperature, 'R Standard air temperature, ºr Measured auxiliary supercharger inlet static pressure, In. Hg. Measured auxiliary supercharger inlet temperature, 'R po Too : PCO Measured carburetor deck static pressure, In. Hg. TCO Measured carburetor air tempera- High Blower Critical Altitude ture, 'R pmo Snitt Point Measured manifold pressure, In. Hg. (P. - Po), In. Hg. Measured torquemeter power RAM * PRESSURURE ALTITUDE, Lou Blover Critical Altitude BHPO Procedure (1) Determine actual auxiliary super- charger pressure ratio BRAKE MEAN EFFECTIVE PRESSURE Pc DCO Рco ) Fig. 6:8 po (pao + + RAM) poo From these data the best blower shift point may also be determined as shown in the figure. (2) Determine auxiliary supercharger inlet temperature under standard conditions as follows: T'ago:tc+ Tos - Too The case of the two-stage two-speed blower is handled in the same manner as described above, and these data appear as shown in Fig. 6:8 excepting that there are now three critical altitudes and two shift points. (3) Determine corrected auxiliary su- percharger blower ratio for standard con- ditions (PC/Pa)s, from curve Fig. 6:5 by entering the curve at observed pressure ratio and auxiliary supercharger entrance temper- ature To., and following nearest slant line to the corrected auxiliary supercharger en- trance temperature Tos. Two Stage Systems To correct to standard power available when using a two-stage supercharger, the 6:12 (4) Determine corrected carburetor deck pressure from the relation is similar to the turbo-driven unit which has no definite critical altitude, and has its per- formance limited by structural consider- ations. PCS = (po + RAM) Menos 3 - par le monde (5) Determine observed main super- charger pressure ratio (Pm/Pc). (6) Determine carburetor air temper- ature corrected to standard conditions, Tos : Tco Tao + Tas (7) Determine corrected main super- charger pressure ratio by entering Fig. 6:5 with observed main stage pressure ratio (Pm/Pc), and observed carburetor air tem- perature Tco. Follow slope of nearest slant line to corrected carburetor air temperature Tcs, and read corrected main supercharger compression ratio (Pm/Pc's. More recently the compound engine has made its appearance. As presently used, this engine is equipped with a single-stage two-speed supercharger together with blow- down exhaust turbines whose output is fed back to the crankshaft through fluid cou- plings. Because the blow-down turbines are not directly used for supercharging, but have the purpose of converting exhaust gas energy to shaft horsepower, the operation of the compound engine is quite similar to that of a conventional reciprocating engine. Indeed, it has been found that the same correction procedure used to determine power available characteristics of the standard engine (as outlined above) also applies to the compound units. (8) Determine final corrected manifold pressure by multiplying corrected main su- percharger pressure ratio by carburetor deck pressure as follows: Pm Pms Dc (9) Determine corrected BHP from re- lation Pms BHP= BHP. : els DCS In the United States, military engines of 2000 cu. in. displacement and above are required to have provisions for the instal- lation of torquemeters (AN 9500a, para.D-8) which also function as BMEP gages. For test work, such gages are always installed. However, this is not true operationally in all cases, and frequently the operational pilot is provided only with a tachometer and a manifold pressure gage for use in deter- mining engine power output. For this rea- son, data on power available are frequently presented in terms of MAP rather than BMEP. To convert the BMEP data to MAP for the part throttle region, we have the constant RPM relation 8 (一​) (==) ( The special charts contained in Figs. 6:6 and 6:7 can be used with modification for the two-stage system. However, this is not usually done because the method becomes rather complex, BMEP ~ MAPCOAT) 6:20* In addition to the types of superchargers described above, we also have the two-stage automatic-shift variable-speed type, which has no definite critical altitudes, and which automatically regulates the engine to prevent development of excessive BMEP values. With this latter type of installation, it is extremely difficult to obtain repeatable data because of the unknown variation of super- charger RPM. This type of supercharger * This relation immediately follows from the well-known equation BHPov = (constant) ( (MAP). (TjV2 since at constant RPM, BMEP is directly proportional to BHPov. . . 6:13 Constant MAP Tests For Critical Altitude For the further condition of constant BMEP, Eq. 6:20 immediately provides the relation OAT test MAP std - MAP test OAT std 6:21 Because of decreasing exhaust back pres- sure at altitude, less MAP is required to maintain a given BMEP at altitude than at sea level. Therefore, a condition of constant BMEP with varying altitude corresponds to a decreasing MAP with altitude. This is il- lustrated in Fig. 6:9. As noted, the observed test values of MAP for part throttle operation are corrected according to Eq. 6:21; full throttle MAP test values are corrected ac- cording to the Wright Case II Method de- scribed in this chapter with the corrected MAP data appearing as in Fig. 6:9. Constant MAP tests are conducted in ex- actly the same way as constant BMEP tests except that the MAP is kept constant rather than BMEP. Runs are continued until full throttle operation will not give rated MAP, and the altitude(s) where rated MAP corresponds to full throttle is (are) recorded. If plotted, the raw test data appear as shown in Fig.6:10. To reduce the test data to standard, no corrections are imposed on the part throttle region. However, the full throttle curves are corrected by the Wright Case II Procedure. The corrections will shift the location of the critical altitude(s), but the general nature of the curves will still be as shown in Fig. 6:10. The corrected plot of MAP vs. Ho deter- mines the critical altitudes, but does not show the location of the shift points, nor does it directly yield information on the variation of power available with altitude. To obtain these data, in the absence of a torquemeter, it is necessary to refer to engine charts. It should be recognized that the use of these charts is not a very satisfactory a procedure, but rather is a necessary al- ternative in the event of unavailability of a When BMEP gages are not available for test work (as is the case when dealing with the smaller engines), critical altitude deter- mination is based entirely on MAP and RPM readings. The procedure is described below. Full Throttle Lines of Constant BMEP High Blower Critical Altitude TH Full Throttle Part Throttle PRESSURE ALTITUDE, Mp PRESSURE ALTITUDE ---- Low Blower Critical Altitudo ABSOLUTE MANIFOLD PRESSURE ABSOLUTE MANIFOLD PRESSURE Fig. 6:9 Variation of MAP with Altitude for Constant BMEP Part Throttle Operation Fig. 6:10 6:14 torque nose installation. Using the charts, the MAP, altitude and RPM data are con- verted to a plot of BHP av vs. altitude, ap- pearing as in Fig. 6:11. The increase in power available with the altitude shown in the figure is due to the fact that constant MAP operation causes an increase in BHP available with altitude. To obtain power available as a function of altitude, runs are made (which may be made as part of the power required test) at various altitudes at rated RPM and either at limiting MAP or turbine RPM. At low altitudes MAP limits are maintained by retarding throttle and partially bypassing the turbine. At the higher altitudes the throttle will be fully open, and allowable turbine RPM restrictions will prevent obtaining allowable MAP limits. Full Throttle Part Throtile For wide open throttle, we may suppose that the MAP obtained at a given test pres- sure altitude and fixed value of engine RPM is the same as under standard conditions. Then as long as we have not exceeded turbine RPM limits (even though the observed tur- bine RPM is not the same as the standard for the test altitude) we shall have the re- quirement that corrections must be imposed for: Critical Altitudes Shift Altitudo PRESSURE ALTITUDE , He Full Throttle (1) Non standard outside air tempera- ture Port Throttle (2) Non standard carburetor air tem- perature BRAKE HORSEPOWER (3) Non standard turbine RPM Fig. 6:11 (4) Non standard exhaust back pressure (5) Non standard BHP (power obtained from torquemeter) Engines With Turbosuperchargers To correct for the above deviations, we proceed as follows: As previously mentioned, turbosuper- charged engines display no well-defined criti- cal altitude; however, the determination of their power available as a function of altitude is still an important matter. From the engine characteristic charts (ob- tained by test or computation) we may obtain a value of the partial derivative When flying with a turbosupercharged in- stallation the pilot must observe three basic limits on engine operation. These are limits on: a(CAT) O (RPM)T (1) Engine RPM (2) MAP or BMEP at the test value of turbine RPM where CAT = carburetor air temperature and (RPM)T • turbine RPM. Knowing A(RPM)T, we may compute A(CAT)t, effective change in CAT due to non standard turbo RPM, from (3) Turbine RPM 6:15 2 in. Hg. decrease in EBP. Further the relation AICATII (CAT) a(RPM) . A(RPM)T. Trest There also exists another increment in CAT due to non-standard OAT. This is BHPS - BHP at constant MAP Tstd A(CAT)2 = (OAT)std -(0AT)test. . holds for turbosupercharged engines so that finally The total change in CAT is САТ, BHP std - BHPtest LE test ) CAT std A(CAT) = A(CAT)2 + A(CAT)T - - (OAT)std +.005 [EBPtest - EBP Pstal (OAT)test + (CAT) A(RPM)T. O (RPM)T Now CAT Std . The effects of changing exhaust back pressure are approximately given by the re- lation that the power increases 1% for each CATtest + (OAT)std O(CAT) - (OAT)test + O(RPM), A(RPM)T. Therefore CAT test BHP std - BHP test (CAT) CATtest + (OAT)std - (OAT)test + J(RPM) ) .(RPM)T OT +.005 (E8Ptest – EBP sta]} where n ~.5 6:22 6:4 DETERMINATION OF Vmax Eq. 6:22 may be used as a basis for cor- recting full throttle turbosupercharged en- gine data to standard. At less than full throttle at constant MAP the power may be presumed to vary in accord with Eq. 6:21. At a selected number of altitudes which should include sea level and the critical and shift altitudes the generalized power re- quired curve BHPew vs. Vew is used to ob- tain actual curves of standard BHP required. Knowing the variation of standard power available with altitude as determined in sec- tion 6:3, the intersections of the power avail- able and power required curves may be at Having determined the standard BHP avail- able, our next step is to obtain the variation of standard Vmax with altitude. This is dis- cussed in section 6:4 below. 6:16 for specific range and endurance given below: once determined. These intersections are the standard Vmax points for the altitudes considered. A typical plot of Vmax Vs. altitude for an airplane equipped with a two- speed single-stage supercharger is shown in Fig. 6:12. specific range dR η C 1 dw Со CO W 6:23 specific endurance 3/2 1/2 dE CL S P dw с CD 2 np. . 1 3/2 W 6:24 where R • Range in feet Critical Altitude W • Weight in lbs. с : bsfc Shift Point р • Air density PRESSURE ALTITUDE, Hp E : Endurance in seconds Critical Altitude ap : Prop efficiency CL • Lift coefficient CD : Drag coefficient Vmax S: Wing area Fig. 6:12 Recognizing that np and c are variables, the quantities to maximize are, respectively, 3/2, 6:5 RANGE AND ENDURANCE TESTING (CRUISING FLIGHT) 03 ) and ( 7 et ). CO Range and endurance test data are fre- quently obtainable along with power required and available data. Therefore, the theory of range and endurance testing will be covered at this point followed by a brief discussion of general data reduction form and required instrumentation. Because the optimum values of "p, c and CL/CD or C32/ co may not occur simul- taneously, it is necessary during tests to consider variations of each of these quanti- ties. Ideally, if engine bsfc and prop efficiency were constants, power required data could be used to obtain directly specific range and endurance information. In any actual case, however, variations of these quantities pro- hibit this simple an approach. To see what is required, consider the Breguet equations At any given weight and altitude, the values of the aerodynamic parameters CL/CD and CL"/2/ Co are dictated by the flight speed. The variations in speed alone suffice to tie these variables down. Engine bsfc is a func- tion of BMEP and RPM which together de- fine power output. Normally at any given values of RPM, bsfc will tend to be least at maximum allowable BMEP, although this is 6:17 not an inflexible rule. Propeller efficiency is a function of the quantity V/nD (advance ratio) and therefore, changes with both for - ward speed and RPM. This is true even though the propeller is of the constant-speed type because such a propeller adjusts its blade angle to produce enough torque to govern RPM rather than to achieve maximum prop efficiency. Therefore, maximum prop efficiency is not necessarily attained at the same RPM required for limiting engine BMEP. Similarly, optimum engine and pro- peller operation may be unobtainable at the aircraft aerodynamic speeds for best range and endurance. Changes in weight and alti - tude change the aerodynamic speeds for best range and endurance, which in turn alter prop efficiency and power required, and therefore, the effects of weight and altitude must be considered along with the effects of speed changes, RPM, and BMEP setting. To attempt to account fully for all the effects of all variables in testing for range and endurance characteristics is generally not practical from the standpoint of time and effort involved. Moreover, the actual varia- tions encountered in carburetor functioning are great enough to make any attempts to attain absolute precision ridiculous. Thus, the procedures prescribed herein are di- rected toward achieving the best results con- sistent with reasonable expenditure of time and effort, rather than toward the achieve- ment of meticulous accuracy. Normally, we are required to obtain fuel consumption data for several different alti- tudes and weights rather than for all possible combinations of these. Because power re- quired and available data must usually be obtained at the same altitudes and weights as those chosen for fuel consumption testing, data for these latter items are most easily obtained simultaneously with fuel consump- tion data. the given RPM (as specified by the engine manufacturer). Speed is reduced by reduc- ing MAP (and the BMEP) until minimum speed is obtained. Normally, 6 to 10 points will suffice to define the curve between the minimum and maximum speed points. The procedure is repeated for different RPM values. Usually, from the point of view of convenience in making the test, it is desir- able to start first with maximum rated RPM and BMEP and then work down with subse- quent runs at lower RPM's. The lowest RPM used is determined by generator cutout speed or similar considerations. Data taken during tests are readings of: MSD Carburetor metering suction differential (a measure of fuel flow) RPM Engine rotational speed MAP Absolute intake manifold pressure BMEP Or torquemeter reading ОАТ Outside air temperature Observed airspeed Fuel flow VO Fuel quantity Pressure altitude Carburetor deck pressure Carburetor air temperature, cylinder head temperature, mixture setting, blower setting, throttle position (part or full), external configuration, When two-stage superchargers with inter- coolers are used, additional data as listed below should be obtained to permit proper reduction of power available data simultan- eously with the range and endurance data: Pressure at inlet of auxiliary super- charger stage Temperature at inlet of auxiliary su- percharger stage The procedure is as follows: At the specified altitude(s), pressure, and weight(s), a series of level flight runs is made at a constant RPM setting, starting out with the maximum allowable BMEP for Pressure at outlet of auxiliary super- charger stage 6:18 Temperature at outlet of auxiliary su- percharger stage Reduction of Range and Endurance and Power Required Data Lines of Constant RPM Because existing fuel flowmeters and car- buretors are rather erratic in their behavior, the first step in reducing the test data is correlation of MSD with observed fuel flow. For this purpose, data on these two quan- tities obtained at all weights and altitudes are plotted as in Fig. 6:13 and an average curve relating MSD to fuel flow is deter- mined. METERING SUCTION DIFFERENTIAL Separato Plot Required for Each Pressure Altitude BRAKE HORSEPOWER Fig. 6:14 METERING SUCTION DIFFERENTIAL o Doc Loon Limit Ayorage Rich Limit WF WEIGHT FUEL FLOW RATE WF" • WEIGHT FUEL FLOW RATE Fig. 6:13 Lings of Constont RPM Separato Plot Required for Eoch Pressuro Altitudo Following the preparation of the above plot, all the data (at varying weights and RPM's) at one pressure altitude are plotted as shown in Fig. 6:14. BRAKE HORSEPOWER Fig. 6:15 Using the data of Fig. 6:14 and the faired curve of Fig. 6:12, prepare the cross plot of Fig. 6:15 which represents the faired fuel flow vs. BHP data for various values of RPM at the selected altitude. The foregoing procedure is repeated for all test altitudes. 6:19 TABLE 6:1 - FLIGHT DATA ANALYSIS SHEET Std. Gross Wt. Model Configuration * Observed Data Corrected for Instrument Error 車 ​. (1) Flight No. (2) Vo (3) CAS (4) Ve (5) * Press. Alt. (6) Pos. Error (7) lipc (8) TOAT (9) Po P ( (10) *MAP Pmo |(11) *RPM (12) *Torque Q (13) BHIPO |(14) *Fuel Flow 157hr (15) GPH |(16) bsfc (17) * MSD (18) Gross Wt. - WO (19) WoW |(20) (W6Wsia W.7W ' (21) (W/W) 3/2 (22) Vew for wt. |(23) BHP s1 (24) (BHP)w for wt. |(25) Vew |(26) (BHP)si Vew (27) BMEP (28) *Pco In. Hg. (29) *Polo In.Hg. (30) (Pc/Polo (31) Too (32) Too (33) Tag (34) To's (35) (Pc/Pa')s (36) Pcs (37) (Pm/Pcle (38) * TCO (39) TCS (40) (Pm/Pc)s (41) Pms (42) BHPs (43) *Head Temp. °C 6:20 > Following the fairing of fuel flow data, the power required data are reduced to standard using the procedures developed in section 6:2. A typical data analysis sheet for this purpose, which also includes the steps required for reduction of power available data, is given as Table 6:1. Use of Flight Data Analysis Sheet 1. Number of flight 2. Observed data 3. (2) and airspeed calibration 4. (3) corrected for pressure altitude 5. Observed data 6. (2), (2) - (3), (5), (8) and correction chart for position error 7. (5) 1 (6) 8. Observed data, corrected for airspeed 9. (7) and (8) and atmosphere charts 10. Observed data 11. Observed data 12. Observed data 13. (11) and (12) and torque constant 14. Observed data 15. (14) + 6(lb/gal.) 16. (14) + (13) 17. Observed data 18. Computed from T. O, weight and fuel burned 19. (18), - standard weight 20. (19) 21. (20) 22. (4) + (20) 23. (13) + (9) 24. (23) : (21) 25. (22)* 26. (22) times (24) 27. (12) times conversion constant 28. Observed data, carburetor deck pressure 29. Observed data, auxiliary inlet pressure 30. (28) · (29) observed auxiliary compression ratio 31. Observed data, auxiliary inlet temperature 32. (8), outside air temperature 33. Standard OAT for (7) 34. [(31) - (32)] + (33) standard auxiliary inlet temperature + 35. (30), (31), and (34) and compression ratio curve 36. (29) times (35) standard carburetor deck pressure 37. (10) + (28) observed main stage compression ratio 38. Observed data, CAT 39. (38) - (31) + (34), standard CAT 40. (37), (38), and (39) and compression ratio curve of engine standard main stage com- pression ratio 41. (36) times (40) standard MAP (38)460 ៤ 42. (13) times (41) = (10) times standard BHP (39) 460] 43. Observed data, cylinder head temperature 6:21 Using the data from Figs. 6:15 and 6:16, plots of fuel flow vs. Ve or CAS are now prepared for each standard weight and al- titude as shown in Fig. 6:17. If the standard test weights specified differ from one another appreciably, as they may in the case of patrol or bomber type aircraft, it is good practice to account for unknown variations of propeller efficiency by prepar- ing more than one "standard” weight plot so that two or more generalized "ew"power required curves are obtained as shown in Fig. 6:16. From Fig. 6:17 the best endurance speed for any given set of circumstances is selected as the speed for minimum fuel flow. The engine RPM setting for this speed is also given. If we now take each curve of Fig. 6:17 and compute the ratio of true flight speed to fuel flow, we obtain data relating specific range to Ve or CAS (the specific range will have units of miles per pound, with V in MPH and fuel flow in lbs./hr.). These data may be plotted up in the form of Fig. 6:18 to deter- mine optimum range information. High Gross Weight Tests BHP Low Gross Weight Tests From Fig. 6:18 the speed and RPM setting for max range may be determined, as well as the engine setting for optimum range, at any selected airspeed. Provided small var- iations from standard weight (of the order of 10% of G.W.) are involved, it is possible to use the data of Figs. 6:17 and 6:18 to determine range and endurance character- istics at other weights, vow Fig. 6:16 The assumption we must make here is that for small weight changes, the quantity 7/c in Eqs. 6:23 and 6:24 does not vary appreci- ably, in which case we have (Specific rangel, Wz 6:25 Specific rangela wi 3/2 Web (Specific endurance), (Specific endurancela ) 6:26 WI Having obtained the curves of Fig. 6:16 (which we note could just as well have been obtained entirely separately from the fuel consumption data), we are in a position to determine corrected curves of fuel flow as functions either of true or equivalent air- speed. We note here that no attempt is made to correct engine characteristics themselves to standard, and that the corrections merely involve correction of the power required data. The procedure is justified on the basis that from a practical standpoint the engine corrections are small in the first place, and in the second, normal variations in carburetor performance produce larger unknown errors than could be corrected for. with the equivalent airspeed remaining con- stant during correction, Thus, to correct the data of Figs. 6:17 and 6:18 for small changes in standard weight each fuel flow or specific range point has its ordinate changed according to Eqs. 6:25 and 6:26 while the abscissa values remain unchanged. 6:22 RATE ( Note: Curves Required For Each Standard Weight And Altitude ) WEIGHT FUEL FLOW ines Of Constant RP Envelope Curve Horizontal Tangent WF Best Endurance Speed EQUIVALENT AIRSPEED, V Fig. 6:17 6:23 Horizontal Tangent Envelope Curve SPECIFIC RANGE IN MILES/LB. FUEL Max. BMEP Points Note: Separate Plot Required For Each Weight And Altitude. Speed For Best Range EQUIVALENT AIRSPEED, Ve Fig. 6:18 6:24 6:6 PILOT TECHNIQUE IN LEVEL FLIGHT PERFORMANCE TESTING (c) Achieving stabilized conditions Now that we are aware of what is required in the way of test data and have followed through the procedure for reducing these data, we are in a position to discuss the tech- niques which should be used to obtain good raw test data. These techniques are as fol- lows: . Since it is easier to slow an airplane at a given altitude value than it is to accelerate it, all level flight performance tests should be started at the Vmax speed. To achieve this speed most conveniently, the airplane is first climbed to an altitude several hundred feet higher than the test altitude. At this altitude the external configuration is set, the position of each external item is recorded, and the altimeter set to 29.92 in. Hg. to make sure it is reading pressure altitude. The engine controls are set to give maximum power at the test altitude and a slow descent is made to the test altitude during which Vmax is attained for the level flight run. (a) Planning the flight Before the flight, make sure the pilot knows exactly what data are going to be obtained and under what conditions he is to obtain it. This will normally entail his knowing: (1) The test pressure altitude(s) (2) The test gross weight range(s) (3) The engine power setting(s) and mixture control setting(s) The airplane should then be flown at the test altitude a sufficient period of time to insure that stabilized flight has been achieved. This will take from three to ten minutes. During stabilization the altitude should be held to within plus or minus 20 feet and the airspeed to plus or minus one mile per hour. After at least three minutes have passed and the airplane is stabilized, data may be taken. (4) The external configuration(s) to be checked. The data items to be recorded should be noted, and a convenient flight data card pre- pared to allow rapid data recording. If a photopanel is being used along with the data card, its operation should be checked prior to flight. We note here that efficient planning will cut down the flight time required to obtain data, and in this connection it is frequently possible to plan flights so that at least two extremes of altitude and gross weight may be tested for in a single flight. Following the Vmax run the engine controls are adjusted for the next lower power setting. The airplane is again stabilized and data are taken. The procedure is continued until the required range of speeds has been inves- tigated. As data for each point are obtained, note should be made of instrument fluctua- tions, and the amplitude of the fluctuation and average reading recorded. The apparent cause of the fluctuation should be determined and recorded. Change in configuration due to creeping of the cowl flaps, bomb bay doors, etc., should also be recorded as well as the time allowed to obtain stabilized flight, (b) Climb to test altitude (d) Specialized technique for obtaining low speed data After take-off and during the climb to the altitude for the first run, engine instrumenta- tion and airplane operation should be checked for satisfactory behavior. This avoids a wasteful climb to altitude should the equip- ment show signs of malfunctioning. It is quite difficult to maintain stabilized level flight over the range of airspeeds ex- tending from the stall to slightly above the minimum power required speed (known as 6:25 Horizontal Tangent Envelope Curve SPECIFIC RANGE IN MILES/LB. FUEL Max. BMEP Points Note: Separate Plot Required For Each Weight And Altitude. Speed For Best Range EQUIVALENT AIRSPEED, Ve Fig. 6:18 6:24 6:6 PILOT TECHNIQUE IN LEVEL FLIGHT PERFORMANCE TESTING (c) Achieving stabilized conditions Now that we are aware of what is required in the way of test data and have followed through the procedure for reducing these data, we are in a position to discuss the tech- niques which should be used to obtain good raw test data. These techniques are as fol- lows: Since it is easier to slow an airplane at a given altitude value than it is to accelerate it, all level flight performance tests should be started at the Vmax speed. To achieve this speed most conveniently, the airplane is first climbed to an altitude several hundred feet higher than the test altitude. At this altitude the external configuration is set,'the position of each external item is recorded, and the altimeter set to 29.92 in. Hg. to make sure it is reading pressure altitude. The engine controls are set to give maximum power at the test altitude and a slow descent is made to the test altitude during which Vmax is attained for the level flight run. (a) Planning the flight Before the flight, make sure the pilot knows exactly what data are going to be obtained and under what conditions he is to obtain it. This will normally entail his knowing: (1) The test pressure altitude(s) (2) The test gross weight range(s) (3) The engine power setting(s) and mixture control setting(s) The airplane should then be flown at the test altitude a sufficient period of time to insure that stabilized flight has been achieved. This will take from three to ten minutes. During stabilization the altitude should be held to within plus or minus 20 feet and the airspeed to plus or minus one mile per hour. After at least three minutes have passed and the airplane is stabilized, data may be taken. (4) The external configuration(s) to be checked. The data items to be recorded should be noted, and a convenient flight data card pre- pared to allow rapid data recording. If a photopanel is being used along with the data card, its operation should be checked prior to flight. We note here that efficient planning will cut down the flight time required to obtain data, and in this connection it is frequently possible to plan flights so that at least two extremes of altitude and gross weight may be tested for in a single flight. Following the Vmax run the engine controls are adjusted for the next lower power setting. The airplane is again stabilized and data are taken. The procedure is continued until the required range of speeds has been inves- tigated. As data for each point are obtained, note should be made of instrument fluctua- tions, and the amplitude of the fluctuation and average reading recorded. The apparent cause of the fluctuation should be determined and recorded. Change in configuration due to creeping of the cowl flaps, bomb bay doors, etc., should also be recorded as well as the time allowed to obtain stabilized flight. (b) Climb to test altitude (d) Specialized technique for obtaining low speed data After take-off and during the climb to the altitude for the first run, engine instrumenta- tion and airplane operation should be checked for satisfactory behavior. This avoids a wasteful climb to altitude should the equip- ment show signs of malfunctioning. It is quite difficult to maintain stabilized level flight over the range of airspeeds ex- tending from the stall to slightly above the minimum power required speed (known as 6:25 the region of reverse command); however, data are still obtainable over at least part of this region, the procedure being as follows: excess speed, flying very smoothly at the selected altitude, and stabilizing all the controllable variables as rapidly as pos- sible. Any deviation from standard tech- niques should be noted in the final report of the particular set of tests. (1) Assume and hold a constant air- speed at the test altitude. (2) Adjust power for as near level flight as possible (holding the rate of climb or descent to less than 50 fpm). 6:7 THE INSTRUMENTATION REQUIRED 7 FOR LEVEL FLIGHT TESTING (3) Make a five-minute run recording altitude at the beginning and end of the run. The table on the opposite page lists the special test instrumentation required for the level flight performance tests. A photo recorder is desirable but it is not necessary if the scope of the project does not warrant such an installation, (4) During the data reduction proce- dures the observed brake horsepower re- quired should be corrected for the time rate of change of potential energy by the relation (Rate of climb or descent) W BHPrea BHP I 6:8 CLIMB AND DESCENT TESTING SAWTOOTH CLIMBS 33,000 np with minus holding for climbing flight and the plus for a descent. Here nap is the prop efficiency which may be assumed to be between 80% and 90%. We note that excessive precision is not required in making the above correction inasmuch as the magnitude of the correction will normally be small as long as the 50 fpm restriction on climb or descent rates is observed. Normally the speeds for best rate of climb at various altitudes for reciprocating engine aircraft are obtained by using the sawtooth method of testing. Acceleration run tech- niques may also be used, but are not as satisfactory for piston engine aircraft as for jets. Since these acceleration techniques are discussed elsewhere in this volume, they will not be considered further here. Because this method is not as accurate as the stabilized altitude and airspeed proce- dure, it should not be used except when data are not obtainable in any other manner. Checking of descent characteristics of piston engine aircraft is generally not ac- complished in detail except for the landing conditions where the rate of sink becomes an important factor insofar as the structural integrity of the airplane is concerned. (e) General comments The outlined procedures for making level flight speed runs should be followed exactly, unless special considerations of a particular project require a different approach. It is anticipated that a shortening of the duration of the stabilization runs will be required for airplanes with high fuel consumption and relatively low fuel supply. This in turn will require extreme care in letting down without In this section we shall consider the nature of sawtooth climbs, and in sections 6:9 and 6:10 the continued climb to ceiling and the pilot techniques involved. It is noted that the maximum energy climb concept is ap- plicable to reciprocating engine aircraft as well as jets. However, since this topic is discussed elsewhere in the manual, no rep- etition will be made at this point, The sawtooth procedure is the oldest flight test method used to determine the speed for 6:26 SPECIAL TEST INSTRUMENTATION REQUIRED FOR LEVEL FLIGHT PERFORMANCE TESTS NECESSARY FOR INSTRUMENT POWER REQUIRED POWER AVAILABLE FUEL FLOW Sensitive Tachometer 1 1 1 Torquemeter (if installation permits) Sensitive Manifold 1 1 1 1 1 1 1 Pressure 1 1 1 Sensitive Altimeter 1 1 1 1 1 1 2 2. 3 1 3 1 1 1 Sensitive Airspeed Indicator Sensitive Outside Air Temperature (OAT) Gage Sensitive Carburetor Deck Pressure (CDP) Gage Carburetor Air Temperature Gage (CAT) Totalizer Type Fuel Quantity Gage (desirable to measure gross weight variation accurately) Fuel Flowmeter Metering Suction Differential Gage Auxiliary Supercharger Pressure Gage (inlet) Auxiliary Supercharger Temperature Gage (inlet) Auxiliary Supercharger Pressure Gage (outlet) Auxiliary Supercharger Temperature Gage (outlet) 3* 3 1 3* 3 1 2** 3** 3*** 3*** 1. Cockpit and Photo Recorder 2. Cockpit Only 3. Photo Recorder Only For Fuel Consumption Only ** For Two-Stage Supercharger Only *** For Intercooler Efficiency Survey Only, of Two-Stage Installation 6:27 best climb rate at a given altitude. In this procedure, one first selects the altitude(s) at which data are desired and then an altitude band of 1000 to 4000 feet (surrounding the test altitude) through which a series of climbs at different stabilized airspeeds are made. For each of these runs the time to climb through the altitude band is measured and the aver- age rate of climb through the band computed. As an example, consider the test pressure altitude of 20,000 feet. For this altitude the test band could be chosen as 19,000 to 21,000 feet or possible 19,500 to 20,500 feet if the rate of climb were low. Further assume that we want to check the airspeed range from 150 knots IAS to 300 knots IAS. . Horizontal Tangent RATE OF CLIMB Speed For Maximum Rate Of Climb V Fig. 6:19 Raw Sawtooth Climb Data at One Altitude 6:28 Our procedure would be as follows: Starting at some altitude less than the lower limit of the test band, full throttle is applied and a climb initiated which is sta- bilized at some airspeed close to 150 knots by the time the lower band limit altitude is reached. This stabilized airspeed is main- tained through the test band and the time required to climb through the band measured. After this first climb, a descent is initiated and the procedure repeated, say at 175 knots, 200 knots, etc., until the full range of air- speeds has been investigated. The fact that succeeding ascents and descents produce a flight path resembling a sawtooth pattern, leads quite naturally to the name “sawtooth climbs.” Now, as we have noted, the sawtooth pro- cedure is used only to obtain the speeds for best rate of climb at various altitudes, not the climb rates themselves. Therefore, no corrections are imposed for non-standard power available or for non-standard gross weight. This procedure is quite satisfactory for piston engine equipped aircraft whose fuel load is a relatively small percentage of the gross weight. However, determina- . tion of the effects of weight change are easily accomplished when necessary. The procedure is to run the sawtooths at various weights and from the data thus obtained to plot a curve of speed for best rate of climb at any altitude vs, weight. From this curve the speed for maximum climb rate at any standard weight may be directly obtained. A rough and ready rule for estimating the speed for maximum rate of climb for a pro- pellered airplane is: 6:9 CLIMBS TO CEILING Vbc - Vstall + ķxVmax - Vstall ) : 6:27 The sawtooth data obtained at various pressure altitudes define a curve of CAS for best rate of climb vs. altitude similar to that shown in Fig. 6:20. PRESSURE ALTITUDE, Hp In conducting the tests, although the nom- inal values of the airspeed may be 150, 175 knots, etc., stabilized airspeed values of 154 knots, 180 knots, etc., are just as satis- factory as the nominal values as long as air- speed is held constant during the climb, and satisfactory data points are obtained. A typ- ical plot of raw sawtooth data at one altitude is shown in Fig. 6:19. In obtaining sawtooth data, care must be taken to insure that the airplane is not climb- ing with or against a wind having a vertical gradient; otherwise, the kinetic-potential en- ergy balance will be modified sufficiently to give false data. Although correction tech- niques for gradient winds are available, in general, insufficient information is available to allow their proper application. We note in passing that the sawtooth method is dif- ficult to apply to aircraft with very high rates of climb because of the small incre- ments of time during which the airplane re- mains in the test band and the difficulties associated with achieving a stabilized climb airspeed. Also, altimeter (static system) lag at high ascent rates introduces errors of frequently undetermined magnitude. RATE OF CLIMB CALIBRATED AIRSPEED FOR MAXIMUM Fig. 6:20 CAS for Maximum RC 6:29 On the basis of the schedule shown in Fig. 6:20 continued climbs to service ceiling (al- titude at which RC: 100 fpm) are made, during which the following data are contin- uously recorded either by photopanel or by using a kneepad data card: aircraft, the actual climb to service ceiling is made in stages, and the data are presented in such a manner as to simulate instantane- ous blower shift. Although such a presenta- tion is not entirely realistic, it is considered to provide the most convenient representation of climb performance for the type of air- plane involved. (1) Pressure altitude (2) Time (3) IAS (4) CAT (5) Torquemeter reading (6) Engine RPM In making a measured climb to ceiling, the aircraft is first trimmed at the indicated speed for best climb for the starting alti- tude of the climb. With the propeller con- trols set at the proper RPM for the climb, sufficient MAP is used to maintain level flight at the climbing speed. The throttle is then opened to provide maximum permis- sible power while the observed airspeed is maintained constant until stabilized climbing conditions are reached. The procurement of useful data commences at the altitude and time at which V, BHP, trim, and configura- tion are stabilized. The altitude(s) at which full throttle is (are) reached should be noted carefully. (7) Gross weight (8) Carburetor deck pressure (9) Manifold pressure (10) Configuration including blower po- sition From the observed data are calculated the following: (1) Pressure rate of climb at various altitudes It is normal practice to continue a climb in any one blower setting to an altitude as much as 2000 feet above the blower shift point. The climb is then discontinued and the airplane altitude reduced to as much as 2000 feet below the shift point during the time the shift is being accomplished. The climb in the next higher blower is begun below the shift point and continued up to the shift point for the next higher blower. In other words, the climb is made in as many sections as there are blower ratios on the airplane, and the various sections are inte- grated to give the final climb curve. (2) Brake HP developed by the engine (3) Density ratio (4) True rate of climb (5) Equivalent BHP (6) Equivalent rate of climb (7) Standard BHP available vs, altitude (8) Standard rate of climb vs. altitude Once the basic test data have been obtained as outlined above, the power available data for the climb configuration are reduced to standard by the procedures previously de- scribed in this chapter to give a curve as shown in Fig. 6:21 which defines the various shift points and critical altitudes for the climbing airplane, For gear-driven supercharger-equipped 6:30 available and BHPro is the observed brake power required to maintain flight speed only, then from basic considerations the true rate of climb is given by Critical Altitudo BHP. - BHP RCT: np Le)33,000 12) ALTITUDE Wo 6:29 Climb Shift Point Critical Altitudo STANDARD where пр is the propeller efficiency. We may define an equivalent rate of climb in the same fashion that we defined equivalent airspeed, i.e. RCE : : RCT vo 6:30 and from Eqs. 6:29 and 6:30 BRAKE HORSEPOWER AVAILABLE BHP. - BHPT 33,000. RCE = np Vo Fig. 6:21 Standard Power Available in Climb . Wo 6:31 Further, define RCew as After the power available data have been corrected, the observed rates of climb (ob- tained from a plot of pressure altitude vs. time, and which are actually the rates of change of pressure altitude) are corrected to tape line or true rates of climb by using Eq. 6:28, which is obtained from the relation Δp : -Pg Δh RCT:vo. Wsz Ws : RCe Wo Wo 6:32 then 33,000 mp ( Mole R Cew : Ws BHP Wo RC true - RCobs Rcobs (1905 ) 6) 2) Ouroove (i) 'obs Ws - BHProvo B 6:28 Wo 6:33 where obs = observed and s: standard. We may further define a term, equivalent brake horsepower available as From this point on data reduction con- veniently follows what is called the “'Equiv- alent Rate of Climb Analysis Procedure" the basis for which is developed below. Ws BHPaoew 2 ВнРаос Wo and recall that Development of the Equivalent Rate of Climb Concept W BHP Proew BHProvo (en If BHP,. is the observed brake power 6:31 Then, on multiplying both sides of Eq. 6:33 by Ws / Wo, and, therefore, Eq. 6:35 does serve as a basis for fairing test data from all altitudes. Moreover, the “ew" presentation allows correction to standard altitude conditions by the following procedure. RCew Ws = 33,000 mp BHP a Rew - BHProew) 6:34 Solving Eq. 6:34 and BHPoo ew From the test data corrected in accord with Eq. 6:35, prepare a plot of observed equivalent power available vs. equivalent rate of climb + BHP roew BHPoCew 33.0001p (RCew) Ws (Note: 6:35 BHP 00 EW , BHP. Ws BHP ovo Wo ( ; - -). Govo (w RCTVO ( ) Now, if under any set of circumstances it happened that the climb schedule provided for a climb at constant propeller efficiency and at constant CL then the quantities Ws /33000 mp and BHProew would be constants leaving only two variables in Eq. 6:35, namely, BHPCO ew and RCew. In other words, we would have a relation of the type y = mx + b where Wsla RCew : Wo similar to that shown in Fig. 6:22.) EQUIVALENT POWER AVAILABLE, BHP. HP ooow y : BHPaoev RCew Ws m = EQUIVALENT RATE OF CLIMB, RCOW 33,000 mp b = - BHProew Fig. 6:22 Equivalent BHP., vs. Equivalent RC In such a case as this, a plot of BHPao ew V8. RCew would be a straight line with data taken from all altitudes falling along this same line. The slope of the line would de- termine the propeller efficiency and the in- tercept the brake power required at sea level for non-climbing flight at standard weight. Actually, one must expect that neither np nor CL will remain constant and, therefore, a plot of Eq. 6:35 will exhibit curvature, The curve of Fig. 6:22 is somewhat anal- ogous to the brake horsepower required curve, in that it represents the basic climb performance of the airplane without regard to the actual power available for climbing. To use this curve, we proceed as follows. Experimental results, however, reveal that if the test data are plotted in the form pre- scribed by Eq. 6:35 only one curve results, At any selected standard altitude we refer 6:32 to our previously determined curve of stand- ard power available in the climb and obtain the standard power available at that altitude. This value of BHPas is then converted to equivalent power available by means of the equation Eqs. 6:36 and 6:37 assume that the equiv- alent rate of climb vs. equivalent power available curve has already been corrected to standard weight, and, therefore, addi- tional weight corrections are not required. BHPasew - BHPasvo 6:36 Knowing the value of BHPasew , we may enter our experimental curve of reduced climb data (Fig. 6:22) to obtain the value of RCew. Then, by virtue of Eq. 6:32, the true standard rate of climb is given by In some cases where the weight correc- tions are of small magnitude, they may be neglected throughout the entire data reduc- tion procedure; however, if the observed weight differs from standard by 5% or more, corrections should be imposed. Data reduction is most conveniently car- ried out in tabular forms of the type illus- trated below: RCT : RCew To 2 6:37 TABLE 6:2 - DETERMINATION OF BHPaew vs. RCew 1. Pressure Altitude 2. OAT °C 3. Tobs : OAT 273° 8 4. Ts, °K (Standard temperature at pressure altitude) 5. Tobs/Ts 6 . 7. P from (1) and (2) vo: (plodle BHP 40 (from torquemeter and tachometer) 8. 9 . RCobs dhp/dt 10. RCT: (9) x (5) 11. Wo from test data 12. Ws/Wo; (Ws = standard weight) 13. (Ws/W.)" (Ws /W.) 8/2 312 14. 15. RCew • (10) x (13) ( 16. BHP. 'aoew (8) X (14) 6:33 TABLE 6:3 - DETERMINATION OF STANDARD RC 1. Standard Altitude 12. BHPas (standard power available) 3. To (standard density ratio) 14. BHPasew = (2) x (3) 15. RCews from plot of Table 8:1 16. RC std = (5)/(3) . 17. 1/RCstd : 1/(6) Following the determination of the standard rate of climb, the standard time to climb may be determined using row (7) of Table 6:3 To do this, first plot 1/RC std vs. standard altitude. This gives a curve similar to that of Fig. 6:23, which represents the case of a single stage two-speed super- charger installation. Critical Altitude . (R.Custo) Critical Altitude Shift Point I. -ΔΗ STANDARD ALTITUDE Fig. 6:23 6:34 The specialized pilot techniques required to obtain raw sawtooth and time-to-climb data are discussed in section 6:10. The area under the curve of Fig. 6:23 taken to some altitude, say H, is the time to climb to that altitude, This area can be found by any one of the graphical integration methods, all of which will serve to evaluate the integral HI dh RC . 6:38 6:10 PILOT TECHNIQUE FOR CLIMB TESTS (a) General In considering climb tests of any type it is essential that the test day provide smooth air without high altitude temperature inver- sions and without gradient wind. Attempts to obtain data under unfavorable weather conditions are almost certainly doomed to failure and represent wasted effort and im- proper planning. The reduced climb data obtained as pre- viously explained is generally presented in a final composite plot similar to that shown in Fig. 6:24. STANDARD ALTITUDE - FEET Time To Climb Rate Of Climb STANDARD BHP AVAILABLE RATE OF CLIMB CAS FOR BEST & TIME TO CLIMB CLIMB RATE Fig. 6:24 Composite Presentation of Climb Data 6:35 (b) Sawtooth Climbs least three different airspeeds in the vicinity of the estimated speed for best climb, and preferably for four or six speeds. (c) Timed Climbs When the sawtooth climbs are to be made, the altitude increment or sawtooth depth to be checked is determined by the rate of climb characteristics of the airplane. For instance, if at a given altitude the rate of climb of one airplane were 700 FPM, a sawtooth depth of 1000 feet would be the maximum to use; however, if the rate of climb were 5000 FPM, an increment of at least 2000 feet would be required. The selection of the altitude band is a matter of initial planning. As noted previously, either sawtooths or acceleration runs are used only to deter- mine the schedule of speed vs. altitude for conducting timed climbs to ceiling. The procedure for conducting these climbs is the same regardless of what schedule is used or how it was determined. In performing an individual sawtooth climb, the actual climb should be started several hundred feet below the lower limit of the test altitude band. Prior to commencing the climb, the configuration should be completely checked and care should be taken to insure that instruments and equipment are function- ing properly. The technique involved in performing measured climbs requires more care and attention to detail than any other part of flight testing. However, this care is useless if the proper choice of weather is not made since climbing performance is very seriously affected by turbulent air thermals, vertical gradient in horizontal wind, and temperature inversions. Climbs should be undertaken only in the most ideal weather because data taken during unsatisfactory weather condi- tions tend to distort the remaining climb data. To avoid the possibility of actually zoom- ing into the test band and to insure stabilized flight through the band, entry into the saw- tooth should be accomplished in the following manner: During the climb the pilot must: (1) At an altitude several hundred feet below the lower limit of the test band, set the configuration and trim the airplane to hold, in level flight, the airspeed desired in the climb. t to wird (1) Maintain the indicated airspeed within plus or minus one knot of the specified value and hold ball-in-center wings level trim. (2) Keeping airspeed constant, smooth- ly advance the throttle until rated power for the test is attained. The airplane will then be climbing. (2) Maintain the proper power setting on the engine in the part throttle region. This will generally mean the maintenance of a prescribed RPM and torque. (3) Observe and note the altitude at which full throttle is reached. (3) Make sure that the airplane is stabilized in the climb as it enters the test band and start recording data (through the test band the speed variation should not ex- ceed 1 knot). Data is best taken with a photopanel; however, stop watch data plus pilot instrument panel readings may be re- corded on a kneepad to give adequate data provided the climb rates are not excessive. (4) Record the data specified in the section on data reduction at specified inter- vals except in the case where an engineer observer is on board or a photopanel is available for recording the data. (4) The procedure is repeated for at 6:36 The technique of entering the climb is as follows: 6:11 MEASUREMENT AND CORRECTION OF TAKE-OFF CHARACTERISTICS Provided the test conditions do not deviate excessively (20% or more) from standard, it is most convenient to correct take-off data by semi-empirical methods. The airplane is trimmed at the indicated climbing speed for the starting altitude of the climb with the propeller governor set at the proper RPM for the climb and sufficient manifold pressure to maintain level flight at the climbing speed. The throttle is then opened smoothly and the airplane speed is maintained constant until stabilized climbing conditions are reached. Correction of take-off data is required to account for (1) Non-standard wind conditions (2) Non-standard weight (3) Non-standard air density (4) Non-standard thrust The procurement of useful data commences at the altitude and time at which V., BHP, trim and configuration are stabilized. If the airplane is properly trimmed with the con- trol tabs the problem of holding observed airspeed within the allowable limits is made much easier. In some of the larger airplanes it has been found practical to utilize the auto- matic pilot in climb, with the necessary ad- justments to the longitudinal trim being accomplished as speed changes are required. The approximate numerical procedure commonly used to correct for the above listed deviations from standard is a ratio- exponential procedure using empirically de- termined exponents. An increment correction procedure is also available. The procedures are theoretically justified as follows: From Ref. 3, Eq. 10:22, the approximate take-off distance of an airplane traveling against a wind of velocity Vw is given by It is good practice to continue a climb in any one blower setting to an altitude as much as 2000 feet above the blower shift point. The climb is then discontinued and the air- plane altitude reduced to as much as 2000 feet below the shift point during the time that the shift is being accomplished. The climb in the next higher blower is begun be- low the shift point and continued up to the shift point for the next higher blower. In other words, the climb is inade in as many sections as there are blower ratios in the airplane, and the various sections are inte- grated to give the final climb curve. s : Sw + sa w (va - Video . Vw W (Vg - Vw) - sg Fm gFm W [Vg - VW]' 29Fm 6:39 where It is true that such a procedure supposes an infinitely fast blower shift in the final climb curve, but this procedure is considered to be the best representation of the airplane performance. It should be pointed out in conjunction with this procedure that the time lost in accomplishing the blower shift and the amount of climb which is repeated in the blower shift region should be kept to a mini- mum to avoid the necessity for weight cor- rection of the final observed climb data. 8 = take-off distance Sw - distance traveled through moving air mass • distance traveled by moving air mass Sa 6:37 W = gross weight V, VW Fm = take-off airspeed (getaway speed) : wind velocity : mean accelerating force during take-off run Next, consider the case of non-standard weight, with P and Fmw standard. In this case V, in Eq. 6:33 is a dependent variable and must be expressed in terms of W. Assume that the data have already been corrected to zero wind conditions, then Eq. 6:39 becomes g : acceleration of gravity WV S 29Fm or, since 2 w Vå Ideally, the parameter which is held con- stant during all take-offs regardless of ex- ternal conditions is the lift coefficient. There. fore, the correction procedure is based on the concept that the take-off lift coefficient is a fixed quantity. To obtain the form of the wind correction formula, consider for the moment that the weight, density and mean accelerating force are all fixed. Then, since PCLS 6:42 w 2 S gp CLSFM 6:43 woževi CLS W Using “sub o" for observed and “sub s" for standard, Eq. 6:43 provides the relation Ws Ss • So where CL is the take-off lift coefficient, we may express our constant W in terms of Vg (also a constant for the conditions of this analysis since Vg is fixed in value by Wip, and CL). )。 Wo 6:44 Thus, Eq.6:39 becomes į PVO CLS Eq. 6:43 also provides the relation for correcting for non-standard air density since Vg (which is a function of p as well as W) has been expressed in terms of p and W and does not appear in Eq. 6:43. Thus, assuming the weight correction has already been ac- complished and that Fm is not altered by density - S: [Vo - vu]'. 29Fm 6:40 If we let Ss . So () so : observed ground run distance and Ss - standard ground run distance for standard wind Vws 6:45 then, from Eq. 6:40, the ratio of ss to so is 2 Ss or so Vg - Vws Vg - VWO Vg – Vws Vg - vwo' In the preceding development, Fm has been held constant to permit isolation of the di- rect effects of wind, weight, and air density. Actually, changes in weight and wind produce second order changes in Fm, while changes in density produce a first order change in Fm. The second order effects are custom- arily accounted for by modifying the exponent 2 in Eqs. 6:41 and 6:43 to fit experimental Ss so 6:41 6:38 . data. The proper values of these exponents are discussed later in the chapter. employed as a means of obtaining standard power during test. The quantity Fmo itself is determined directly from Eq. 6:39 by re- writing this equation in the form Wo 2 As noted before, changes in density pro- duce a direct aerodynamic effect on take- off distance, which is essentially independent of the change in Fm due to density. Conse- quently, it is possible to handle changes in Fm separately from changes in the other variables. Fmo : [Vg. - Vw.)? 2950 where Wo, so, Vgo and Vwo are the observed test values, The important factor altering Fm is vari- ation of engine thrust due to non-standard atmospheric conditions, improper engine set- tings, or any of the other causes which in- fluence power plant operation. Proceeding as before, the effects of variation of Fm are best isolated by holding the quantities W, Vw, and and p constant and permitting Fm alone to vary. In this case, Eq. 6:34 provides As noted, the procedures developed here presume that the corrections made are all relatively small, and within the framework of this assumption, the foregoing work may be employed for "JATO" take-offs as well as normal take-offs even though the 'JATO" unit firing time does not equal the take-off time. sofmo. Fms Ss To conclude this brief analysis of take-off characteristics, the preceding work is sum- marized below in a collected form, and the data reduction procedure outlined. 6:46 But Fms differs from Fmo by an amount AFm, the deviation from standard, so that AFm : Fms – Fmo or Because the corrections for weight and density were obtained on the basis that the observed data had previously been corrected to zero wind conditions, the wind correction is made first and is followed by the remain- ing corrections. From an examination of E28, 6:41, 6:44, 6:45 and 6:47 it follows that the complete correction formula to zero wind standard is Fms 8 Fmo + AFM. Substituting in Eq. 6:46, we obtain Vgo Ws 1 Ss so -) ( ) Ss : So Voo AFM W - vwo AFm 1+ 1+ Fmo Fmo 6:48 6:47 where "sub o" indicated observed (as tested quantities) and “sub s" the standard quan- tities. For practical purposes, the quantity AFm may be taken as the difference between the thrust available on a standard day and the actual thrust available during the test. This increment may be computed using engine charts and standard engine correction pro- cedures or torquemeter equipment may be If it is desired to correct to a standard wind velocity at a standard weight, an addi- tional wind correction term is required based on standard getaway speed Vgs - 6:39 Since 2 w va PCLS The data on which the exponents of the wind and weight correction terms are based has been drawn from many sources and has been accumulated over a long period of time. Early test data (which are presented in a dif- ferent form than used here) are given in Fig. 216 of Ref. 4 and computed data in Fig. 217 of the same reference. By plotting the data of Fig. 216 in the form log (so/ss) vs. log (1 - VW/Vs ) we obtain the relation we have that Ws Ves Vg . Wo Ps ) i 1.85 Vgo Ws2 : Vgs Vgo ܘܘ ss so wo s - Vwo V90 6:49 6:51 (Note: V, is not a function of Vw or Fm.) Using Vgs defined by Eq. 6:49 and multi- plying Eq. 6:48 by the additional correction term Vgs - VWS Vgs It is believed that these data were obtained for tail skid type alighting gear having higher ground friction coefficients than found in present day airplanes. More recent data indicate exponents of 1.8 to 1.9 with some evidence favoring the value of 1.9. On the basis that the corrections are of small mag- nitude, using a value of 1.85 produces neg- ligible errors. Considering the weight correction term, various references give values of from 2.1 to 2.6 for its exponent. For modern aircraft a value of 2.2 or 2.3 seems most realistic. Thus the equation the following is obtained: M90 Vgs - Vws ss so Х Vgs Vgo - two Ws 1 ) 6:50 Wo ss W. 2.2 AFM - ( So Fimo Wo 6:52 where Ws Vgs : ( Wo vo Cewek ( en [V8o – vwol". . 's Wo Fmo 2950 (6:50 cont'd) In the preceding analysis it was mentioned that second order variations in Fm due to wind and weight effects had been ignored and would be accounted for by modification of the exponents of wind and weight correction terms. These modified exponents are dis- cussed below. is recommended. Ref. 5 indicates that the exponent of the weight term of Eq. 6:49 should be .476 rather than the theoretical 1/2. This indicates a modification of pilot technique due to change in gross weight rather than a change in any other factor. The data determining the value of .476 were, it is believed, obtained for conventional gear reciprocating engine air- craft, and as such are probably not directly applicable to tricycle alighting gears. Ac- cordingly, the factor of 1/2 is recommended pending further investigations. 6:40 Summarizing, Eq. 6:50 with empirical exponents becomes 1.06 1.85 2.2 Vgo Ws Pobs 1 ss so Vgs - Vws Vgo - Vwo ) Vgs Wo DFm Ps 1+ Fmo 6:53 where Ws Vesivo choses ) i and Fmo 8 Wo 2 g so [vco -vo) wo Vgo Wo Ps (6:53 cont'd) Test Procedures There are two basic approaches used to obtain take-off data; these are: (1) Determination only of total distance and end speed before take-off by causing the aircraft to travel over tar strips whose separation dis- tance is known. Knowing the separation distance and the time taken to cover the distance between adjacent strips, the speed just prior to take-off can be computed; more- over, the point of take-off is fixed between two strips by the disappearance on the ac- celerometer trace of blips and roughness caused by passage over the tar strips and runway surface. (2) Determination of a space time rec- ord of the entire run. Various other schemes have been proposed using limit switches on the gear to indicate full extension and so forth. Most of these can be made to work, if sufficient attention is paid to detail. For routine testing, the first procedure is all that is required, and the data are obtain- able in several fashions. The most obvious procedure is to estimate first the take-off distance from some fixed reference point, and then to place a motion picture camera at the approximate take-off point so that pictures of the actual take-off are obtained. Knowing the film speed (using a timing pendulum or stop watch in the field of view) and the dis- tance of the airplane from the camera, the take-off ground speed is obtained. Problems arise here in the determination of the exact point at which the airplane leaves the ground, and generally several observers must be employed to obtain a reasonable estimate of the take-off point. When more information is desired, space- time records can be obtained using a the- odolite tracking camera with or without a grid screen or some device similar to the new Fairchild Associates distortion cor- rected tracking camera. In the event that such data as take-off distance over a fifty-foot obstacle are re- quired, the use of a tracking camera is al- most mandatory. Analysis of Obstacle Take-offs Another procedure utilizes accelerometers mounted on the shock struts as a basis for determining the take-off point, and revolution counters attached to the wheels. The method is based on determining ground speed shortly - The problem of determining the ground distance required to clear a fifty-foot ob- stacle (or an obstacle of other height) under 6:41 The ground run portion of the data may be corrected by the procedures previously developed; however, the air distance re- quires further consideration. For this anal- ysis, the actual curved flight path shown in Fig. 6:25 as a dashed line is represented by a solid straight line also shown in the figure. The required wind correction on sa is apparent from the figure. Suppose that the take-off were conducted into a wind of ve- locity Vw, and a time "q" was required to cover the observed distance sao. The wind has no effect on the time, as will be seen on a moment's reflection. However, it shortens so by an amount Vwt. . Thus, the observed air distance is cor- rected to standard no wind conditions by the relation SOS Sa. + Vwt. 500 6:55 standard conditions is more complex than determination of standard ground run itself. There are a number of reasons for this with the principal one being the large possible variation in pilot technique. However, pro- vided one is satisfied with reasonable ap- proximations it is possible to correct to standard distance required to clear an ob- stacle of a given height. The problem is analyzed below. First, it should be realized that the prob- lem of achieving minimum distance from the start of the run to over the obstacle is dif- ferent from the previously analyzed problem of minimum ground run. In achieving mini- mum distance on the ground, the pilot is not nearly as concerned with the initial climb characteristics as he must be when he desires to clear an obstacle at the end of the run- way. For good climb-out characteristics, it may pay to remain on the deck longer when at- tempting to clear an obstacle than when seeking minimum ground run. The reasons for this are simply that the speed for best climb angle is generally somewhat greater than the minimum getaway speed, and the possibilities of performing a zoom maneuver are better with more speed available. This means that the ground run data ob- tained for minimum take-off distance on the deck is not usable for obstacle take-offs even for the ground run portion of the ob- stacle take-off runs. The procedures for obtaining data may vary; however, the use of some type of photo- tracking device which provides a space time record is mandatory. In addition, records of wind velocity, air density, weight, etc. should be kept, The actual distance component parallel to the ground from the start of take-off to over the obstacle is best broken into two parts, the ground run distance Sg and the projec- tion of the flight path from getaway to over the obstacle so. Thus, the total distance is Similarly, the corrected value of the aver- age flight path angle is given by tan ys . h Sao + Vwt and since tan yo Yo : h/800; ton Yo (1+ Vw I ) tony's Sao 6:56 Assuming that the observed data have been corrected for wind effects, the remaining corrections can be made on the basis of Eq. 9:25 of Ref. 3, which is rewritten below in modified form. T CO T CL w sin y : Со nCL s.sg + sa 6:54 6:57 6:42 h VW so Flight Path Sg a h : Obstacle Height S td -Take-off Point Fig. 6:25 where 9:26 of Ref. 3 for CL, whence n* : 4 for values of y less than 30° 2 WCOS Y CL : . n : 2 for values of y between 30º and 45° pVPS (Note that y is not a function of air density except as this factor alters the thrust.) The value of Co corresponding to CL may be obtained from wind tunnel test data or performance flight test data. For propeller driven aircraft, the power required curve serves the basis for determining the equation for the lift drag polar, as ) In Eq. 6:57 the observed value of y is the angle between the straight line approxima- tion to the flight path and the horizontal and is the approximate average value for the actual flight path. cz CO : CDE + TAR The average value of CL, at which the aircraft flies during the time the distance So is traversed, is obtained by solving Eq. from which CD is readily calculated for any test value of CL. * Note: This improves the approximation given in Ref. 3. The variation in observed thrust from standard may be obtained from the engine 6:43 Procedure for Correcting Obstacle Take-off Data charts using a reasonable assumed value of propeller efficiency. Letting AT be the thrust increment, defined by the relation Ts : To + AT + 6:58 8 (a) First reduce the ground run data to standard by means of the corrections dis- cussed earlier in this section. (b) Reduce the observed air distance and flight path angle data to no wind conditions by Eqs. 6:55 and 6:56 where Ts : standard thrust To : observed thrust SOS Sao + Vwot To + AT Ws tanys • ton Yo 1 sinyo Vwt CO CL Со n CL 1 + soo (c) Compute the average value of CL dur- ing the climb-out using the relation 6:59 Then 2 Wocos Yo CL Pobs tas sin Ys To + AT CD Ws CL то Ср Wo CL sinyo - 6:60 where the value of Yo is the observed value corrected to zero wind conditions. (d) Knowing the average test value of CL, obtain the value of the average drag coefficient from level flight performance data. where Ws: standard gross weight, W. = observed gross weight. From Fig. 6:25 (e) Using Eq. 6:57 solve for the average T/W ratio, and knowing the gross weight, obtain the average thrust during the climb- out. h Soo tanyo CO ().." sin yo () CO + Yo 1 - пCL CL h SOS tonys so that (f) Determine the deviations from stand- are thrust from engine charts and compute the standard no wind value of y from Eq.6:60 TO + AT Со Ws SOS ton Yo CL 3 Sinys sin yo 500 To CD Wo CL tanys 6:61 6:44 (g) Compute the standard no wind air distance from Eq. 6:61 for weight are attempted inasmuch as the- oretical considerations alone do not seem to give proper answers. Accordingly, landing tests should be made at weights as close as possible to standard. tonyo SOS Soo tonys (h) If data on obstacle take-offs into standard wind are required, compute the average true air speed under standard con- ditions from Since the landing run analysis is precisely the same as the take-off run analysis, ex- cepting that there is no effect due to chang- ing engine operating conditions (except for airplanes with propeller pitch reversing mechanisms), Eq. 6:53 may be rewritten for the ground run during landing in the following form: 2 Wscos Ys Voss PS CLS INDO 1.85 1,65 Sgs sgo (VDs VDS - Vws VDO - VW x - (i) from Compute the new air distance time 2.2 Ws Pobs Ps Wo 6:62 sos ts : Viass when Sgs 8 : standard ground distance; (j) Compute the wind correction sws ts Vws and subtract from the distance in step g to give air distance to clear an ob- stacle in the standard wind sgo : observed ground distance, For aircraft equipped for propeller pitch reversal an additional thrust correction may be added. sas SO 59 SWS (k) The total corrected distance from the start of motion to over the obstacle is given by either the sum of a + g or a + j. Eq. 6:62 provides for the ground run cor- rection. If landings over obstacles are to be analyzed, we must also consider the air distance corrections. According to Ref. 2, the only correction on air distance which can properly be imposed is that accounting for the effect of wind. Thus, if 6:12 MEASUREMENT AND CORRECTION OF LANDING CHARACTERISTICS sao = observed air distance - Sas : standard air distance Actual test measurements of landing run characteristics are accomplished using the same equipment and similar procedure as those used for take-off measurements. LOO : observed time for sa VWO = observed wind velocity Because of the wide variation in pilot technique encountered during landings, re- peatable data for landing runs are more difficult to obtain than for take-off runs. This shows up particularly when corrections Vws : standard wind velocity, then - sog soo + (Vwo - Vws) too 6:63 6:45 Combining Eqs. 6:62 and 6:63 voo 1.85 Vws 1.85 Ws 2.2 Pobs : + sos Sgs Ss sgo \ VOS vo var en meg) ** () VDO Vwo Wo Ps + SOO + ( Vwo - VWs) too 6:64 The data reduction process required with Eq. 6:64 is similar in all respects to that required for landing analysis, and, therefore, need not be discussed in further detail here. utilized. In this case, the test data obtained over a wide altitude band may be expected to display scatter over and above that attrib- utable to propeller efficiency variations and similar effects considered earlier in the chapter. 6:13 CONCLUDING REMARKS In this chapter we have summarized in- formation on performance testing of recip- rocating engine aircraft, with all important phases being analyzed. Special topics, such as engine cooling investigation, were not in- cluded since these are matters which depend more on specifications than on fundamentals. Finally, we note that although the effects of thrust produced by the engine exhaust stacks were not considered here, significant amounts of power may be generated by this thrust even though an ejector system is not If desired, exhaust thrust may be con- sidered in the data analysis using methods similar to those described in Chapter 5. However, it should seldom be necessary to go to such extremes; rather, an attempt should be made to separate high and low altitude data by reducing all data taken above 20,000 feet (for instance) to a 20,000-foot level and data obtained at altitudes lower than 20,000 feet to sea level. This procedure provides two sets of “generalized" data curves and generally may be expected to reduce the scatter of data. ( 6:46 REFERENCES 1. Dommasch, D. O., “Flight Test Manual, Part I," Revised Edition, Preliminary Copy, NATC, Patuxent River, Maryland, August, 1953. 2. Herrington, R. M. and Shoemacher, P.E., AFTR No. 6273, Revised January, 1953, "Flight Test Engineering Manual.” 3. Dommasch, D. O., Sherby, S. S., and Connolly, T. F., “Airplane Aerodynamics," Pitman, 1951. 4. Dichl, W. S., “Engineering Aerodynamics,” Ronald, 1936. 5. Forry, J. E. and Horton, C. F., “Manual of Flight Test Methods and Procedures, Part I," Revised 1 June 1948, NATC, Patuxent River, Maryland. 6:47 1 AGARD FLIGHT TEST MANUAL VOLUME I, CHAPTER 7 TOTAL ENERGY METHODS Ву Kenneth J, Lush AFFTC, United States Air Force PARTI Bernard Davy Centre d'Essais en Vol, France PART II D, O. Dommasch Princeton University FOREWORD and PART III J. F. Renaudie Centre d'Essais en Vol, France APPENDIX A 3 VOLUME I, CHAPTER 7 CHAPTER CONTENTS Page FOREWORD TERMINOLOGY PART I, OPTIMUM CLIMB THEORY AND TECHNIQUES OF DETERMINING CLIMB SCHEDULES FROM FLIGHT TEST 7:1 7:1 INTRODUCTION 7:2 7:2 A SIMPLIFIED THEORY 7:2 7:3 END CONDITIONS 7:5 7:4 AN EXACT APPROACH 7:5 7:5 MISCELLANEOUS CASES 7:8 7:6 EFFECT OF OPERATING CONDITIONS 7:9 7:7 DERIVATION OF THE OPTIMUM CLIMB SCHEDULE FOR THE PARTICULAR CASE 7:9 7:8 INSTRUMENTATION AND ANALYSIS 7:11 (a) Climbs 7:11 (b) Level Accelerations 7:11 7:9 STANDARDIZATION OF RATE-OF-CLIMB CURVES 7:15 7:10 REFERENCES 7:15 APPENDICES 1 FLIGHT TEST TECHNIQUE FOR ACCELERATED LEVELS 7:16 II ANALYSIS OF AIRSPEED INDICATOR TIME HISTORY FROM LEVEL ACCELERATION 7:17 PART II, ANALYSIS OF THE OPTIMUM ENERGY CLIMB SCHEDULE 7:19 7:11 INTRODUCTION 7:19 7:12 DEFINITIONS 7:19 CHAPTER CONTENTS (Continued) Page 7:13 THEORETICAL ANALYSIS OF THE OPTIMUM CLIMB SCHEDULE 7:20 7:14 GEOMETRIC REPRESENTATION 7:21 7:15 PROPERTIES OF THE GRAPH 7:23 7:16 DISCUSSION OF ASSUMPTIONS 7:24 7:17 CORRECTIONS 7:26 7:18 PRACTICAL DETERMINATION OF THE OPTIMUM ENERGY CLIMB SCHEDULE 7:27 7:19 CONCLUSIONS 7:31 PART III, CORRELATION OF THE ENERGY CLIMB ANALYSES 7:32 7:20 REVIEW OF PRECEDING ANALYSES 7:32 7:21 THE THIRD METHOD OF EXPRESSING THE CONDITIONS FOR OPTIMUM ENERGY STORAGE 7:33 7:22 PRESENTATION OF THE THIRD SET OF CONDITIONS IN TERMS OF V AND h. 7:35 7:23 SUMMARY 7:36 APPENDIX A, PROCEDURES FOR DETERMINING THE ENERGY CLIMB PERFORMANCE OF TURBOJET AIRCRAFT 7:38 7A:1 INTRODUCTION 7:38 7A:2 COMPUTATION OF TAPELINE OR GEOMETRIC ALTITUDE AND ENERGY HEIGHT 7:38 7A:3 DETERMINATION OF W: dhe/dt 7:40 7A:4 CORRECTION PROCEDURES 7:41 7A:5 DETERMINATION OF CORRECTED w= dhe/dt AND CORRECTED TIME-TO-CLIMB 7:45 7A:6 SUMMARY OF PROCEDURES 7:46 7A:7 CONCLUDING REMARKS 7:48 VOLUME I, CHAPTER 7 TOTAL ENERGY METHODS FOREWORD Starting with the original German concept of energy height, separate analyses of the optimum energy climb schedule have been carried out independently by the several NATO nations. Because of different approaches to the problem and the use of different notations, at first glance there appear to be some basic variations in the suggested procedures. However, the several analyses do arrive at the same fundamental conclusions, and it is worthwhile to point this fact out at the beginning of the chapter. The actual correlation of the methods is presented in Part III of this chapter. In an effort to preserve originality, the papers, sub- mitted by Mr. Lush of Wright Field (which also reflects the British approach to the problem) and Mr. Davy of the Centre d'Essais en Vol of France, have been left in essen- tially the form submitted (although certain idiomatic changes were of course required in the translation of the latter paper). No attempt was therefore made to correlate the two papers by modifying the papers themselves. Rather, it was felt that a more satisfactory correlation could be achieved by adding a third section to the chapter, which would serve to summarize the joint thinking expressed by both authors and to introduce a third approach to the prob- lem. ) The order of appearance of the first two papers in this chapter was determined by the date of their arrival, rather than by any other considerations and therefore no signifi- cance should be attached to the sequence. TERMINOLOGY h True Altitude he, H Energy Height V True Airspeed along Flight Path E Total Energy W Weight m Mass W dhe/dt t Time F, T Thrust, also Temperature Ft Excess Thrust o Flight Path Angle N RPM T Temperature n Load Factor RO Rate of Climb with Zero Flight Path Acceleration PART I OPTIMUM CLIMB THEORY AND TECHNIQUES OF DETERMINING CLIMB SCHEDULES FROM FLIGHT TEST ABSTRACT Consideration is given to the problem of determining the optimum technique by which an airplane will pass from given initial con- ditions to assigned final conditions in the minimum time. It is shown that introduction of the concept of energy height enables a simple approach which takes proper account of the kinetic energy of the airplane and appeals to physical intuition. It is concluded that on the quasi-steady climb the speed height schedule should be such that at each level of total energy level attained, the speed is that which gives the most rapid increase in total energy. This is equivalent to the condition ORO av V ORO g on . Other climb cases are discussed briefly. It is concluded that the climb technique for minimum time will give a range performance very close to the best possible. The problem of achieving the greatest speed in a given time is discussed briefly. Determination of the optimum schedule from flight test is considered. It is concluded that this is most conveniently done from level accelerations or from full climbs to a range of speed height schedules. The latter have advantages if it is reasonably practicable to make a performance climb on most flights, but it is essential to have reliable wind grad- ient corrections, either from weather data or from radar tracking on a plotting table, Otherwise level accelerations, in which the airplane accelerates horizontally in substan- tially level flight with the engine at the re- quired rating, are better. Instrumentation and analysis for level ac- celerations is discussed at length, methods considered and the following conclusions reached: (a) Accelerometer. Convenient if satis- factory instrumentation is achieved, but very high sensitivity is required. As yet unproved. (b) Airspeed Indicator and Clock. Incon- venient, but available and in successful use. Analysis routine given in appendix. (c) Time Trace of Dynamic Pressure, Available but unproved. Should work and is much more convenient than (b) and probably as good as (a). (d) Rate-of-Climb Indicator Connected to Total Pressure. Unproved. May be very convenient for a quick check, but unlikely to be suitable to give absolute values. (e) Air Temperature Thermometers with Different Recovery Factors. Unlikely to be sensitive enough. Probably not sufficiently superior to (c) to be worth developing. where V is the true airspeed, h the tape- line height, and Ro the rate-of-climb which would be attained at the revelant speed and height with zero longitudinal acceleration. The transition between the end conditions and the quasi-steady climb is discussed, and it is concluded that it is sufficient to ensure that these phases are as brief as possible. This requires horizontal or near horizontal acceleration at the end of a climb. The above approximate approach suffices until, with very fast climb, limited maneu- verability, and low aspect ratio the technique in the initial and final phases of the climb becomes so important that it must be deter- mined quantitatively. A more complex ap- proach is then required. A possible method is indicated, but it appears unlikely that any general solution is possible, and numerical solution is not attempted here. 7:1 7:1 INTRODUCTION In the days of the biplane fighter the climb consisted essentially of increasing the poten- tial energy of the airplane, any incidental changes in kinetic energy being small. It was then customary, and entirely legitimate, to ignore the kinetic energy of the airplane both in selecting climb schedules and in predict- ing climb performance. these reduce to differential equations of the second order and second degree in the load factor (lift/weight) which have varying coef- ficients. Routine solution of these equations is not practicable, and does not give much physical insight into the problem. For the moment, therefore, we will consider a simple theory. The climb equation for an airplane in near straight flight in still air may be written in the following forms, assuming a consistent system of units: The advent of the monoplane of relatively high wing loading increased climb speeds and also increased the speed range between climb speeds and maximum level speeds. This in- crease was sharply accelerated by the adop- tion of jet propulsion. W dV V(F-D) : W. dh dt S + ·V. g dt 7:1 V dV dh : W: dt ( (+) 9 dh 7:2 Consequently, the kinetic energy changes associated with the acceleration from take- off to climb speeds, with the quasi-steady climb and with the passage from climb to, for example, maximum level speed, became im- portant. Attention was first directed to the second of these, and it became normal prac- tice to make allowance for the change in kin- etic energy with altitude when predicting per- formance in the quasi-steady climb. It was not until German engineers introduced the concept of "energy height", fairly late in World War II, that a simple but satisfactory presentation of climb performance was ob- tained d : W wit (nt h + o dt 7:3 where V : true airspeed F = total net thrust The aspect of climb performance of most general interest is the achievement of a pre- scribed change of height and speed in the least practicable time. This will be consid- ered at length below, and brief reference will be made to other problems. Given a climb technique, the establishment of climb per- formance is simple enough. We will, there- fore, study the theory of selecting climb tech- niques. D = drag h altitude above some arbitrary datum 10 W = weight 7:2 A SIMPLIFIED THEORY In a given atmosphere and at a given en- gine rating, F and D are functions of W, V and h. If initial gross weight is now fixed we may omit W, as effect of climb technique on weight variation with height would not be ex- pected to have any appreciable effect on the optimum technique. It is possible to write down the exact equations for the technique for minimum time between specified end conditions, but 7:2 With the above in mind, we now write and hence dh dV ffv, moet on n, dhe , di dh 11,2 hezo V2 x (V, he) ne, ovi 7:6 and proceed to find the climb technique which minimizes the time required to pass from a speed V and a height h, to a speed V2 and a height h2 by minimizing the integral h2o V2 - Minimization of this integral is in principle very simple, and is achieved by making X as large as possible. At those points on the climb away from the influence of end condi- tions, the necessary condition for maximiza- tion of X is ах dh 1,2 hvi f( V, h, ) dV dh o v 7:7 using the Euler-Lagrange conditions. How- ever, a clearer physical picture is obtained if we substitute for h in terms of the quantity "energy height", he, defined by the equation v h + where the partial differentiation is, of course, made at constant he. That is, the schedule of the quasi-steady climb is to be such that at each level of total energy achieved, the rate of increase of total energy is as large as possible. (Such a schedule will not neces- sarily give the most rapid increase of height, or potential energy.) : he siä 7:4 Energy height is clearly the height equiv- alent of the total energy of the airplane. Its use amounts to a recognition of the fact that the kinetic and potential energies of an air- plane are fairly readily interchangeable and that it is the total energy, rather than either the kinetic or the potential energy, which must be increased as rapidly as possible to mini- mize climb time. Substituting from Eq. 7:4 into Eq. 7:3 we have, of course, This conclusion may be drawn semi-intu- itively if it is first agreed, as indicated above, that the division of the airplane's energy between kinetic energy and potential energy can be adjusted fairly rapidly by climbing or diving, and is therefore of no consequence except at the beginning and end of the climb. dhe Eq. 7:7 is simple in appearance, but is not readily applied in practice as both calculated and experimental values of X are most read- ily obtained as curves of X against V at con- stant height. It is therefore desirable to cast the equation in another form, VIF-D) W () dt With our assumption of constant weight, F and D are functions of V and h only, and hence, from Eq. 7:4, of V and he only. We may therefore write As X is a function of V and he, and he is itself a function of h and V, we may write the identity dhe dhe = X(V, hel x(he, V): $(h, v). dt dt 7:5 7:3 But We may now rewrite Eq. 7:7 in the more convenient form 1 2 ola he Sht дф у дф • O ду 9 on 7:14 By definition we therefore have ох ax dV+ dhe = dx . av ane Given computed or experimental curves of $ against true airspeed V for various heights we may estimate 0/0 h, which is fairly small, and thus deduce the slope of the rele- vant curve at the optimum speed. дф дф = d$ = dV+ av dh dh 7:8 and It is of interest at this stage to compare the schedules deduced from Eq. 7:14 with those deduced by earlier methods. Two meth- ods have been used. These are, in historical order: (a) to take as the schedule speed, the speed which gave maximum rate of climb in a “sawtooth climb,” at constant indicated airspeed, through the relevant height; and (b) to take the speed at which o/a V is zero. dhe = dh + tř. av dV 7:9 Substituting from Eq. 7:9 into Eq. 7:8 to eliminate dhe , we have ox ах Ох V + 9 .) DV + dh ane av ahe дф дф av dV+ ola dh. Consider (b) first. With air breathing en- gines 04/0h is negative, so the correct op- timum speed as given by Eq. 7:14 will be higher than that given by (b). In the absence of compressibility effects, this difference usually amounts to between 5% and 10% in speed, which is insufficient to cause any no- ticeable loss in climb performance. If the optimum speed is dictated by compressibility effects, the two conditions will give closely similar speeds and the loss incurred by using condition (b) is again very small. 7:10 Equating coefficients of dh and dV we have ax . ola ane 7:11 and Method (b) is thus adequate and may be used as a convenient approximation instead of using Eq. 7:14. This, of course, does not in any way detract from the energy height approach to the climb problem, as it is only by reference to the exact condition that we can evaluate the approximate condition. дф ol Vax + ov 9 dhe av 7:12 Eliminating a X/he we now have ах at v дф However, condition (a), which was used for many years with slower airplanes, is more seriously in error. When climbing at constant indicated airspeed, the proportion of the available thrust power which goes into in- creasing the kinetic energy of the airplane ov ду 9 on 7:13 7:4 increases with increase in forward speed. Consequently, the sawtooth climb curve peaks at a lower speed than the curve of $ against V at constant height and condition (a) gives an even lower speed than condition (b). airplane approximates the required final value. The airplane should then be pushed over into a moderate dive so as to level out at the required altitude and speed. The increased negative error (for air- planes with air breathing engines) leads to an appreciable loss in climb performance of the order of 5% to 10%. This is not tolerable. Thus, condition (a) must be rejected. This does not, of course, mean that sawtooth climbs may not be used to determine optimum climb schedules, only that correct interpretation is required. This is a vague statement, but fortunately quite a moderate dive angle will result in a short acceleration period and should, there- fore, give results approximating the best possible. If the required final speed is low, or if the aim is to achieve altitude quickly regardless of speed, the climb will, of course, be ended by zooming and leveling off at the required altitude. 7:3 END CONDITIONS It may be noted that with some airplanes which operate up to, but not past the transonic drag rise, climb and maximum speeds may be very close together and acceleration in level flight rapid. It would then be sufficient in many cases to level off at the required altitude. We have derived above a condition which accurately determines the optimum schedule for the quasi-steady part of the climb. This simple theory cannot determine quantitatively what the techniques of entering and leaving the climb should be. Useful qualitative conclu- sions may be drawn, however. It is worth remarking that with present military airplanes, no precise meaning can be attached to the time to reach operational altitude unless initial and final speeds are quoted. The initial condition is probably best fixed by quoting the time taken to reach climb speed from the start of the ground roll. The final condition is arbitrary. One may quote time to reach altitude at climb sched- ule speed provided that it is made clear what the end condition is. During these parts of the climb, the rate of increase of energy is necessarily less than the maximum possible. The time spent in these unfavorable conditions should, there- fore, be made as short as is practicable. The extent to which this aim is realized is limited by the maneuverability of the airplane. Its acceleration cannot exceed that in a vertical dive and is further limited immediately after take-off by the presence of the ground, while its deceleration is limited to that in a verti- cal climb. Further, it can only change the angle of its flight path at a limited rate, par- ticularly at high altitude. The above qualitative approach is likely to suffice for airplanes of the class presently operational, For supersonic airplanes of very low aspect ratio, however, it may be necessary to adopt a more quantitative approach. This will be discussed briefly in a later section. 7:4 AN EXACT APPROACH It is clear that after take-off, the airplane should be kept at as low an altitude as is practicable until it approaches the climb speed, then pulled up fairly quickly into the climb. If the final speed is to be much great- er than that given by the climb schedule at the final altitude, the climb should be continued past that altitude until the total energy of the We will now examine briefly a more exact approach to the problem, which will take into account the limited longitudinal acceleration 7:5 and and deceleration available and also the varia- tion of induced drag with normal acceleration. V. dy dy dt 8 용 ​la . 비 ​9 dhe 9 dt dhe We will retain he as the variable instead of h, inasmuch as he increases monotonical- ly with time over the whole climb in almost all cases of interest, and we thereby avoid difficulties with double values. $ (n - cos y, 7:18 We may write We now denote total differentiation with respect to he by a dash (for example, dy/dhe = Y'). Also we write dhe V (F-D) dt W 5 + $ sin y : f(V, he,n) 8 = - \ IV, V, he,n,y) 7:15 in a given atmosphere and at a given engine rating, where n is the load factor, defined as the ratio of the lift to the gross weight, řx'- $ln - cos y) = XIV, he, n, y, y'). > Integrating between initial and final con- ditions we may write Then our problem is to minimize the integral dt 11,2 dhe dhe sød $ dhe 3 subject to the additional conditions $=0=X. s'o V, ne,n) dhe. 7:16 We write F: + 4,W + 12x We have auxiliary conditions concerning the longitudinal acceleration and the rate of pitch, which we we may write in the form 7:19 dt where , and 2 are functions of he which are to be determined. 1 O ov dhe DV dt g dhe D $ { Standard text books on variational calculus show that a necessary (but not sufficient) set of conditions is sin y {; } W where y is the angle of climb ola : 0 av dhe lov olo to 15 • {+ - () () d $ sin y = 0 an dhe on 7:17 7:20 7:6 d af d dhe ar dhe d OF ela ( 105) (125 . +42 (+ - + siny). : 0 dhe ду x2 $ 7:26 g (7:20 cont'd) Substituting from Eq. 7:21 through Eq. 7:26 into Eqs. 7:20 we have, after rearrangement, where V, V', n, n', Y and y' are to be trea- ted as independent during differentiation. . PLY + 1, sin y - 121n-cosy) , ys} av Considering each function in turn, we have + t OF av 1.0. + 12 j In-cosy) - дф OV + + sin Y 9 히 ​긍 ​ov 7:27 + 0 % -12 - do (n-cosy) av FE+ ( ) na iyo o $ {1+11 sin 7-12 (n- cos x)} , Y y +d2. in-cos y), 1 + 1, sin y -d21n-cos coom) - 2 - 0 : 0 on 7:28 3 In-, X2 1, $ cos y 7:21 - ه - . 0. V 7:29 d OF dhe love (0) : dhe : “Ai 7:22 Eq. 7:29 enables substitution for , in terms of dę and 'g. After differentiation it also enables substitution for ti. We then have one equation in ' , l'2 and 12 and one in X'q and 12. Both equations will have varying coefficients. Solution of these equations in practice would probably be best made by some step-by-step process. af : дф on Y 비​" on {{1+, sin ) - 121n- cos 1)} -$12, 7:23 It is of interest to apply to the above equa- tions the simplifying assumptions used in the preceding section. We assume that is in- dependent of n, and deduce from Eq. 7:28 that (5) 。 8 (0) : 0, dhe dhe 7:24 $d2 = 0, OF that is, 1, $cos y - 12$ sin y ay da :0. 7:25 7:30 7:7 As this is to be true throughout the climb, we also deduce that examination will be made which will sug- gest that a precise treatment is not worth- while, dà = 0. 7:31 Substituting in Eq. 7:27 and Eq. 7:29 we then have Consider a modification of the climb tech- nique for minimum time in which the airspeed at each point on the climb is increased in the constant ratio (1 + i). The fuel flow at each height will be sensibly unchanged, so the dis- tance covered for each pound of fuel consumed will be increased by approximately 1 times average miles/lb. on climb. дф {1 + 1, sin Y} 1+ } - o g ду 7:32 1,0 cos y = 0. 7:33 On the other hand, the time required to reach cruising height and speed, and hence the fuel used on the climb, will be increased by an amount of the second order in . The cruising range will be reduced by the product of this increase in fuel used on climb with the average miles per pound cruising. From Eq. 7:33 either COS Y O, y : $ 90° 7:34 Or , = 0 Thus it is evident that at least a small in- crease in the climb speed schedule will be profitable. Examination of particular cases tends to show, however, that the most profit- able increase in speed is only around 5% giving an increase in range of rather less than 5% of the distance covered on the climb, which is itself only a fraction of the total range. and hence from Eq. 7:32 дф : 0. ду 7:35 We thus have the two solutions discussed in the preceding section, namely Thus it would appear that only small im- provements in range can be achieved by de- parting from the climb schedule for minimum time and that a precise examination is hardly worthwhile. a. (oflav) = 0 during quasi-steady climb b. vertical ascent or descent to pass be- tween end conditions and quasi-steady climb. 7:5 MISCELLANEOUS CASES A third problem which may arise is the achievement of the maximum possible speed, or Mach number, in a given time. This prob- lem will be most likely to arise with air- planes having high induced drags at high lift for which a more exact approach, similar to that outlined in Section 7:4 would probably be required. It could be stated in the form of maximizing the integral Determination of the climb technique for minimum time, considered above, is the prob- lem of most general interest. We will, how- ever, consider some other cases briefly. 11 First, we will consider the climb tech- nique for best range. An exact treatment will not be attempted, but an approximate dt (dt/ dv) 7:8 subject to the same sort of continuity condi- tions as those of section 7:4. Solution of this problem will not be attempted here. Ro against speed. When this is done, the optimum technique for standard conditions is readily deduced. Even when this is not done, it is evident from the above discussion that standardization of the climb schedule is rare- ly necessary. If it is required, it may be done by methods given in R & M 2576. 7:6 EFFECT OF OPERATING CONDITIONS It is of interest to know over what range of operating conditions a fixed schedule of indicated Mach number or indicated airspeed against indicated height may be used without a serious loss in value of performance. 7:7 DERIVATION OF THE OPTIMUM CLIMB SCHEDULE FOR THE PARTICULAR CASE We are presently concerned with a quant- itative determination of the optimum climb schedule on the quasi-steady climb only. This schedule is therefore defined by the equation This matter is considered in British Aero- nautical Research Council Report R & M 2576 (Ref. 2). It is shown that the optimum climb technique of a jet airplane is very insensitive to gross weight, but that in the absence of compressibility effects on drag, it does vary considerably with thrust, and hence, with air temperature. However, the curves of rate of climb against airspeed are very flat-topped, and considerable departure from the optimum is possible before a serious loss of perform- ance is incurred. ORO V ORO av g ah 7:36 This is the exact condition. It is for con- sideration by the test engineer whether he will use this or the approximate condition It is shown in R & M 2576 that within approximately the ranges of gross weight, engine speed and air temperature tabulated below, the loss in performance which will result from keeping to a fixed climb tech- nique will not exceed the greater of 1% and 1/2 ft. per sec. ORO : 0. av 7:37 Limits of Variation Parameter Low Altitude Near Ceiling Gross weight Engine speed Large 3% 15°C 12% 3% 10°C Given a family of curves of Ro (or of ex- cess thrust power) against airspeed or Mach number for various heights, derived from per- formance estimates or from flight tests, the schedule defined by Eq. 7:37 is very readily deduced. However, if the curves are against true airspeed, the exact solution is itself readily deduced by a rapidly convergent ap- proximation process as follows: Air temperature The loss in performance varies as the square of the departure of the operating condi- tions from those appropriate to the schedule used. When the optimum schedule is deter- mined by the onset of compressibility effects, It will be insensitive to changes of thrust. (a) At each height, take the speed for maximum Ro as a first approximation to the optimum speed. Estimate Vg times Rolah at this speed, by cross plotting if desired. (This curve is often nearly linear. Inasmuch as a rough value of V/g times Oro/dh suffices, it would usually be satisfactory to assume a Rolah = ARO/Ah, where ARO is Some test units standardize the curves of 7:9 should, if possible, be flown perpendicular to the wind gradient, the change in height between adjacent curves for heights separated by Ah.) (b) Find the speed at each height at which the slope of the Ro - V curve is equal to the corresponding value of V/8 times o Rolah. (c) Plot this speed against height. It will be found that a second approximation to the optimum speed is not necessary. For pre- sentation to the pilot, the curve must be con- verted to one of either indicated Mach number or indicated speed against height. An attractive technique of recording and analysis, if suitable equipment were readily available, would be to track the airplane by radar on a plotting table and derive the rate of change of true speed with height from the radar record. This technique is being con- sidered for check climbs. In view of the above, the problem of the flight test engineer is to deduce curves of Ro against V, at constant values of height, from flight tests. Such curves may be deduced from sawtooth climbs, from level accelera- tions (as is routine British practice for jet airplanes), or from full climbs covering a range of schedules which bracket the optimum speed at each altitude. Level accelerations avoid or reduce most of the difficulties mentioned above in connec- tion with sawtooth climbs. They need less flying for equivalent accuracy, are negligibly affected by vertical gradients in headwind, and can be made satisfactorily when the available rate-of-climb is very high. Pilots usually require some time for familiarization, but not a great deal. Bad runs are sometimes obtained, often for no obvious reason, but once the pilot has practiced the technique, the pro- portion of bad runs is not too high, Sawtooth climbs have been in regular use for this purpose for many years, and can still yield satisfactory data. However, they have several disadvantages. For example: For equivalent precision, the flying time required was not more than one-half of that required for sawtooth climbs. One level ac- celeration run takes about as long as one saw- tooth climb run, but four runs at most should be sufficient at each altitude and configura- tion. The main difficulties arise frm rather unsatisfactory instrumentation, a problem discussed in the next section. (a) They require a long test program, much of it at high engine output. (b) On jet airplanes, which climb best at high forward speeds, they are susceptible to large errors from wind gradients or from small irregularities in flying. (c) They are not too practical at rates of climb exceeding, say, 6000 ft. per min., be- cause a large height range must be covered on each climb to get good data. The essential points of the test technique are to fly as smoothly as possible through the important part of the speed range, maintain- ing the engine speed (or whatever determines the engine output) as constant as is practic- able. Tests are flown at constant indicated pressure height to simplify flying and elimin- ate static pressure lags, and correction made for any inclination of the flight path resulting from changes of position error with speed. Objection (b) can be largely overcome by suitable flight planning and analysis and by careful flying. The airplane heading should be reversed between runs, as this will show up any wind gradient effects present as a nearly symmetrical scatter about the mean. Care must be taken to enter the sawtooth climb with the speed thoroughly stabilized. At low altitudes it may be convenient while setting up engine speed either to extend dive brakes or to climb from below the intended test altitude. At high altitudes acceleration is slow and the pilot will have ample time to set up his engine speed. If speed variations are present, the anal- ysis should take account of them. The tests 7:10 Weight corrections for induced drag are usually small except near the ceiling, but if their value is in doubt, they can be eliminated by flying the tests at a constant ratio of lift to air pressure. With current operational airplanes, the effect of flight path angle on induced drag is small because high angles of climb are available only at low lift coeffic- ients where the induced drag is low. acceptable. This is presently under consider- ation by the U.S. Air Force Flight Test Center. The analysis should preferably be based on the rate of change of total energy of the air- plane, allowing for test wind gradients and test air temperature gradients. The latter may vary considerably from standard near the ground in very hot or very cold areas. (b) Level Accelerations For airplanes with which this is no longer true (for example, rocket-powered airplanes of low aspect ratio), it may be necessary for this and other reasons to re-examine the climb problem using the more exact approach discussed in section 7:4. The flight test tech- nique is discussed in slightly more detail in Appendix I. Instrumentation for measurement of level accelerations presents some difficulties, in- asmuch as there is not yet a proven method which is entirely satisfactory. The choice of method may depend on whether the tests are solely to give optimum climb technique or whether they are also to give standardized curves of Ro against speed for inclusion in a report. Instrumentation which has been suggested includes: a. Accelerometer A further method which has been used a little is to make full climbs to a range of schedules sufficient to cover the interesting range of airspeed at each height (or each ratio of weight to air pressure). To fly such tests specially would probably take about as long as a program of sawtooth climbs. However, when it is practicable to measure climb per- formance on every test flight, they might be economical. It would no longer be possible to eliminate wind gradients by reversing the air- planes heading between runs, but the method would largely avoid the difficulty of stabiliz- ing sawtooth climbs. . b. Airspeed indicator and clock c. Time trace of dynamic pressure d. Rate-of-climb indicator connec- ted to total pressure e. Air temperature thermometers It is essential to have reliable wind gradi- ent corrections, either from good weather data or from radar measurements of ground speed. Accelerometer: 7:8 INSTRUMENTATION AND ANALYSIS 3 An accelerometer mounted with its axis along the flight path has the considerable merit of giving a direct measure of the excess Te of thrust over drag, since it measures (a) Climbs dv Te + sin y dt w 7:38 Instrumentation and analysis for sawtooth climbs or full climbs require little comment. Good height or pressure measurement with small or predictable lag and hysteresis is required. Radar tracking on a plotting table would be very useful if enough tracking range is available and the method is operationally where y : angle of flight path to horizontal and W = gross weight. 7:11 have been recorded using an airspeed indica- tor and a clock, usually on a photo-panel. This method has been in routine use in England for about five years. It requires relatively little analytical effort and does not require conversion of a rate of change of IAS reading to a rate of change of true speed, or a correction for any inclination of the flight path resulting from changes in altimeter position error with speed. The main difficulty is that it is nec- essary to measure very low accelerations and provide high sensitivity. At a true forward a speed of 400 knots, an acceleration of 0.01 g is equivalent to an unaccelerated rate-of- climb of 400 ft. per min. Measurement of the acceleration to within around 0.002 g, there- fore, seems desirable. Its main merits are availability and a reasonable degree of reliability. The instru- ments commonly form part of the standard equipment of the test airplane. The airspeed indicator is a robust and reliable instrument, usually very consistent in behavior. As test runs are made with the pressure at the static source constant, lags can only arise from pressure lag in the pitot lines (which will be negligible inasmuch as the volumes of the instruments are small) and instrument lag, which should not be serious in the present- case, with so simple an instrument. It may be noted that it is not essential to measure the inclination of the accelerometer axis to the flight path, as the effect of this may be determined from the accelerometer readings in steady level flight at the relevant Mach numbers and weight/pressure ratio. . The practicability of this method depends on the availability of a suitable accelerometer and the practicability of determining, with sufficient precision, the inclination of the accelerometer axis to the flight path, Pre- sently the method is unproved, but there are grounds for hope of its success. Differentiation of the record of indicated speed against time, to give the excess thrust power, stretches the instrument to the limit of its sensitivity and readability, but with suitable techniques of analysis, satisfactory results are obtained. Any errors give visible scatter and are not hidden as accelerometer errors might be. The procedure for analyzing accelero- meter records would be simple. It would be: (1) Correct for instrument error. The disadvantages arise from the fact mentioned above, that the airspeed indicator is barely sensitive enough and that much film reading and computation are required. In general, altimeter position error varies with airspeed, so that an acceleration run made at constant indicated height will be made along an inclined path. If the technique of analysis does not take account of this, errors will be introduced in the absolute values and some- times in the slopes of the curves. (2) Correct for inclination B of ac- celerometer axis to flight path, by subtract- ing g sin B from the reading (B positive if the accelerometer axis is pitched up rela- tive to the direction of flight). (3) Multiply the reading so corrected by 101.3 VT (Vt being the true airspeed in knots, accelerometer reading being in ft. per sec.? ) to give the corresponding unaccelera- ted rate-of-climb Ro in ft. per min. The effect should be allowed for, if the position error change is appreciable and if accurate absolute values of the rate-climb are required. If the tests are only to give the optimum technique, the effect is less im- portant, Airspeed Indicator and Clock: Two kinds of procedure are possible: The great majority of level acceleration runs which have been made up to the present (a) Compute and plot total energy against time, then take the slope as the rate of change 7:12 of energy Tos ICAO or NACA ambient air tem- perature K 2.5 (b) Take increments of indicated airspeed over suitable internals of time At and com- pute the corresponding changes AE in total energy, taking AE/At as the rate of change of energy. 1 +0.2 Bulan 285.2 AI - gos 1 + 0.2M? 7:40 Experience suggests that procedure (b) is preferable. In particular, it ensures that any scatter in measurement and analysis is present in the final plot and so enables an intelligent appreciation of the agreement be- tween repeat tests. Az 1 + { (1 + 0.2M2)3.5 -1 (1 +0.2 M2)2.5 7:41 It has been found desirable to have a speed change of at least 10 knots in calibrated airspeed, and a time change of at least 0.2 minutes in each interval. It is, of course, possible to overlap the intervals, if the total speed range is small. About twenty such intervals should be taken. where M = Mach Number a = speed of sound at 15° C, knots os • standard relative density at test pressure altitude. It is desirable to use automatic computa- tion if it is available; however, if computing is by hand, the unaccelerated rate-of-climb Ro may be deduced from the change in cali- brated speed Vc and pressure altitude Hp by the relation (assuming an instrument cali- brated to read correctly at sea level). An analysis routine is outlined in Appendix II. A, V AVC Tat A time trace of airspeed and height would materially reduce the labor of reading the record, in that it would be possible to deter- mine the rates of change of Vc and Hp much more easily. Ro 100 Δ: Tos + A2A HP 7:39 where Ro - unaccelerated rate-of-climb, feet It may be noted that the two-turn type of airspeed indicator is likely to be better than the multi-turn type of linearized scale, in that it should have less friction and inertia. 8 per minute At : time interval, minutes Vc : calibrated airspeed, knots Time Trace of Dynamic Pressure: AVC = increase in Vc in interval At : AHp: increase in pressure height in interval At A time trace of dynamic pressure (given, for example, by a transducer recording on an oscillograph) is potentially a better and more convenient form of record than photographs of an airspeed indicator in that the slope of the trace is roughly proportional to the de- sired quantity. Tay - test ambient air temperature : °K 7:13 dh Ro : -Ri +12 +0.2MP) : We have, assuming a consistent system of units, 1 d (Pt-Pol g - ) RTA RO 8 (A2-1) Py-Podt R; + 2.1 approximately RTO 1 dia 7:43 -A2 O 9 Po dt 7:42 where A2 is defined by Eq.7:41 and as dh/dt is small, If, on the other hand, the air entering the instrument were at ambient temperature, we would have Po = ambient air pressure Ro : -(1 + 0.2 M?) Ri +(2+0.2M?) 2+0.2m2 (car) dt Pr: total air pressure : 7:44 R : gas constant (= 3090 ft /sec? /ºK in ft-sec units) O Neither of the above assumptions will hold exactly. Only tests can show which is more nearly true. To: ambient absolute air tempera- ture t : time It will be convenient when applying Eq.7:42 to have position error corrections to the in- dicated pressures plotted against indicated dynamic pressure. It is worth noting that when the optimum Mach number for climb is high, it will often be sharply defined by a rapid drag rise. Use of Eq.7:43 would then give a good indication of the optimum speed. We may also note that with many installations the pressure error at the static pressure source will vary closely as the square of the airspeed, making dh/dt proportional to the rate of change of the square of the speed. The term in dh/dt in the above equation could then be omitted with- out altering the position of the peak of the curve. Given suitable instrumentation, this would seem to be a promising method of recording level accelerations, particularly if the trace displacement is linear with regard to the dynamic pressure. Determination of d(Pt-Pa) /dt is then fairly simple, Rate-of-Climb Indicator Connected to Total Pressure: We may summarize by saying that a rate- of-climb indicator connected to total pressure is promising as a means of making a very quick but rather rough assessment of the optimum speed, but that it is as yet unproved. A rate-of-climb indicator connected to total pressure will give a reading which will approximate, after a change in sign, the de- sired quantity Ro. This makes it very attrac- tive as a method of making a quick but fairly rough check of the optimum speed, Air Temperature Thermometer: Let us assume that the instrument reads correctly when used in the normal way. Then if the temperature of the air entering the in- strument were equal to the stagnation temper- ature of the ambient air, we would have the following relation for an instrument reading Ri: An idea which is attractive at first sight is to record temperatures from two thermo - meters with widely different recovery fac- tors, because the temperature difference is proportional to the square of the true air- speed. This is a time differential that would lead fairly directly to the required data. Further consideration suggests, however, 7:14 that such an approach may be impractical. We would have 334 dh RO- OST, -Tzlt K, -k2 dt к dt 7:45 where K1, K2 are the recovery factors rate-of-climb R, is straight forward, using established methods. If the aim is to pre- sent curves of Ro as part of the published test results, this must be done. If, on the other hand, the tests are only to produce data from which to determine a climb tech- nique, it may not be considered worthwhile to go through the whole standardization pro- cedure when the test air temperatures do not differ too much from standard (section 7:6). However, even then, it is probably worthwhile to plot test rate-of-climb mul- tiplied by the ratio of standard to test weight; this makes an almost complete weight cor- rection and so will bring repeat runs made on the same day (and hence at the same air temperature) together. T1, T2 are the indicated tempera- tures and Ro is unaccelerated rate-of- climb, ft. per min. . The factor 334/(K,-K2) is unlikely to be much less than 500. Therefore, to deter- mine Roto within, say, 100 fr. per min, one must determine d(T-T2)/dt to within 0.2° per minute. This would be difficult. It would seem that relative to the dynamic pres- sure trace technique the temperature method seems too unpromising to be worth develop- ing, at least for subsonic and transonic air- planes. It would, of course, be necessary to determine (K1-K2) from flight test. For airplanes with simple jet engines, an alternative approach is to make the tests at the desired value of NWT, and, possibly, of the ratio of weight to air pressure. Correc- tion for temperature would then consist only of multiplying the test unaccelerated rate-of- climb by Tor/Tas, where Toy and Tas are respectively 'the test and standard ambient air temperatures. 7:9 STANDARDIZATION OF RATE-OF-CLIMB CURVES If tests were at the required ratio of weight to air pressure, no weight correction would be required; otherwise, the usual weight cor- rection procedure would be used. Standardization of the test unaccelerated 7:10 REFERENCES 1. Lush, K, J., "A Review of the Problem of Choosing a Climb Technique, with Proposals for a New Climb Technique for High Performance Aircraft," British Aeronautical Re- search Council Report R & M 2557. 2. Lush, K, J., "The Loss in Performance, Relative to the Optimum, Arising from the Use of a Practical Climb Technique,” To be published as British Aeronautical Research Council Report R & M 2756. 7:15 APPENDIX I FLIGHT TEST TECHNIQUE FOR ACCELERATED LEVELS FLIGHT SCHEDULE to the intended values. If photopanel record- ing is used, records should be taken about every five (5) seconds. 1. At least two and preferably four runs should be made at each of three or four heights, the heights being: a. Fairly near ground level b. At a medium altitude 2. At low or medium altitudes, start each run at about the minimum drag speed (say 180 knots) at or below the desired altitude, trim suitably, bring the engine to the required conditions, level off at desired altitude and let the airplane accelerate well past the expected optimum speed. Start recording at any convenient time before settling down to the acceleration. C. At a height at which the maximum rate-of-climb is expected to be about 1000 ft. per min. d. At the tropopause, if this is much below (c). Tests at the two extreme heights at least should include repeats made on dif- ferent flights. 2. If a number of configurations are to be checked, it may be sufficient to run tests at two heights only for some of them. Also, if the climb schedule is expected to become one of constant Mach number at high alti- tude, fewer tests may suffice. 3. It is important to have the airplane under smooth control before the important range of speed is reached. To give time to set up the test conditions, it is often helpful to extend the dive brakes during this pre- paratory phase or alternatively, as indicated above, to start below the intended test alti- tude and set up the conditions while climbing. The trim setting should be so chosen that retrimming will not be necessary until late in the run. TECHNIQUE 1. Each run will consist of an accelera- tion through a range of airspeed covering the expected best speed at constant altimeter reading, and engine speed (or other critical engine parameter). Small changes in alti- meter reading will not matter, but it is de- sirable to hold the engine conditions closely 4. Runs made near the ceiling will differ from the above in two ways. First, if the initial speed is too low, the airplane may decelerate instead of accelerating. If this occurs, try again using a higher initial speed. Second, acceleration is usually slow in terms of indicated speed and the pilot has ample time to make adjustments. 7:16 APPENDIX II ANALYSIS OF AIRSPEED INDICATOR TIME HISTORY FROM LEVEL ACCELERATION altitude computed for the run as a whole. This should approximate the mean altitude, * 1. A procedure is outlined below by which to analyze an airspeed indicator time history. This type of recording is presently the only one in established routine use; other methods of recording promise to be more convenient, but it would be somewhat premature to set up procedures at this stage inasmuch as too many details will vary between particular cases, For similar reasons a procedure is given for hand computing only, since machine techniques vary too much with the type of computer available. Choose a speed interval of at least 10 knots, big enough to cover a time interval of at least 15 seconds (very large time inter- vals may be needed near the ceiling). Pick out from the record some twenty intervals of about the above length. These may be consecutive or overlapping if necessary. 2. The basic equation (Eq. 7:39) is At the beginning and end of each interval read airspeed, altitude (if required), engine speed or equivalent. Read air temperatures with one set of readings around the middle of the run. A VCAVC Ro #azono + 100 PREPARATION where Ai is a function of calibrated speed and Mach number and A2 a function of Mach number only and Correct the above readings for instrument error and, where relevant, position error, and so deduce, among other things, calibrated airspeed Vc , calibrated altitude (that is, true pressure altitude) Hp. Vc : calibrated airspeed AVC : increment in Vc in time At AHP : increment in pressure height in time At. At the speed at which the temperature was read, evaluate: 1. Mach number, from Vc and Hp, PRELIMINARY DATA 2. Ambient temperature, from in- dicated temperature and Mach number, and Preliminary data required are position error corrections to indicated airspeed and altitude, the air thermometer recovery fac- tor and, of course, instrument calibration. 3. Test weight. ANALYSIS TEST DATA If the height range covered during the run is less than about 200 feet, the steps involv- ing AHp may be omitted and one calibrated Denote values of parameters at the begin- ning and end of any interval by subscripts 1 and 2 respectively, the mean values by subscripts m, and differences by A. 7:17 For each interval evaluate (10) AI Vcm AVC 100 (1) AVc = AVc = VC2 Vc2 - VCI (11) A20Hp (2) ΔΗρ: Ηρ2 (2) Hp2- HAI (12) AL VCM AVC () + A2 AHp 100 (3) At 8 12-1 (4) Vom - ¿ (Vc, + Vcz) (13) Al Vem AVC 100 { ལ་ + A2 Ho}() 6+ - ) (5) Hpm (Hp, + Hp2) { (6) Os (14) Ro (13) Standard relative density at height Hpm () To Tos (7) Mach numbers and Vts, the true speed which would give the test Mach number on a standard day, from Vcm and H Pm Check that the engine speed (or other en- gine operating parameter) was held substan- tially constant in the region of peak rate of climb (say, within 14% in the case of engine speed). If not, evaluate the mean engine speed į (N, + N2) or equivalent for each in- terval and plot it against Vts. This assists subsequent evaluation. (8) Aſ from Mach number and Vom and os (9) A2 from Mach number € 7:18 PART II ANALYSIS OF THE OPTIMUM ENERGY CLIMB SCHEDULE 7:11 INTRODUCTION 7:12 DEFINITIONS Total energy is defined as the sum of the kinetic and potential energies and is given by For a given airplane with a given power- plant, the climb schedule may be expressed in the form V = f(h) which relates airspeed : to altitude. It is apparent that similar equa- tions can be written in terms of the cali- brated airspeed, Vcal, or the Mach number M. WV 2 E = Total energy - Wh + 29 7:46 where h is the geometric or “tapeline” alti- tude and V the true airspeed of the airplane. The origins for h and V are arbitrarily taken as h = 0 and V : 0. It is well known that for any type of air- plane an increase in flight path speed during a climb decreases the instantaneous rate of climb and also that the converse is true, In most cases the flight path speeds for piston-engined airplanes are small compared to those achieved by jet airplanes, and there- fore, the effects of flight path speed are more important in the case of jet aircraft than for piston-engine planes. For propeller-driven aircraft, the deter- mination of the best climb schedule is simply accomplished by using the "sawtooth” climb technique to determine the speed for maxi- mum climb rate. Energy height (total height) is designated by he and is defined as the total energy per unit weight. Thus v2 he-h + slo 7:47 which is seen to be independent of the air- craft mass. Por jet-propelled airplanes, it is known from analysis and experiment that the opti- mum climb angles are less than those cor- responding to the maximum rate of climb at a given altitude, and thus represent greater flight path speeds than the speeds for so- called maximum rates of climb. We next define w as the rate of change of energy height, thus dhe W dt 7:48 ) Under these conditions, the time required to go from one altitude to another depends on the variation of flight path speed between the two altitudes, a factor which has gen- erally been neglected in the climb analysis of propeller-driven airplanes. (Note that w is the instantaneous rate of climb when the flight path speed is main- tained constant.) Explaining this thinking in terms of energy, we see that when the flight velocity is large, variations of kinetic energy (mV2/2) can no longer be ignored in comparison to the po- tential energy due to height. At speeds in the neighborhood of the op- timum climb speed, potential energy may be transformed to kinetic energy, and vice- versa, in a reversible fashion, i.e., without appreciable loss in the conversion. 7:19 7:13 THEORETICAL ANALYSIS OF THE OPTIMUM CLIMB SCHEDULE This is not a definite integral, since we have not as yet provided a specific relation between w and he. We shall assume that W = f(h,V). (This assumption is discussed in section 7:16.) We shall begin by making certain sim- plifying assumptions with the understanding that these assumptions must be re-examined later to correlate the analysis with actual operating conditions. (See sections 7:16 and 7:17.) Thus, we have w :flh,V) 7:50 and y he : h + 29 Our assumptions are as follows: (1) The control of the engine is not at the disposal of the pilot because the throttle is either fixed or must be operated by the pilot in some scheduled manner depending on speed and altitude. For example, a jet engine must be regulated to maintain a given constant RPM or to hold a give tailpipe temperature. 7:51 or, on eliminating h 1 v (he,v). W 7:52 (2) The climbs occur in a standard at- mosphere. Then nez '1,2 s elhe, vidhe . (3) Atmospheric turbulence is negligible in the case of measurement with respect to the ground as reference, or any atmospheric motions are in the form of a constant hori- zontal wind, which is a function only of alti- tude, in the case of measurements with re- spect to the air mass. nei 7:53 This integral will be minimized if V is a function of he such that (4) The climb is made in a single vertical plane. Oy : 0 av 7:54 According to these assumptions, the sin- gle parameter determined by the pilot during the climb is the extent to which he can pre- cisely control his plane to follow the pre- scribed climb schedule, V = f(h) while av- . eraging the load factor, a topic to which we will return later. (See section 7:16.) provided the function satisfies certain conditions of regularity apparent in the geo- metric representation of section 7:14. Eliminating he from Eqs. 7:51 and 7:52, we obtain an equation of the form V = $(h) Consider the two energy heights, he, and hez with hez > hej i since w: dhe /dt, dt = dhe/w, the time required to go from he, to he becomes 7:55 ne2 dhe *1,2 - s W which is the condition under which the in- tegral (Eq. 7:53) will be minimized. Accord- ingly, the values of V between the limits her and hez are fixed for all intermediate values of he. he 7:49 7:20 (2) By means of accelerated level flight runs during which the height h is held con- stant, in which case d v2 V dV W = The climb schedule of Eq. 7:55 is the op- timum schedule, i.e., it gives the minimum time required to go from one combination of speed and altitude of the schedule to another combination also satisfying Eq. 7:55. How- ever, the problem of passage from one com- bination of speed and altitude not satisfying Eq. 7:55 to any other combination in minimum time is not resolved. We shall see in sec- tion 7:15 how this latter problem can be solved practically. dt \2 g gdt These two methods serve to provide a good definition of the surface (E), which for turbojet aircraft has the general form illustrated in Fig. 7:1. 7:14 GEOMETRIC REPRESENTATION Eq. 7:52 can be written, in the light of Eq. 7:51, From the foregoing, the optimum energy climb schedule has its projection in the (h, V2/2g) plane defined by the Eq. 7:55. The optimum climb schedule can also be repre- sented on the surface (£) by Eq. 7:54 or the equivalent equation obtained by rewriting Eq. 7:56 as w : $ in, he-h). W on) V 2 , 29 - 7:56 The condition for minimization of the in- tegral in Eq. 7:49 becomes This functional relationship can be de- termined practically from flight tests during which the assumptions made in the preceding section are valid. Actually, it is sufficient to measure V and h at each instant to obtain don{tin, ne – n}= 0 ch 2 or h he :ht sla 29 od dlhe-h) dø dh + дь dh : 0. and then dlhe-h) dh dhe W dt But at V = constant ane : dh on the curve representing he as a function of time. or 0$ дф dlhe-n) O, ah Thus we obtain points on the surface (3) represented by Eq. 7:56. In fact, we see now that there are two methods which can be used to obtain the surface (2): and finally дw 318 on 2 (1) By means of continued climbs follow- ing schedules in the neighborhood of the op- timum schedule. olemas 29 7:57 7:21 4 W Po 6 M P. slo hezo H=0 h Z 2 1 : cte 29 21+ H, = 2,7 zig V. constant 29 he, hi + Hq = cte : constant nez Fig. 7:1 This Eq. 7:57*, which defines the optimum climb schedule (C) on the surface (E) of Eq. 7:56, indicates that at each point of C, the plane tangent to (E) is parallel to the direction established by the intersection of the planes w : 0 and V2 * constant. h + 2g * By interchanging variables, it can easily be verified that Eq. 7:57 is identical to Eq. 7:54. 7:22 T Geometrically this result is simply in- terpreted and it follows that the integral by the points of maximum w along intersec- tions between the surface () and planes per- pendicular to w : 0 erected through lines of constant he ne2 dhe 11,2 8 W hey 7:15 PROPERTIES OF THE GRAPH will be a minimum if for each value of he, 1/w is a minimum, i.e., w is a maximum. (1) The projection (c) of the curve (C) on the horizontal plane (w : 0) gives the opti- mum climb schedule relation V = f(h) 7:55 On projecting C onto the bisecting plane h : V2/2g (perpendicular to the direction of the line defined by w: 0, he = constant), ) we obtain the following graph. which is the practical definition as far as the pilot is concerned. . From Fig. 7:2 it may be seen that the area A between he, and he will be a maximum when the curve w : f(he) is the projection of the outline of (2) viewed parallel to the di- rection of he constant lines in the plane w = 0. (2) Examination of Fig. 7:1 reveals that if P is the maximum of the curve w f(V) obtained with he = constant, and if M is the maximum of the curve w = f(V) obtained for h : .constant, then the point P corresponds to a greater speed than that of M. : The curve (C) representing the optimum energy climb schedule is, therefore, given (3) In the neighborhood of the optimum climb schedule w does not vary appreciably from its maximum value along the he • con- stant intersection. From this it follows that small variations in speed from the optimum schedule V = f(h), do not materially change : the value of the integral representing the time for ascent from he, to hez (i.e., the area A of Fig. 7:2 is not appreciably altered by small changes in V from the optimum schedule V = f(h)). h Therefore, at a given height, provided that the speed does not vary too much from the speed for best energy climb, a variation of V is equivalent to a variation of h, each pro- ducing the same variation in energy height. That is to say, the kinetic and potential en- ergies may be interchanged in a reversible fashion provided the climb schedule does not deviate too much from the optimum.* On any other portion of the surface (E), it is apparent that this reversibility does not exist. H H H₂ por * This fact is also expressed analytically by Eq. 7:57. Fig. 7:2 7:23 (1) Atmospheric Conditions This condition of equivalence of energy transformation permits us to deduce the nature of the surface (E). (4) Figs. 7:1 and 7:2 allow us to study schedules other than those corresponding to the optimum. If we depart from a point on (2) not situated along C, and which repre- sents an energy height he, it is evidently impossible at first to follow a path along the surface with he, a constant, since in this a case We have assumed that the only atmospheric disturbance present has been a horizontal wind of constant velocity at any one altitude. In the event that measurable vertical move. ments exist, it is evident that the measure- ment of w will be in error. It should be possible to correct our test data for vertical motions. However, the complications in- volved cause rejection of this idea. In the case where the horizontal wind grad- ient with altitude is not zero, the airplane will gain or lose energy in the climb. If the horizontal wind gradient is known at all al- titudes, we can compute the acceleration due to the wind dhe W : : 0. dt Rather, one should attempt to maximize the area of Fig. 7:2 by approaching as closely as possible to the intersection of the he. con- stant plane and the curve (C) without increas- ing the load factor sufficiently to risk a de- crease in w. () dt wind at any point of the climb and correct w ac- cording to the relation V dV Aw: g dt wind. ) Therefore, it will generally be advan- 'tageous: (a) If one is at a speed less than that of constant he on (C), to speed up to the point (C). (b) If the airplane takes off from the ground to gain speed close to the ground. (c) If one wishes to obtain an altitude h2 and a speed V2 greater than the optimum, to first climb to an altitude greater than h2 without exceeding hey and then to descend slightly allowing the speed to increase to V2. To insure that all data points form a sin- gle surface (2), it is necessary to correct data obtained under various observed con- ditions to standard atmospheric conditions. The correction is considered in section 7:17. (2) Operation of the Airplane In the maneuvers described above, it should be realized that because of load factor variation, there exists the risk of departing from the region wherein our assumptions are valid. We have already noted that the engine must be operated in accord with certain re- strictions, which must be well defined for each set of flight conditions. It is evident necessary and sufficient condition (from the point of view of airplane operation) for the existence of a well-defined surface (E) is that the engine limits be specified for all values of V and h.* In section 7:17 we 7:16 DISCUSSION OF ASSUMPTIONS To simplify the discussion, we have made certain assumptions which we must inves- tigate further to assure the validity of our results. * Since the atmospheric conditions are de- fined in Part I of this section, the engine op- erating restrictions in effect call for a spec- ified variation of thrust with speed and alti- tude, 7:24 shall examine the practicability of achieving this in the case of a turbojet airplane, The most important operating restriction on the airplane involves the value of load factor during the climb.* If we call F, the resultant of the forces along the flight path (thrust-drag), we have In the event that drag varies little with n (generally the case at low altitudes at the optimum climbing speed for jet airplanes), Fy is insensitive to n and no corrections need be made for variable load factor. In cases where large variations of drag are encountered, F, is a function of n and cor- rections may be required if level flight runs are used to determine the climb schedule. However, we shall not concern ourselves with this topic here. dv m dt FT - (mg sin y) where mg : airplane weight y : flight path inclination to the horizontal. With Vsin y = dh/dt, we have, on multi- : plying the preceding equation by V, To determine the influence of non w (which is required for a rigorous definition of the surface (£)) n must be defined for each value of h and V. In practice, slight variations of energy level are acceptable without producing noticeable thickening of the surface (E) es- pecially at low altitudes. This permits ac- ceptance as points on the surface, data ob- tained in different fashions with different flight path slopes or curvatures, provided the variations in n are small. . dh d FTV 1 dt dt la mg or FTV. W mg In general, the method of determining the optimum energy climb schedule by continued climbs (see section 7:18) will be more pre- cise than the method of using accelerated level flight runs, since the load factors at- tained in the former case are quite close to those of the actual climb; for instance, the normal load factor in a steep climb differs appreciably from the n : 1 of level flight. Fy is the instantaneous excess thrust over that required for level flight at the same load factor, and w is, therefore, the instan- taneous excess power per pound weight, If we now consider a variation of load fac- tor between two flights carried out under otherwise identical conditions (same weight, same h and same V), w will not vary from one case to the other if it remains constant. The difference in angle of attack in the two cases generally produces negligible differ- ences in thrust, but possibly noticeable dif- ference in drag especially at high altitudes. Finally, the weight variations due to burn- ing of fuel can generally be accounted for as a given function of altitude in a continued climb, which provides another advantage of the continued climb method. ) (3) Analytic Conditions * We shall not consider the variation of air- plane mass during the climb due to burning of fuel, this variable being readily handled by determining the relation m = f(h). ) When the preceding conditions are ful- filled, the surface (E) is well defined, but certain theoretical paths on the surface are not possible without changing the load factor in which case the actual path is no longer on the surface. 7:25 Similarly, certain paths are analytically impossible, since w - dh /dt, and, there- fore, if w is positive, one cannot move ex- cept to increase he (i.e., decreases in he are not permitted). Here we shall consider only the case of a turbojet with a single variable parameter (assumption 1, section 3). . (1) Correction for Operating Conditions If we write If during an actual climb the thrust of a turbojet (determined as a function of RPM, N) deviates from standard, a correction may be applied. In effect, we have shown that dhe dhe dh dhe 24) dt an dt V dt д 20 W FTV mg . If we now increase the thrust T by an amount AT, the excess thrust Ft and, there- fore, w will be increased proportionally. Thus we see that at a point on (2) where the par- tial derivatives dhe/ah and one/a(V2/2g) are defined at the same time as w, there remains only one variable parameter (for example, ascent speed or flight path slope). Further, this parameter defines the tangents to the curve representing the displacement on the surface, or, since the maximum slope of the flight path is vertical, one has the theoretical extreme case corresponding to the values of dh/dt and d(V2/2g)/dt in vertical flight, the load factor corresponding to the point under consideration. Δw V AT . mg AT - It is sufficient then to know the variation of thrust AT due to a correction AN on N to obtain Aw.* It follows that the analytic conditions re- quire not only a positive variation of he for w > 0, but also an angle of less than 180° which corresponds to the slopes of the de- fined surface. Without going into further detail, we note that the possible paths on the surface (E) are generally still further lim- ited by load factor limitations. To accomplish the correction, one gen- erally refers to the engine manufacturers' charts presenting thrust as a function of RPM, speed and altitude. The accuracy ob- tained (using such charts) is often sufficient for correction procedures. However, it is always possible to make precise corrections provided the necessary engine parameters are measured. This method is used when certain anomalies are noted in the engine operation or when complex engine designs are employed. 7:17 CORRECTIONS (2) Temperature Corrections It is not possible to consider here all the practical cases which require corrections to make them conform to the theoretical cases considered; i.e., where our assumptions are satisfied in a convenient manner. In particu- lar we shall not consider the important case of correcting for load factor. This correc- tion would require a special chapter by it- self treating "reduced coordinates." A variation of ambient temperature T, . * Editor's note: The use of RPM as a thrust correction parameter is not an entirely sat- isfactory procedure. However, it is con- sidered acceptable for small variations from standard. Each type of power plant requires a spe- cial study of the corrections applicable to it, 7:26 between two climbs through a given point (h, V) effects w in two ways. It modifies the engine thrust and also the airplane drag. 7:18 PRACTICAL DETERMINATION OF THE OPTIMUM ENERGY CLIMB SCHEDULE We recognize that the variations of air- plane drag with temperature are small. How- ever, since the drag decreases with the tem- perature, neglecting drag variations provides optimistic data from trials in warm air, in which cases it may be necessary to consider drag effects in the analysis. Let us now assume that we are able either to conduct our flights according to the theo- retical assumptions or are able to make the necessary corrections to standard conditions. Under these circumstances, the experimental determination of the surface (E) is then pos- sible from which the curve C (determined by the intersections of the planes perpendicular to w = 0 and including the lines he = constant) is obtained. The projection (c) of (C) on the (h, V) plane gives the desired optimum energy climb schedule. As far as the thrust is concerned, the effects of temperature can be accounted for by noting that thrust can be expressed as a function of NWT where N = RPM and T - tem- N : perature. Other factors being constant, the same results are obtained if NVT = N'NT', where N may differ from N' and T from T'. Since the engine charts are presented in terms of NWT, the RPM and temperature correc- tions may be made simultaneously. We shall now consider the two methods of determining this climb schedule, which we mentioned earlier in section 7:14. (3) Weight Corrections (1) The Method of Continued Climbs Several (3 or 4) continued climbs to ceil- ing are made during which the true airspeed is linearly increased with altitude. Provided the optimum climb schedule lies within the range of speeds and altitudes tested for, the surface (2) can be deduced from an analysis of the continued climb data, the test data ap- pearing as in Fig. 7:3. If the mass m of the airplane differs from the standard mo, the relation w = FTV/mg shows that the simplest way of determining Wo corresponding to mo will be to conduct tests at mass m under conditions such that Ft remains unchanged; i.e., by making the test at the same h and V while maintaining the same angle of attack by variation of the load factor. The relation between the meas- ured w and wo is then given by wo - w(m/m). . Finally let us recall that the wind can be other than horizontal and constant. In sec- tion 7:16 we have developed the corrections required for the case of a horizontal wind of known vertical gradient (naturally, with the wind oriented along the flight path). 2 ) 3 Corrections for vertical components of known winds would be simple to make, but, since vertical air currents are practically always unknown, and, since their influence is large, the only rule is not to conduct tests under conditions when strong vertical air movements exist. Fig. 7:3 7:27 The test data are corrected to standard conditions and we then obtain (for each con- tinued climb) curves of he as function of time (Fig. 7:4). Values of h and V as functions of time can similarly be presented for each climb. H M3 Hi MI If we now consider constant values of he, we shall have, for example, at hej, inter- sections M1, M2, and M3 (see Fig. 7:4) and the slope of the curves he or t at each inter- section may be at once determined at four different values of V for the case shown. Since w : dh /dt, we may cross plot the slopes as a function of V as illustrated in Fig. 7:5. TEMPS TIME From Fig. 7:5 we may at once obtain the optimum speed at the selected value of he (he), and, therefore, at the corresponding h, which gives us the optimum climb sched- ule V : f(h). The value of w is also known, 80 that considering different values of che, we may determine Wmax = f(he) and from : this the time to climb may be obtained by direct integration. In practice the time to climb should be verified by a climb made according to the optimum schedule. This time will be obtained very accurately, even if the pilot is not very precise, provided the speed remains close to that called for by the optimum schedule. Fig. 7:4 HOH We note in passing that Fig. 7:4 is gen- erally sufficient for the determination of the optimum schedule, for the magnitude of the slopes may be directly observed from this figure without recourse to Fig. 7:5. Max. (2) Acceleration Run Method We shall present here the principle of this method*, which consists of making ac- celerated level flight runs from a speed con- siderably less than the optimum to a speed We shall not consider the difficulties which this type of testing often entails. Fig. 7:5 7:28 approaching Vmax at various different alti- tudes. From the data (since dh - 0), we may calculate d W : love you V dV gdt di 12 . At each altitude, h, at which tests were conducted, we shall have cross sections of the surface (E) averaging certain correc- tions, in particular on the load factor as shown in Fig. 7:6. At each point on each curve, h and V are known and, consequently, he is defined as a function of h and V. Therefore, the data can be cross plotted in the form w = f(he) as shown in Fig. 7:7. In this system (Fig. 7:7) the curves rep- resent the projections of section of (a) on the planes h = constant, and, therefore, the optimum climb schedule is defined by the envelope of the curve (see Fig. 7:7). Another more precise method of determin- ing the optimum schedule (C), utilizing the information of Fig. 7:6, consists of con- structing the cross sections w = constant by cross plotting the data of Fig. 7:6 as shown in Fig. 7:8. 2, 22 23 Ni na ng Fig. 7:6 nz (C) NE Hecto ho constant 22 ܕܐܢ ha v2 29 Fig. 7:8 Thus we obtain the intersections of planes of constant w with the surface (E) projected on the (h, V2/2g) plane of Fig. 7:1. The Fig. 7:7 7:29 slope of the line he : constant in this plane is given by the normal to the bisector of the axis system (i.e., by a line with a negative 45° slope) and the tangent parallel to this direction determines points on the optimum curve C. nil. Moreover, the proper load factor and weight conditions are very well simulated during the test ascents. Also the required data measurement is reduced to a minimum, records of speed altitude and time generally being sufficient (if one used meteorological data describing the actual variation of tem- perature with altitude). If we plot the data in the (h, V) plane rather than the (h, V2/2g) plane , the curves of constant he will be parabolas displaced from one another by translation along the h axis, and in this case the points of tangency will also be well defined thus establishing the curve C. The method of accelerated level flight runs is desirable in spite of piloting difficulties frequently encountered, if one wishes to de- termine the optimum energy climb schedule very quickly without resorting to the climbs required by the other method. Temperature corrections may then be ignored, for a first approximation to the optimum speed at the altitudes of the level flight runs (but obviously not for the ascent time). For these latter cases, we note that the curves of Fig. 7:7 need not be obtained from level flight runs. Any other family of curves obtained holding some quantity constant, for example constant load factor, could be pre- sented in terms of w and h with the envelope curve giving the optimum climb schedule. * This thought is very useful for resolving cer- tain difficulties, for example, when it is dif- ficult for the pilot to maintain constant load factor in level flight. Level flight runs serve to provide other data as well; such as, acceleration time (obviously), maneuver boundaries* and max- imum level flight speeds. Finally, certain difficulties encountered in the operation or control of complex engines or those having poor regulation are avoided in level flight runs so that useful data may be obtained from these tests in cases where the con- tinued climbs may produce inaccurate data, (3) Choice Between the two Methods The two methods of testing described above are sufficiently different to justify some indecision on the part of the test en- gineer who must choose the one most adapt- able to the given test program. Ideally both methods should be used, but rarely are suf- ficient time or justification available to per- mit this procedure to be used. We shall, therefore, present some of the advantages and disadvantages of the two methods to pro- vide some basis for a choice. Summarizing, the corrections to be made and the supplementary measurements re- quired are determined by the particular problem to be solved and these factors should be considered when choosing one method in preference to the other. The complications of a given program and the total number of tests required to supply the data (both for the climb schedule and for other purposes) should be carefully considered in selecting which method is best in any given instance, The method of continued climbs is de- sirable when other tests require climbs under similar conditions, since under these con- ditions the loss of flight time is practically * Editor's note: It would be necessary to have h and V defined precisely in terms of the given constant quantity to determine V = f(h) from this scheme. * Editor's note: It can be easily shown that the ability to attain a given load factor with- out loss of altitude is directly related to the acceleration characteristics provided suf- ficient maximum lift coefficient is available to permit attainment of the load factor. 7:30 7:19 CONCLUSIONS assumptions which establish the validity of the theory. This should serve to provide a basis for the further study required in each particular case. We have presented the essential prin- ciples and also the methods of determining the optimum energy climb schedule using the energy height parameter. It is important to recognize on one hand, the theoretical com- plexity of the problems encountered in the utilization of these methods, and on the other, the large number of possible results obtain- able, which give more than just a knowledge of the optimum climb schedule. Each type of power plant introduces new problems because of the way it operates (the rocket motor, for example) and also because of the changes of order of magnitude of per- formance which it is likely to produce. We have considered here only the fundamental When this study has been completed in any particular case, the performance of the plane may then be well defined and the influence of the various parameters which affect its performance clearly established. The man- ner of most effective operation of the air- plane (from the performance viewpoint) can be determined from the tests, as well as the effect of practical departures from the theo- retical case. Finally, using energy methods, we may solve problems hitherto not con- sidered amenable to solution by flight testing methods. 7:31 PART III CORRELATION OF THE ENERGY CLIMB ANALYSES 7:20 REVIEW OF PRECEDING ANALYSES maximize dhe/dt we required that ow ow . On V 2 29 7:59 In both Parts I and II of this chapter it has been demonstrated that the optimum en- ergy climb schedule is determined by the condition that dhe/dt shall be a maximum at all times during the climb so that the air- plane will have stored the maximum pos- sible amount of mechanical energy (kinetic plus potential) at the end of any given time, where dhe 키 ​W : fln, ( dt 29 It is obvious that $ and w are the same quantity so that Eqs. 7:58 and 7:59 are the same relation expressed in different ways. It is easy to verify the truth of this statement for we know that d It has been pointed out that the optimum schedule can be determined by two different flight testing procedures; one involving con- tinued climbs to ceiling and the other, level flight acceleration runs conducted at con- stant altitude. We note that at the onset of the climb, it is not possible to follow the op- timum schedule immediately because of the necessity of accelerating the airplane from rest to optimum speed. For climbs to low maximum altitudes therefore, energy climb techniques are of little value. On the other hand, the use of these techniques applied to high altitude interceptors provides signifi- cant gains in operational performance. It is this fact that explains the great emphasis recently given to the study of the energy height concept. ) dV 29 9 or VOV d : ) 29 g 7:60 so that Eq. 7:59 may be written In Part 1 of this chapter, it was estab- lished that since he is a function of both h and V, we shall maximize dhe/dt when V ow дw . gon av дф V дф 7:61 av 9 on 7:58 which is the same as Eq. 7:58 (since $ : w). where dhe V 2 olo ( Thus we see that identically the same conditions are specified in both Part 1 and Part II for the determination of the optimum energy climb schedule. h + dt 2g/. Similarly in Part II, it was shown that to In Part II an important geometrical in- terpretation is made of Eq. 7:59; namely, 7:32 1 T d²ne d²ne that this equation defines the maximum of w for a given constant value of energy height. Although the analytic development of this fact was not detailed in Part II, it is easy enough to establish the truth of this interpretation. This is done as follows: () a ) level climb d12 di? 7:63 7:21 THE THIRD METHOD OF EXPRES- SING THE CONDITIONS FOR OPTI- MUM ENERGY STORAGE Writing Eq. 7:58 in terms of w, we have Eq. 7:59 defines any line which lies in a plane parallel to a plane perpendicular to w : 0 and and oriented in the direction h : V2/2g, w: 0. However, we also know that this plane is tangent to the surface (E) of Part II. The plane can only be tangent to (E) and still include the line defined by Eq. 7:59 when it passes through the maximum of the curve obtained by cutting the surface (E) along an he = constant line (which is perpendicular to the direction of the line, h : V2/2g). 31 V dw gan av o with dhe dt Therefore, since the maximum of w along a constant he line is given by a horizontal tangent parallel to w : 0 and parallel to the he h+V2/2g = constant lines, it follows that the tangent plane, defined in part by Eq. 7:59, includes the line parallel to w : 0 and parallel to lines of constant he. This establishes the interpretation of Equation 7:59 given in section II. now since dw/OV is obtained at constant altitude, we have д dh eleon de av -181 av d12 / level (2) hieveloped level " dt l level / We shall show in the following paragraphs that there is yet a third way of specifying the optimum energy climb schedule which involves time derivatives rather than deriva- tives of h and V. Although this procedure may be derived directly from physical consid- erations, we shall obtain it using Eq. 7:58 as a starting point since, by so doing, we also demonstrate that this third way of expressing ourselves is merely a restatement of what has already been developed. where a : acceleration. Similarly, ow . Com dhe I climb dh on dt dt This third specification is that the follow- ing equations must be satisfied if the airplane is to fly at the optimum energy storage speed at a given altitude during a continuous climb, Therefore, our original Eq. 7:61 becomes dhe dhe dh dt dhe (2) climb 9 dhe v dt? / level ) dt level dt climb level 7:62 7:64 7:33 which is satisfied when Now, since the excess power can be used just as well, to produce a rate of change of potential energy as to produce a rate of change of kinetic energy, we have that at a given speed and altitude d²hel d²ne dt? I climb dt2/level 7:67 d (Wh) :d dt dt WV2 29 ) (he) level dhe - dt I climb 7:68 where W : weight. Since the weight is the same in both cases and since dw/dt will also be about the same regardless of the method of energy storage, we find that which are the conditions we sought to esta- blish in Eqs. 7:62 and 7:63. dh V dv ole . . 10 dt 9 dt Eq. 7:68 also expresses the reversible interchangeability of kinetic and potential energy at points in the neighborhood of the optimum climb speed. (This was tacitly implied when we set dh/dt = (dhe/dt)climb in the development of Eq. 7:66). 7:65 but in the neighborhood of the optimum climb schedule To amplify this latter statement, we note that since both true airspeed and altitude will be simultaneously changing when the airplane dh dhe . dt dt I climb and dhe dt v 위은 ​dhe II dt g (%。 dt level di 'climb so that Eq. 7:64 becomes RATE OF CHANGE OF ENERGY HEIGHT d²he اههه () dhe dt dt? I climb level a²ne dhe 1. time () d12 level dt climb 7:66 Fig. 7:9 7:34 is following the optimum energy climb sched- ule, the quantity (dhe/dt)climb is a function of both hand V, although in the region of the optimum energy climb speed, (dhe/dt) will be the same regardless of what method is used to change he. ( dhe/dt)climb curve with this latter curve being determined for the optimum energy climb condition. Therefore, if the airplane starts to climb at point (E), we shall have at all times the maximum possible value of ( dhe/dt: w), which maximizes the value of he as a function of time. A physical interpretation of the conditions 7:62 and 7:63 may be obtained by referring to Figure 7:9. First, we note that relations 7:62 and 7:63 locate point (E) as the optimum energy climb point, since only at point (E) are the speci- fied conditions satisfied. The physical rea- sons why this must be so are as follows: We note that this analysis has ignored the topic of load factor variation, which would practically be involved if we were first to accelerate to either point (A) or (E) of Fig.7A:1 and then attempt to follow the given climb schedule. It is apparent, however, that the analysis is nonetheless valid for the determination of point (E) when the airplane is in a continued climb. In this case, Fig. 7A:1 represents a true comparison of the climbing airplane and the equivalent airplane in level flight at some given altitude. Suppose that an airplane accelerates in level flight to point (A) (which determines the speed for so-called maximum rate of climb, conventionally determined in the past by sawtooth climbs) and then starts to climb. Due to the decrease of excess power with altitude and the necessity of accelerating along the flight path (dhe/dt)climb de- creases initially at a greater rate than (dhe/dt level so that at some time ti, the climbing airplane, will be at point (B). Had it remained in level flight, it would have been at point (C). 7:22 PRESENTATION OF THE THIRD SET OF CONDITIONS IN TERMS OF V AND A To complete the energy climb picture, it is desirable to write equations which illustrate the relationship between w as ob- tained in a climb and w as obtained in level flight tests. The energy height being the integral By definition we know that s S'Care ) at hen dt, v2 di he : h + to 29 so that in level flight with h = constant we see that climb out at point A cannot store the maximum amount of energy in a given time, since in this case, he, is the area under the curve DAB which is less than the area DAC representing the air- plane in a level acceleration run. dhe d V dV to W : ) - en dt level dt 29 gdt 7:69 In the general climb, both V and h are variable, so that in the climb It is easy to demonstrate that the (dhe/dt climb curve is concave upward at speeds exceeding those for maximum climb rate, and moreover, that at speeds exceeding the optimum energy climb speed the (dhe / dt )level curve falls below the dhe dh $18 V + g dV dt W dt / level 7:35 Moreover, if we fly in accord with a given schedule, V: f(h), then the preceding = equation becomes from which it follows that Eqs. 7:67 and 7:68 may be written V dV dh dt Go dhe dt. I climb (1 + ý g dt level W : dh V dV dh dh V + 7:72 do 1+ g dh g dh dt dt wdw 7:70 (i+) dw dt level hanno 9 dh dh climb • from which it follows that the actual rate of climb is given by 7:73 dh W dt 1 + V dv g dh These relations may be used to determine the optimum climb schedule using flight data from acceleration runs conducted at various altitudes. However, normally a trial and error procedure similar to that suggested in Part I is required. 7:71 Note that if we make dv/dh = 0, we will then have that the rate of change of potential energy equals the total rate of change of energy. For in this case, there is no accel- eration along the flight path. 7:23 SUMMARY Moreover, if we make dv/dh negative, then we can achieve a rate of climb greater than the rate of change of energy height, since we are converting kinetic to potential energy and under these conditions the air- plane is a "zoom". As noted previously, provided we do not deviate excessively from the optimum schedule, the energy interchange may be made reversibly. In this chapter we have shown that the conditions for maximum energy storage may be presented in several different fashions, each having some advantages from the point of view of visualizing the physical nature of the problem. Using the foregoing, we have In Part II it was shown that if the quantity dhe/dt was considered a function of h and V42g, then a surface representing the rate of change of energy height could be construc- ted which allowed geometric interpretation of the optimum energy climb requirements. Using this surface as a basis of discussion, direct procedures (not involving trial and error) were developed for determining the optimum energy climb schedule using several different flight techniques. (5) dane dw v2 29 at² - d12 / level dt /level d and dw d dhe *** le to climb d+2 I climb dt In both Parts I and II, limiting cases were considered which revealed that although accelerated level flight runs were a satis- factory means of determining the optimum energy climb schedule in many cases, a dt / climb dhe dh dt dh ols () climb dt 7:36 more accurate technique was that of making continued climbs. It is pointed out here that this is particularly true when the climb angles become large (as is the case with certain modern high performance jets, par- ticularly at low altitude). In such case, the equality of lift and weight existing in level flight is a rather poor approximation to the actual conditions existing in a climb. It is apparent that much more work is yet to be done on investigating applications of the energy height concept, both in tactical evaluation of aircraft and in analytic treat- ment of performance. However, it is felt that here we have set down the fundamental thoughts, and further amplification of the topic of energy height does not come within the scope of this manual. . 7:37 APPENDIX A PROCEDURES FOR DETERMINING THE ENERGY CLIMB PERFORMANCE OF TURBOJET AIRCRAFT Editor's Note Because of a revision in the original concept of the AGARD Flight Test Manual, it has been necessary to reduce the size of Mr. J. F. Renaudie's excellent paper, “Procedures for Determin- ing the Energy Climb Performance of Turbojet Aircraft." In reducing the size, every effort has been made to maintain the es. sential features of the original work. D. O. DOMMASCH 7A:1 INTRODUCTION to conduct flight tests at constant he to provide curves directly such as shown in Fig. 7A:2, we find that alternate schemes for obtaining the optimum energy climb schedule are employed. There are two basic procedures, which are: In Chapter 7 it was established that a surface (E) may be constructed represen- ting the overall climb performance of a given airplane. In that chapter, the surface (E) was presented in terms of a coordinate axis system consisting of the orthogonal rectangular coordinates, w: dhe/dt, V2/2g and h. (1) Level flight acceleration runs (2) Continued climbs It is apparent that this surface can also be presented in terms of the set of coordi- nates, w, he and V2/2g as shown in Figure 7A:1. As discussed in Chapter 7, there are advantages and disadvantages to both methods and the selection of one depends on the test circumstances. For any case, it is necessary to compute the value of w: dhe/dt and, in the case of continued climbs, the true tape- line altitude. These items are considered below. The cross section of (E) at a given value of energy height appears as in Figure 7A:2. 7A: COMPUTATION OF TAPELINE OR GEOMETRIC ALTITUDE AND EN- ERGY HEIGHT From this figure we see that the speed corresponding to maximum rate of change of energy height is at once determined by constructing a horizontal tangent to the cross section curve. Since it is not convenient To determine the true altitude we start 7:38 W w =h-h 29 (8) constant Section along he=constant plane h: constant Villa ILA V for optimum energy climb Fig. 7A:1 Fig. 7A:2 with the well-known equations dh T dhp Tst dp :-podh 7A:5 7A:1 P: PRT 7A:2 Either of Eqs. 7A:3 or 7A:5 may be in- tegrated between the limits (1) and (2) to give (for average values of temperature and pressure) where in the English engineering system R : 1718 ft.? /sec? °R. RT Δη : Ap Combining these, we find that 9 P 7A:6 dp = g p RT dh. or 7A:3 т T An : Anp. T st In the standard atmosphere 9 р RTst dp dhp 7A:4 The average temperature is (T, + T2)/2 and the average pressure (Pi + P2)/2 and Ap : P2 - P. Thus Eq. 7A:6 becomes when Tst • standard temperature at hp R an : hp filho Hom) - pressure altitude. Tat 2 + T + PI (P2 - DD 9 P2 Dividing Eq. 7A:3 by 7A:4 7A:7 7:39 and R ή : ΣΔ ) T2+ TI R Σ g El Under subsonic conditions the Mach num- ber may be determined from readings of im- pact pressure and static pressure (suitably corrected for instrument, position and lag error) using the isentropic relation P2 + P (P2-PI). Y/Y-1) Y-1 Di Ds = (1+1=1 me) *Ar=1 부 ​M2 7A:11 The pressure and temperature may be measured at intervals of several thousand feet of pressure altitude during the climb (or these data may be determined from radio- sonde transmissions) and Ah computed for each interval. The sum of the interval heights gives the true altitude corresponding to each pressure altitude. where y = 1.4. 7A:3 DETERMINATION OF w: dhe/dt Graphical integration may of course also be used to determine the altitude, since 2 T ከ : f dhp Tst The value of w for any value of he is ob- tained as the slope of the curve of he vs. t. Although the slope of this curve is often obtained graphically, greater precision may sometimes be obtained using analytic pro- cedures. The finite difference interpolation formulae of Stirling or Bessel may be dif- ferentiated to provide equations for obtaining derivatives of experimental data. Consider Fig. 7A:3, which is a plot of energy height vs. time for a given test run. 7A:8 and the area under the curve of T/Tgqvs. hp gives the true tapeline altitude. The energy height being defined as V2 ne : energy height = h + 29 We first break the time interval into con- venient equal increments At. At some time In the energy height, has a value hen. Simi- larly at tn+i, we have hent, etc. The dif- ferences hen - henti-hen are now computed for each in and from these the coefficients 7A:9 - is simply obtained by adding the tapeline alti- tude to the value of V?/2g existing at a given test point. An . Ahe, Ahen- Bn + Ahen-2 Ahen tl Since we must know the true speed, we make use of the equation given in Chapter 1, V = Mc where V : true speed, M : Mach number, C: sonic velocity. The sonic velocity is given by are determined. The value of wn : by the relation dhen/dt is then given c: KVT т 7A:10 where K is a constant depending on the units used for c and T. 7An -Bn : wo 12 AT 7A:12 7:40 During test, dE/dt is the excess power given by he Ww' : (F'-D') whereas, under standard conditions henta menti WW : (F-D) V nen subtracting neni hen-al Wow' -w) = [F'-F-D'-D] v or Wdw : VIDF-dD) 7A:13 where dF is the variation of thrust from standard and DD is the variation of drag from standard. +1 +2 Fig. 7A:3 We now consider the problem of reducing the data to standard atmospheric conditions using a constant weight, constant speed, con- stant pressure altitude procedure. In cor- rection the following are held constant 7A:4 CORRECTION PROCEDURES pressure altitude h In general, flight data must be corrected for the following variations from assumed standard conditions static pressure P true speed V (a) Thrust variations (b) Non-standard atmospheric conditions weight W The table on the following page gives a listing of the relation between test and cor- rected conditions. (c) Thermal air current effects (d) Gradient wind effects Under standard conditions the energy height is he : hp + ala 29 7A:14 In the following work a prime superscript designates uncorrected data and the symbol by itself represents corrected data, We shall consider first the effects of variation of thrust and drag brought about by a "non- standard engine' and by atmospheric varia- tions. ) and during test V2 he : h' + šla By definition 29 7A:15 de wdhe so that : W w di do where the notation is as in Chapter 7. ne che : n'-ho. 7A:16 7:41 The corrected value of w may be com- puted using relation 7A:13 The test Mach number may be obtained by standard procedures; moreover, since we are holding true speed constant during cor- rection, the corrected Mach number is given by Wdw : VdF -dD) 7A:13 M : M Normally, the value of dF greatly exceeds that of DD 80 that the quantity dD may be ig- nored. However, in certain instances it may be necessary to compute dd to determine its order of magnitude. This may be accom- plished as follows: ww where Tg is the standard atmosphere tem- perature corresponding to the test pressure altitude. Provided the lift-drag polars at various Mach number are known (we note that it is possible to obtain these from flight tests), the quantity dD may be computed as follows: The test lift coefficient is given by dD :D-D ' W CL: and Yog MB S 2 Pom'' sco o 8 2 and the corrected lift coefficient by Ps M² SCO D: W 2 CL YD, M м? where Ps : test static pressure : standard static pressure. S 2 Test Conditions Corrected Conditions Energy height time derivatives w': W + dw W Energy height he he True altitude h' h : • hp Ambient temperature T': Tg+ dTg Ts • standard temperature at hp Mach number, M': M + dM 8 M Thrust, F' : F + df, under the test conditions P's, T's and N' F : standard thrust at Ps, Ts and standard RPM 7:42 Referring to the airplane polars, using the known values of CL and M, the drag cor- rection dD is readily obtained. To obtain the thrust which the engine would produce in a standard atmosphere, standard engine correction procedures are used. When thrust measuring equipment is not available, engine manufacturers' data may be used to compute standard thrust, with the procedure being as follows: operation within the airplane, although the use of the derivatives of/a (N/00) and of/oM does provide improved accuracy, since gen- erally the slopes of the curves do not change as much as the values of the functions them- selves. In the event that thrust measuring equip- ment is available, the actual thrust developed may be determined. This is accomplished using the following considerations: For choked or unchoked flow in the tail pipe we find (from the De Laval equations) that the gross thrust FG can be expressed by the relation The relation between thrust and the vari- ables on which it depends for a standard jet engine (without variable exit area or twin spool arrangement) is the well known equa- tion N, M, F : f VO 7A:17 FG 8 18) 8 AB Ps 7A:19 where FG 2 gross thrust where 8 = Ts/Tso 0 To/Tso: ambient temperature/ sea level standard temperature : 00 Ag : tail pipe exit area 8 P/P. Pry : total pressure in tail pipe aft of turbine M : Mach number ~ VIT 8 = Ps/P80 / : pressure ratio Ps ambient static pressure 3 and For this analysis, V : constant and Ps • constant (therefore, 8: constant). However, the temperature varies so that N/V and M are variable. Thus, 《) C**-(32)(7=- : KD for unchoked flow af N so dF : F'-F 4 ) ( AM ам DIN./0) 7A:18 7A:20* and : KD SO so (1. 26 )ror -1 Pt 1.26 Ds for choked flow. where AM is computed using the test speed and the test temperature in comparison to the test speed and standard temperature at the test pressure altitude, In Eq. 7A:18 we may correct at constant N if this corresponds to the standard value of N or A(NWO) may be computed for both variations in N and Jo. Generally the de- rivative df/oM is small and the first term in Eq. 7A:18 is predominant. Frequently engine charts do not present too accurate a picture of the actual engine 7A:21* . * These equations have been rearranged somewhat from their original form to permit the utilization of the gross thrust meter de- veloped at the USNATC. 7:43 () PS (+1)/Y : 27 * W = KQGVy-D)R open your ly1 Pr The transition from choked to unchoked flow occurs at about P1/Ps : 1/85 for con- ventional nozzles. The factor Kappearing in Eqs. 7A:20 and 74:21 is an efficiency fac- tor accounting for non-isentropic conditions, whose value(s) should be determined by static thrust stand calibration as a function of (Pty/Ps). for unchoked flow 7A:27 Using the test values of (P/Ps) the test value of FG may be computed from Eqs. 7A:19, 7A:20, and 74:21. The constant Kog is the discharge effi- ciency factor comparing the assumed isen- tropic flow to the actual. For many nozzles KOG * 0.9; however, its value should be de- termined from ground tests as a function of pressure ratio. The gross thrust is defined as FG QVex 7A:22 where Q 8 mass flow per unit time From the above considerations, we may determine the mass flow through the engine when values of Tto, Ps and Pr, are measured. Therefore, using'Eq. 7A:24,' the value of F' may be determined. . Vex : effective exit velocity : Now we are interested in the net thrust which is F: FG-FR 7A:23 where FR 8 : Ram drag : QV, and V is the V flight velocity. Thus, The question now arises as to what are the desired standard values of Pta Tt, and Ae (for a variable tail pipe area installation). The engine manufacturers' data are fre- quently the only source of this information, however, the specified values of tail pipe pressure and temperature at a given RPM may not actually be attainable in flight, so that these data may be of limited use. Nonethe- less, in the absence of other data, such as is attainable by flight under nearly standard at- mospheric conditions, the manufacturers' data must be used. . F: FG - QV. 7A:24 To determine Q, we again turn to the De Laval nozzle equations, from which The QTTE AePt, o Regardless of the method followed to ob- tain the standard values of pit, Tt, and Ae, we have the following relations (assuming Ps and V are constants): 7A:25 where Tit • total tail pipe temperature (a) From Eq. 7A:19 PT. Pr : total tail pipe pressure FG : Ap 84 Ap80 (: FG = A' 84 sos y : KOG 금 ​(유 ​$)**" 2 lly+1)/(Y-1) y+1 R :. dFG : FG-FG constant for choked flow Аеф Ps 7A:26 7:44 Similarly, 7A:5 DETERMINATION OF CORRECTED W = dhe/dt AND CORRECTED TIME TO CLIMB dFR = FR-F' : VIQ'-Q) - VdQ ) = where, from Eq. 7A:25 The test or observed value of w must be corrected for thrust, drag and horizontal wind variations. From Eq. 7A:13 we have that the change in w due to thrust and drag variations is AePtt Q = y Т. dw = (DF-dD) W and Frequently dD is small and may be ig- nored so that V dw : DF W Aept, dQ Aept /те W (5) TH where dF is given either by Eqs. 7A:18 or 7A:26. 7A:28 Due to gradient wind effects, we have an additional increment, (Eq. 7A:29) so that dF : dFG - VdQ. V dw = Obviously, the correction may also be accomplished using equations obtained by differentiation of Eqs. 7A:19 and 74:25. co dV dh dh/wind dt g Thus, dh W w >13 dF ch in /wind wind dt 7A:30 It remains here to consider the effects of atmospheric disturbances. As previously noted in Chapter 7, it is difficult to account for vertical air motions since their measure- ment is not easily accomplished. Therefore tests should not be conducted under known conditions of significant vertical atmospheric motion. To account for a known vertical gradient of horizontal wind during continued climbs, we have the relation (from Eq. 7:16) that ) We have demonstrated that the standard atmosphere energy height corresponding to the test pressure altitude and speed is given by Eq. 7A:16, which we rewrite as ne-he-n' the . 7A:31 v dw : co). g dt/ wind. Using these two equations, 7A:30 and 7A:31, each raw data point may be corrected to give the standard variation of wvs. he. Then since hea If the wind gradient is dv/dh and the ac- tual climb rate is dh/dt then dhe I hel 7A:32 dw : o ) V dV dh dh wind dt ) 7A:29 we may prepare a plot of 1/w vs. he as shown 7:45 in Fig. 7A:4 and the shaded area represents the time to climb from he, to hez: Data may either be recorded using a photopanel or may be transmitted to a ground station. Appropriate measures should be taken to identify runs and to insure proper functioning and continued calibration of the instrumentation. Note should be made by the pilot of operating limitations other than those specified by the manufacturer (for ex- ample, excessive build up of tail pipe temper- ature preventing attainment of rated RPM). After it has been ascertained that the flights were properly made and that the in- strumentation was functioning properly dur- ing the tests, the data may be reduced to standard, First, consider the case of the continuous climb to ceiling. The raw data at each point are first corrected for (a) instru- ment errors, (b) position errors and (c) lag errors using the procedures described in Chapter 1 of this manual. Following these corrections, the test data (again using the procedures explained in Chapter 1) are con- verted to: nez Fig. 7A:4 7A:6 SUMMARY OF PROCEDURES (1) P's test ambient pressure We shall here indicate briefly the proce- dures to be followed in obtaining and reduc- ing energy climb test data. (2) hp test pressure altitude First, we must provide means of meas- uring the following basic quantities: (3) Vcal test calibrated airspeed (4) M' test Mach number (a) Time (5) T's test ambient temperature (6) V' test true airspeed (7) c' test speed of sound (b) Static pressure (system calibrated at all altitudes) (c) Dynamic pressure (d) Total temperature (e) Engine RPM (1) Fuel flow (g) Normal acceleration (8) W' test gross weight (from fuel flow data). We now compute the tapeline altitude for the points in question using either Eq. 7A:7 For engine thrust determination the fol- lowing should also be measured: R Alt. h : Ean : S.L. (h) Tail pipe area (i) Tail pipe total pressure (j) Tail pipe total temperature T2 + TI P2 + PI # =( Tato) (P2-DID 7:46 2 or Eq. 7A:8 of (12) and (14) h : T dhp • To st (15) df If it is considered necessary to correct for gradient wind and drag effect, we com- pute (dw)wind from Eq. 7A:29 For continued climbs, a systematic pro- cedure may be worked up for calculating Ah between succeeding points with h being given by the sum of the Ah's to the given point. Thus, (dw)wind V dV dh g dh /wind dt (9) h, tapeline or geometric altitude. Where dh/dt is obtained from (a) and (9); dD is computed according to the method de- scribed in section 7A:4. This gives Knowing h and V, we compute he: (10) he h + V2/2g : (16) dD From a plot of he vs. time, or using Eq. 7A:12 (17) (dw)wind. Tan Bn From (15) and (16) and Eq.79:13, we com- pute dw = V/W (dF-dD) wn 12 DI (18) dw : 8 V/W (DF-dD) then we compute w' (19) w: w' (18) - (17) . (11) w' . dhe /dt. : From Eq. 7A:31 From the engine data we next determine the test values of net thrust using either Eq. 7A:18 or 7A:28 he ne : the ch' the, thus né -n't ho (20) he = (10) - (9) + (2). : O (12) F': engine net thrust. We now correct to standard atmosphere conditions holding the following quantities constant: If we are correcting data for continued climbs to ceiling, the standard time to climb must be determined. This is accomplished using items (19) and (20) for all data points up to the given point and Eq. 7A:32 2 dhe 1 hp : pressure altitude 3 V = true speed : W W: weight (21) t : corrected time to climb to he. At hp and V we determine After the data from all runs have been corrected we proceed as follows: () Τ (13) Ts : standard temperature at he (14) Pst: standard thrust at hp and V. (a) For continued climbs to ceiling the data from all runs are plotted in the form of Fig. 7:4; i.e., in the form of he vs, time. Cross plots are then prepared as in Fig. 7:5 We next determine dF from the difference . 7:47 to determine the velocity V for wmax at any value of he. Knowing V and he for wmax we may compute the standard altitude for Wmax from 7:8 then prepared. This immediately deter- mines the climb schedule V = f(h). When this method of testing is used, the time to climb should be determined by a continued climb to ceiling according to the optimum schedule V = f(h) with the time to climb and other data corrections being made by the methods outlined in this section. v2 h = he - 2 29 and continuing the process we thus deter- mine the maximum energy climb schedule V = f(h). 7A:7 CONCLUDING REMARKS For pilot information this schedule should be finally presented in terms of Vcal vs. f(h). (b) If tests were conducted utilizing ac- celeration runs, the corrected data are plotted as in Fig. 7:6 and the cross plot of Fig. Mr. Renaudie's paper contained consider- ably more detail than this abbreviated ver- sion; however, it is felt that the missing details can readily be filled in by the flight test engineer working with a particular problem. 7:48 AGARD FLIGHT TEST MANUAL VOLUME I, CHAPTER 8 TAKE-OFF AND LANDING PERFORMANCE PART I FLIGHT TEST ANALYSIS By F. E. Douwes Dekker National Aeronautical Research Institute Amsterdam, The Netherlands PART II THEORY By D. Lean Royal Aeronautical Establishment United Kingdom ش ش کن TERMINOLOGY (Continued) Lift Coefficient Drag Coefficient Power Coefficient Ground Run Wind Correction Factor Unit Fw FB Ground Run Slope Correction Factor rad. rad. Y Flight Path Angle rad. rad. B Slope (uphill positive) P Air Density kgs2/m+ (lb) (sec2) ft+ σ Density Ratio relative to sea 8 Absolute Temperature Ratio level standard 8 Pressure Ratio μ HB Subscripts 0 1 2. S . t A Propeller Efficiency Coefficient of Rolling Resistance Coefficient of Effective Braking Resistance for All Wheels Sea Level Standard (also standstill) Take-off or Touch Down Point At Obstacle Height Standard Test Increment of Test Quantity to Reach Corrected or Standard Quantity A1, A2 etc. are Constants 8:12 8:13 (c) Lift Coefficient Corrections (d) Landing Assistance SUMMARY OF EQUATIONS CONCLUDING REMARKS CONTENTS (Continued) Page 8:18 8:18 8:18 8:19 PART II THEORY 8:14 GENERAL PRINCIPLES 8:20 8:15 THE GROUND RUN 8:20 8:16 THE TRANSITION AND CLIMB 8:23 8:17 PILOTING TECHNIQUE IN THE AIRBORNE PHASE 8:29 8:18 APPLICATION TO FLIGHT TEST PROCEDURE 8:31 8:19 CRITICAL SPEED DURING TAKE-OFF 8:31 8:20 THE LANDING MANEUVER 8:33 8:21 ESTIMATION OF LANDING DISTANCE 8:33 8:22 THE CHOICE OF THE LANDING APPROACH AIRSPEED 8:37 8:23 SOURCES OF VARIATION IN TOTAL LANDING DISTANCE 8:39 REFERENCES 8:40 TERMINOLOGY Unit m/s ft/s RC Rate-of-Climb RS Rate of Descent D Total Aerodynamic Drag = 0.5 v2 CDS kg lbs E Total Energy W/2g (V²+2gh) kgm lb ft L Total Aerodynamic Lift = 0.5 v2CLS kg lbs N Engine Speed RPM RPM S Wing Area ft2 S Horizontal Distance in Take-off or Landing S $2 Ground Distance Air Distance EE E ft ft ft V True Airspeed m/s ft/s < True Airspeed at Take-off Point (getaway) or Touch Down m/s ft/s V₂ True Airspeed at Obstacle Height m/s ft/s W All-up Weight or Gross Weight kg lbs P Power BHP Engine Brake Horsepower F Net Engine Thrust kgm/s ft lb/s kg lbs J Advance Ratio a Acceleration m/s2 ft/s2 d Deceleration m/s2 ft/s2 60 C g Acceleration Due to Gravity m/s2 ft/s2 h Obstacle Height m ft t Time S 's Vw Surface Headwind m/s ft/s TERMINOLOGY VOLUME I, CHAPTER 8 CHAPTER CONTENTS Page PART I FLIGHT TEST ANALYSIS 8:1 SUMMARY 8:1 8:2 TAKE-OFF DEFINITIONS 8:1 8:3 TAKE-OFF MEASUREMENTS 8:2 8:4 BASIC TAKE-OFF EQUATIONS UNDER NO WIND CONDITIONS 8:4 (a) Ground 8:4 (b) Air Phase 8:5 8:5 PARAMETER THEORY 8:7 8:6 TAKE-OFF PERFORMANCE REDUCTION 8:9 (a) Wind Corrections 8:9 (b) Slope Corrections 8:10 (c) Lift Coefficient Corrections 8:11 (d) Take-off Assistance 8:12 8:7 SUMMARY OF EQUATIONS 8:12 8:8 LANDING DEFINITIONS 8:15 8:9 LANDING MEASUREMENTS 8:15 8:10 BASIC LANDING EQUATIONS 8:16 (a) Ground Run 8:16 (b) Air Phase 8:17 8:11 LANDING PERFORMANCE REDUCTION 8:17 (a) Wind Corrections 8:17 (b) Slope Corrections 8:18 Although the choice of h is fixed by appli- cable specifications, it is supposed for this discussion that the air phase is a transition maneuver, which begins when the aircraft becomes airborne and continues until the altitude h and speed V2 are attained simul- taneously. It is assumed that take-off power is maintained until the air phase is completed, that the flap setting remains constant during this phase and that the landing gear is retracted as soon as safety permits. To keep the scope of this chapter within practical limits, the discussion will pertain exclusively to land planes with take-off weights of more than 2000 kg (4400 lbs) operating from hard surface runways (μ≈0.02). Light aircraft utilizing hard or soft turf require special analysis for each given case as do multi-engined airplanes under simulated engine failure conditions. The determination of critical engine failure speed and accelerate-stop distances will not be considered here. The accuracy of a given measuring device or instrument or method is assumed to be such that there is a probability of 95% that the difference between the actual and measured values will be less than the stated error (95% confidence limit). Throughout the following work, suffixes s and t are used for standard and test conditions respectively; however, subscripts are omitted in evident cases. The increment A, added to the test quantity of a variable, will give the standard or corrected quantity, i.e., sc = S+As. The definitions of the take-off mentioned above are generally used for take-off per- formance reduction. The formulae appearing later in the chapter are based on these definitions. However, a different approach PART I FLIGHT TEST ANALYSIS 8:1 SUMMARY In this chapter the basic principles of take-off and landing performance testing are presented. The object, means and accuracy of the measurements are discussed. The basic equations underlying the performance reduction methods are derived and explained. 8:2 TAKE-OFF DEFINITIONS A take-off is defined as the process by which an airplane is safely brought from standstill to a safe flight condition. The safe flight condition is defined as that point in a climb at which the airplane has first reached a height of h meters or feet (obstacle height) above the point of departure from the ground and an instan- taneous true airspeed of V2 meters per sec. or ft. per sec. The take-off is divided into two parts: the ground run and the air phase. The air phase begins when the airplane becomes airborne. In the past, the choice of the obstacle height has been made somewhat arbitrarily. Usually 15.2 m (50 ft) has been taken as the required height above the runway, but there is now a tendency to choose 10.6 m (35 ft). The choice of h affects the reduction method because it determines whether or not a steady climb will be achieved before the obstacle height is reached. Generally, high average acceleration during take-off and/or low obstacle height h eliminate the require- ment for subdivision of the air phase into transition and steady climb phases. 8:1 to the problem is presented in section 8:5. 8:3 TAKE-OFF MEASUREMENTS The principal aim of flight testing is to collect reliable data which may be used to predict airplane performance under standard conditions as well as under certain non- standard conditions. Uncontrollable factors such as pilot skill must be eliminated from the tests as far as this is possible so that the uncontrolled variation in the quantities to be measured may be reduced to a minimum. Take-off tests should be delayed until a late stage of the prototype test program in order that some agreement between pilots regarding a standard take-off technique may be arrived at. Moreover, for a take-off test the airplane should be operated in such a way that repeatability of data is assured, rather than in a manner leading to the best take-off performance.* This means that during a take-off sudden changes in altitude and configuration should be avoided unless repeatability of the action is assured. The whole maneuver should be carried out gently and safely. The basic features of take-off performance of the air- plane are the topic to be investigated and not the abilities of the pilot. The accelerations obtained during the ground run and air phase may be regarded as fundamental characteristics of the take- off. However, it is usual to measure the take-off performance in terms of ground and air phase distances under standard and other conditions. These distances détermine the mean accelerations defined later in the text. Generally, the standard take-off distances must be determined within about 5%. The accuracy of the final reduced data is deter- * Editor's note: This statement, it is believed, applies generally to commercial aircraft but not necessarily to all military machines. mined mainly by the accuracy of the reduc- tion process, the test accuracy and the number of tests. The test accuracy depends on the accuracy of airspeed measurement at the point where the airplane leaves the ground and at the point where the airplane has reached the height of h, and thus on the test equipment available. Supposing a 4% natural scatter in the true distance (due to uncontrollable factors), it can be determined from a probability analysis that with high accuracy test equipment a minimum of five tests for one configuration are required. Generally, a test accuracy of 2.5%, obtainable with six runs, should be aimed for. One or two additional runs allow a decrease in the required test accuracy to 3.6% and 5% respectively while maintaining the overall 5% accuracy. The actual ground speed at the point where the airplane leaves the ground, the ground speed at obstacle height, ground distance, air distance, air phase time, surface head- wind, air temperature and air pressure are the principal quantities to be measured. A full time history of the take-off is usually accepted as a must, but accurate take-off analysis shows that the measurement of the speed and height at the two stations defined above is sufficient for the purposes of this analysis. In the latter case the test equip- ment may be simplified considerably without reducing the accuracy of the results. Records of only the two portions of the airplane's flight path are necessary: (1) where the airplane becomes airborne; (2) where it attains the obstacle height. For detailed investigations, a complete record of time and distance or acceleration is required. All test methods require knowledge of • actual net thrust at about 0.7 V and at 0.5 (V₁+ V₂).* This involves measurement *V is the speed at which the airplane leaves the ground and V2 is the speed at the altitude h. 8:2 of at least the following: engine speed, manifold pressure (jet tail pipe temperature), air temperature, air pressure and humidity. The power setting during take-off must be recorded either by the pilot, observer or automatic recording equipment. For jet- powered airplanes the use of the latter is preferable to provide an accurate record of engine speed. The actual net thrust as well as the standard net thrust may be estimated from engine manufacturer's data or may be measured by means discussed elsewhere in this manual. All controllable variables should be held as nearly standard as possible with maximum operational take-off weight usually being chosen as the standard weight. It will be worth the effort required to refuel frequently in order to hold each test take-off weight within 5% of the standard value. Tests on very hot or wet or iced runways should be avoided: moreover, tests should not be attempted under crosswind or gusty conditions or when the windspeed exceeds 10% of V2. The initial acceleration does not affect the ground distance appreciably so that the use of maximum power with the airplane restrained by the use of brakes prior to the start of the take-off is not required. However, the throttle should be fully opened as soon as possible after commencement of the ground run. It is noted that initial acceleration does affect ground run time; however, the use of measured ground run time is avoided in the following analysis. Determination of the exact point at which the airplane leaves the ground has always been a difficult and rather confusing problem. Direct observation, camera, or accelerome- ter recordings do not provide the means for accurately determining this point. Only eventmarkers on the main undercarriage legs might solve the problem. However, for this analysis the exact location of the point at which the airplane first becomes airborne is considered unimportant, compared with the location of the point where the airplane has achieved the obstacle height and safe flight conditions. For the purpose of this anaysis, therefore, it is emphasized that accurate measurement of the ground speed near the point where the airplane leaves the ground is more valuable than determination of the exact location of the point. The ground distance required to achieve a certain kinetic energy level, based on a safe true getaway speed appropriate to actual wind, weight and density conditions is considered a reliable perform- ance parameter, as is the total horizontal distance required to reach a certain total energy based on obstacle height and minimum safe equivalent climbing speed. Some pro- posals based on this concept are referred to in section 8:5. Present day equipment used for actual airport-level air pressure, air temperature, humidity, wind speed and wind direction measurement is adequate for purposes of ob- taining take-off test data. Self-counting ane- mometers giving the average wind speed during the take-off are preferred. The track or path of the airplane is usually recorded by motion picture cameras placed either at the side of the runway or at some distance up or down the runway. Plate cameras with a moving diaphragm and a very accurate timing device are also in use. All photographic systems, however, lose accu- racy for very long take-off distances which require more cameras and complicated syn- chronization. When ground and air distances under the test conditions can be estimated well enough beforehand, it may be sufficient to measure the ground speed by photographic or photo- electric means at only the two predetermined stations, (1) and (2), in which case the problem of synchronization can be simplified. 8:3 8:4 is: BASIC TAKE-OFF EQUATIONS UNDER NO WIND CONDITIONS (a) Ground Run The ground distance s, (no wind, no slope) V₁ Sds = s₁ = √ VI s ds ᏧᏙ 1 dt dt dv 2 s dv² 2a = vi 2a 8:1 The actual acceleration a during the ground run is a function of V and decreases from the initial value ao to a final value of a₁. A mean acceleration ā, giving the same overall speed-distance change as the actual variable acceleration, can be defined by the equation For the average jet-powered airplane, V/V₁ = 0.70±0 02 for a₁/= 0.8. For special power units as well as mixed power systems(turbo-prop) the value of V/V₁ can be obtained from Fig. 8:1 for an estimated value of a₁/a。. Using the above approach, actual solution of the basic integral (Eq. 8:1) is avoided, and the performance reduction can be based on the use of the simplified concept of an average acceleration a When the coefficients of drag, lift and rolling resistance are kept constant during the reduction process, it can be stated that V² = A₁ W/o and µ(W− ¯ ) +Ō = A₂W 10 = // [F - μ (W - [) - ō] W where A, and A2 are constants. Then the expression for the ground distance becomes 11 A₁ W/O 2g (틉​-A2 8:3 The difference between the ground run distance obtained during test and the distance existing under standard conditions of weight, density and thrust may be approximated by logarithmic differentiation of Eq. 8:3, whence F ASI A(W/o) W SI W/σ 8:2 where F, L and D are mean thrust, mean lift and mean drag, respectively, at a mean speed V. The ratio V/V depends mainly on the acceleration ratio a/a。 and the functional relationship between a and V. However, including all possible functions a = f(V), i.e., linear, square, etc., it can be stated that: For the average reciprocating-engined airplane, V/V₁ = 0.74 ± 0.02 for ª¡/。 = 0.4. a/ao 8:4 where and A(W/G)= W (AW - AG) A (F/W) - F/AF AW W F W s₁ = ground distance corrected for wind, runway slope and C₁ V₁ = true ground speed at the point where the airplane becomes airborne (cor- rected for C₁) W = F 11 ။ so that ā = Asi Δω Δσ Fg + SI W Wa (AW - 4F) ΔΕ م test weight test density ratio calculated mean thrust at V(for test power setting and atmospheric con- ditions) v² /2s, (b) Air Phase The air phase distance s₂ (no wind, no 8:4 slope) is: $2 S s2 = SI The true speed V₁ will change by an amount AV₁, consequently ۵۷۱ AVI = 0.5 AW W Aσ 0.5 م dE dE dt 1745 - 7 (1/4) E ds = as 013 E2 s WEI dE v2 – v³ + 2gh 2ā 8:6 w/2g where the total energy at point lis E₁ = (Vi) and at the height h (point 2) is E2 = W/2g (V² +2gh). The energy increase dE/dt is equal to the useful work accomplished per second (T-D)V, where the excess thrust (T-D) is equal to the product of mass and flight path accelera- tion a, relative to the gravitational field. The mean acceleration ā at the mean true speed V can be expressed as 이 ​g W [(F-D)]. 8:7 As the speed increase during the air phase is usually small, the mean speed V can be 8:5 It should be emphasized that Eqs. 8:4 and 8:5 are only valid for small increments ▲, and for cases where s and V have been previously corrected for wind, runway slope and C₁ as discussed in section 8:6. Similar relations for larger corrections and an anlysis of the relation between net thrust and effective accelerating force are presented in Ref. 1. Generally, the increments AW, Ao and AF should be determined with an accuracy of about 0.5% of the test values of weight, density ratio and mean thrust, respectively. Summing up, the following test data are required when the reduction formulae 8:4 and 8:5 are used: 8:5 When taken equal to 0.5(V₁+ V2). The terms F and D are mean net thrust and mean aerodynamic drag, respectively, at V. CL, CD and h are assumed constant during the reduction process, the true speeds squared V and V2 are proportional to W/o, and the drag is proportional to W so that v2 - v² + 2gh + 2gh = Az W + 2gh so that As2 $2 v2 - v? Vẻ - Vĩ + 2gh and Δσ (AW - D=) gF /Aw AF -(-) W 8:9 Because the observed true speeds V₁ and V2 also change by amounts AV, and ▲V2, we have D = A4W where A3 and A4 are constants. Accordingly, the expression for the air distance becomes 52 11 A3+2gh 2g W 29-A4 8:8 An expression approximating the change in air distance for small variations in weight, density and thrust is obtained by logarithmic differentiation of Eq. 8:8, whence A(W/o) A(F/W) As2 A3 2052 - g 52 where and A(W/σ) = १|ह Δω Δσ W م ā Δω ・E (AF - AW) W A(F/W) = AVI AV2 V2 = 0.5 AW W - 0.5 ala 8:10 Similar relations for large corrections and details about the kinetic/total energy increase ratio and the net thrust/accelerating force ratio are given in Ref. 1. The required accuracy for the increments AW, Ao and AF can be taken to be 0.5% of the test values of weight, density ratio and mean thrust respectively. Application of Eqs. 8:9 and 8:10 requires the following test data: s2 = air distance, corrected for wind, runway slope, and CL V₁ true getaway speed, corrected for CL V₂ = true speed at obstacle height, cor- rected for CL W = test weight b 14 test density ratio F = calculated mean thrust at V, cor- responding to test conditions = (V² − v² + 2gh)/282 8:6 8:5 PARAMETER THEORY Dimensional analysis has often been used to determine the controllable parameters af- fecting engine performance, in particular when dealing with jet engines. When the results of such an analysis are combined with the functional relations for the airplane in a specific flight condition, valuable per- formance reduction equations are obtained. The basic ground run Eq. 8:3 SI ** 2g A, W/O F W 1 A2 8:3 becomes, on substituting σ = 8/8 S1 AWS 8 F/8 29 W/8 - A2) 8:11 Thus, ignoring the influence of airspeed* in this case, the general relationship will be N 8 W 8 where Ñ = mean engine speed at V. 8:14 This latter Eq. 8:14 is for a specific value of C at V and a fixed value of runway friction coefficient μ. Also the dis- tance s is the test value, corrected for headwind, runway slope, and preferably for C₁ at V₁. A similar relationship applies to the air distance (see Ref. 17) $2² = 1. ( W ) f 8 8:15 or more generally SI = f 8 8 W S 8:12 or for a jet airplane, ignoring speed (Mach number) variation 52 = f 8 N (뽕​) which is a "non-dimensional" relationship, with the constants S, CL, CD and g omitted. For a jet-propelled airplane, Eq. 8:12 may be combined with the well-known relationship (see Ref. 6) F 등 ​(.岩​) N 8:16 8:13 It should be noted that poor intake effi- ciencies at low airspeeds may affect the validity of the basic Eq. 8:13. Some experi- ence concerning this is reported in Ref. 17. * Editor's note: The quantity V/ is actual- ly proportional to Mach number, hence the statement that we ignore the influence of airspeed is equivalent, in this case, to stating that the Mach number effects are unimportant. This is, of course, a proper assumption for take-off analysis. 8:7 For reciprocating-engine airplanes and F = " P V n = f(Cp, J) = f P so that F/σ = f(P/o, N, V) and the mean take-off thrust F/o= f(P/o, N,V). For constant C, the mean speed V is proportional to W/o and the take-off mean thrust for the ground run or air phase is N W f σ 흥아 ​(흥​,지​, 뽕​) 8:17 Combining Eqs. 8:17 and 8:3 gives for the ground run (E. W. N) (P and N at V) Ñ S₁ = f 8:18 which equation also applies to the air phase and pertains to specific C values at V₁ and V2. Thus, it is seen that the take-off distances can be related to controllable parameters, being combinations of controllable and free variables. Theoretically it should be possible to carry out every test at standard values of these parameters, thus avoiding any correc- tion. In practice, however, accurate control of the parameters is sometimes not feasible, and cross plotting is required. Possibly the functions derived here might be calculated during the design stages and checked, as well as corrected, by take-off tests. This, however, would require a great number of tests, at least three for every combination of two parameter values. The advantage is that a three-dimensional "car- pet" plot obtainable from the test data would enable us to predict the take-off performance for any non-standard condition in a relatively simple way. In the performance reduction method pre- sented in the preceding sections, ground and air distances must first have been corrected to specific C₁ values at V and V2. This means that exact determination of the test distance at which the airplane becomes air- borne is not really required. Rather, atten- tion should be directed to the accuracy of speed measurement at the point on the runway where the kinetic energy is expected to reach a level sufficient to lift the airplane from the ground under the given conditions. Following this same line of thinking, accurate measurement of speed and height in the air are required only at the point where the total energy of the airplane is expected to reach the standard desired value (at this point speed and height separately should be near their respective standard values). Cor- rection for wind effects requires measure- ment also of the time required to travel between points 1 and 2. If the philosophy of the analysis presented here is applicable to the type of operation for which the airplane is intended, the given test reduction procedure leads to high accu- racy at low cost. Fully normal and repeata- ble test take-offs can be performed without requiring attainment of the shortest possible take-off distance. The forced pull-off climb as well as the held-down zoom are not needed as part of the take-off test techniques. The test data are corrected to optimum CL values at points 1 and 2. This requires prior agreement about the standard optimum take-off technique, and an investigation into the efficiency of transformation of kinetic energy into potential energy, and conversely. 8:8 8:6 TAKE-OFF PERFORMANCE REDUCTION (a) Wind W When the ground run is made into a constant surface headwind of velocity Vw, the initial condition of the airplane is such that the ground speed is zero and the air- speed is equal to the windspeed V. For a still-air ground run, a ground distance Sw is required to accelerate the airplane from standstill to the airspeed Vw to reach the initial condition mentioned above, assuming that the coefficient of rolling friction μ is independent of ground speed. From this point on, the distance traveled relative to the air to reach getaway air- speed V₁ will be the same in both cases with or without wind; the airplane accel- erates from an airspeed Vw to an airspeed V₁. In the first case, however, (ie., with wind) the ground speed will always be Vw less than under equivalent conditions during a still-air ground run. Therefore, the ground distance will also be Vwtw less, where tw is the time required for the acceleration from Vw to V true airspeed. The influence of a headwind on the total ground distance can be expressed by the relation SIW 11 = | Sw - Vw tw 8:19 or or W V VdV a VW VdV Vw Vw dv (V-Vw) dV a 11 Vw VI-VW (V_Vw) d (V-Vw) (VI-Vw)² 2ā where the mean acceleration ā occurs at the mean ground speed (V/V₁ )(V¡ -Vw) or the mean airspeed (V/V₁)(V₁ -Vw)+Vw for the still air ground run (see section 8:4 for values of V/V₁). W The basic equation for the still-air ground run was S 11 v² 2a where a = acceleration at V. Therefore, the ratio Fw of the ground distances without and with headwind is where W = ground distance with headwind V 2 V W の ​S| ground distance in still air =s₁+Ası Sw tw = ground distance from standstill to speed Vw, still air = time elapsed from speed Vw to VI Fw = 5 | W or in exponential form = still air 8:9 8:20 where V₁ -V = C = ground speed at getaway with wind Vw = surface speed at getaway with wind acceleration at (V/V₁){V₁−Vw)+Vw acceleration at V (for still air) An analysis of the acceleration ratio C for different acceleration characteristics re- veals that the exponent p is practically in- dependent of Vw and can be regarded as a function of the final initial acceleration ratio a/a。. In Fig. 8:2, Fw is given as a function of Vw/(VI -Vw) and aι/ª。. If no other data are available, the acceleration ratio a/a。 can be assumed 0.8 (p = 1.92) for jet airplanes and 0.4 (p = 1.68) for propeller airplanes. W Thus, for purposes of this analysis, the air distance correction for headwind is simply As₂ = Vw 12 8:21 For rather strong surface winds, the wind gradient might be accounted for by taking the magnitude of the effective surface wind Vw to be 10% higher than the actual. (b) Slope Corrections The influence of a runway slope B on the ground distance is given by = s₁ + As₁ = FB • S1 8:22 where Fg ground run slope reduction factor and s₁ = test ground run with slope ß. An up-slope ẞ reduces the effective mean excess thrust by the amount BW. This is equivalent to an acceleration increment of = Aã = gß where B mean runway slope at the test, uphill positive. So From Eq. 8:1 it follows that Asi Δα SI ASI 2gs, B v} The influence of a surface headwind Vw on the air distance can be divided into two parts: (1) The distance travelled by the surface wind during the air phase time t2. (2) The increase of effective accel- eration, due to wind gradient. With a headwind of speed Vw at the ground and speed Vw+dVw at obstacle height h, the acceleration due to wind gradient is roughly equal to dVw/t2 According to the Prandtl wind gradient equation, the wind at 15 m (50 ft) is equal to 1.25 times the surface wind at 3 m (9 ft). In which case, the decrease in air distance due to the extra acceleration should be roughly (Vw/8)†2, which is less than 1% of the air distance when the surface wind is less than 0.1 V2. Therefore, the influence of the wind gradient is usually ignored; moreover, actual conditions can differ appre- ciably from the Prandtl law. 8:10 and Thus s₁ + As₁ = s AS (1-20618) FB = | 2gs, B v² 2 8:23 correction to specified getaway and climb speeds, respectively. The differential equation for ground dis- tance can be written ds 의스 ​dv₁ which when combined with the basic equation v²/2ā gives S dsl ā dvi 2 S 8:25 The influence of B on the air distance depends on the definition of h. There will of course be no influence in case his defined as the tape line altitude above the point where the airplane becomes airborne. Sometimes, however, h is measured per- pendicular to an average runway plane with slope ẞ. In that case, the excess thrust reduction BW will in effect continue during the airphase and can be included in AF of the general reduction Eq. 8:9 or treated separately A52 52 Δα gB 이 ​For constant weight and density it may be deduced from W = 0.5p。 S CL, Vic that *CLI CLI 11 - 2 consequently, for small increments Δ Asi SI a ō ACLI CL 8:26 Due to the change in C, the getaway speed is changed by ACLI ۵۷۱ - 0.5 CLI 8:24 (c) Lift Coefficient Corrections Up to this point the performance reduction equations have been developed keeping Cand CD constant. In most cases, the pilot some- what arbitrarily chooses the lift coefficients at getaway and during the air phase, and it is probable that the C values at getaway and at the obstacle height will be different for all test take-offs introducing some scatter in the reduced performance data. Therefore, ground and air distances should be corrected to specific values of CL, which amounts to 8:27 Later on in this analysis when corrections for weight and density are considered, the standard value of CL will be held constant and V₁ be further modified. 8:11 When no other data are available, the factor a/a, in Eq. 8:26 can be taken as 1.55 for reciprocating-engine airplanes and as 1.1 for jet-propelled airplanes. The influence of V2 on the air phase may be expressed by the equation V2 dV2 · ds2 02 8:28 air distance and so the total air phase correc- tion is As2 11 I vi v3 − v² + 2gh ACL2 CLZ S2 - As · Assuming the acceleration to be constant, Eq. 8:28 can be combined with the basic air phase equation $2 11 v2 – v² + 2gh 20 giving dV2 ds2 52 2v2 2 v3 − v² + 2gh V2 and at constant weight and density dCL2 CL2 Thus approximately - 2 dV2 V2 As2 $2 v32 v2 − vi + 2gh ACL2 CL2 8:29 8:31 Due to the CL2 correction, the airspeed at obstacle height will be altered by AV2 V2 -0.5 ACL2 CL2 8:32 (d) Take-off Assistance Full duration take-off assistance and the use of mixed power systems do not alter the fundamental reduction formulae. However, any change in the acceleration ratio a/a。 affects some constants in the equations and this must be accounted for. When part time assistance is used, it may be necessary to subdivide the ground run and/ or the air phase for purposes of analysis. In Ref. 1, it is proposed to use an effective mean assisting thrust in the reduction equa- tions, equal to the actual thrust multiplied by the (actual operating/total phase time) ratio. Accurate space time recording at the time of the start and finish of the use of assisting thrust is required. Therefore, in some cases a complete space time record may be neces- sary. 8:30 To account for the kinetic energy change, i.e., change in V₁ due to a change in CLI the ground distance increment As, due to the CL correction must be subtracted from the 8:7 SUMMARY OF EQUATIONS (a) Ground Run Measured values s₁, V₁, Vw, W, o, ß. All corrections are to be applied in the 8:12 and $2 order given below; all values SI appearing in the formulae are related to the respective distances corrected for the pre- vious effects. (1) Wind correction As -1 = Fw - 1 (See || | Fig. 8:1) (2) Slope correction Asi 2gs, B SI v} FB-1 s, corrected for Vw WIND CORRECTION FACTOR Fw 1.4 1.3 1.2 1.1 5 1.0 0.9 0.8 10 -5 -5 -10 15 (3) CL, correction ASI SI ā ACLI CLI AV ACLI – 0.5 V₁ CLI reciprocating a = 1.55 a jet a = 1.10 a s₁ corrected for V and B GROUND RUN ACCELERATION RATIO a 1.0 0.8 0.6 do 0.4 5 10 15 SURFACE HEADWIND GETAWAY GROUND SPEED VW V₁ = Vw)% EXPONENT P 2.0 JET 1.8 MIXED PROP 0.2 0.4 0.6 0.8 ACCELERATION RATIO 1.6 1.0 P Fig. 8:1 The Wind Correction Factor Fw= V1 Vw and Exponent p as Functions of Acceleration Ratio for the Ground Run 8:13 = 0.78 MEAN SPEED ▾ GETAWAY SPEED VI 0.74 PROP MIXED 0.70 JET ACCELERATION LINEAR v2 ACCELERATION LINEAR V (ACCELERATION) LINEAR v2 0.66 1.0 0.8 0.6 0.4 ACCELERATION RATIO བད " Fig. 8:2 The Mean Speed as a Function of Acceleration Ratio for Different Acceleration Characteristics During Ground Run Combination of (1), (2) and (3) gives: s, (corrected for wind, slope and CL)= Fw. - FB (1 - -1 - ACLI) (b) Air Phase Measured values s₂, V1, V2, Vw, σ, [2, W (1) Wind correction As₂ = Vw 12 S¡ test where SI test = Slw (2) CL2 correction and thrust V를 ​ACL2 52 - As As2 v2 −vî + 2gh CL2 (4) Weight, density correction ASI AW Ao Fg [Aw AF + م Wa W F SI where a = W vi 251 where As is the correction on ground run for ACLI V₁ corrected for CL s, corrected for Vw, B and CL, F = mean thrust at V = 0.74 V₁ for reciprocating engine and V= 0.70 V₁ for jet engine (see Fig. 8:2) AV2 V2 = − 0.5 ACL2 CL2 82 corrected for Vw 8:14 (3) Weight, density and thrust correction As2 $2 where AW Δε gF /Aw + wã \ w VZ-V₁ v2-v²+2gh \ w It م v2 − vi + 2gh 252 V₁ corrected for CL V2 corrected for CL2 S2 corrected for Vw, C CL and CL2 F mean thrust at 0.5 (V + V2) AF In Ref. 1, empirical constants are given for a proposed simplification of the reduction equations. These constants are based on extensive tests of several groups of airplanes having about the same characteristics. somewhat different approach to take-off data analysis is given in Chapter 6 of this manual. A 8:8 LANDING DEFINITIONS A landing is defined as the process in which an airplane is safely brought from a safe flight condition to a standstill. The safe flight condition is defined as a condition in which the airplane has a height of h meters (feet) above runway level and descends steadily with a true rate-of-descent of RS meters per sec. (ft. per sec.) or main- tains an angle of approach of y degrees with a true airspeed of V2 meters per sec. (ft. per sec.) in both cases. The landing is divided into two parts: the ground run and the air phase. The ground run begins as soon as the airplane touches the runway for the first time. To simplify the problem, the following assumptions are applied in the landing per- formance reduction theory outlined here. The air phase is considered as a transition maneuver from the steady safe flight condi- tion to touch down. It is assumed that the engine setting is constant and that the lift coefficient and deceleration increase until touch down, at which point the engines are throttled back. Moreover, it is supposed that the flap setting remains unchanged throughout the landing. During the ground run the coefficient of rolling friction is taken to be constant. The possible reversal of propeller or jet thrust is regarded as a safety feature and is not considered in the standard landing perform- ance analysis. Investigation of reverse thrust requires separate tests, preferably with ac- celerometers. The use of a brake or drogue parachute might be taken as a standard landing aid, provided that the moment of application is controllable. The theory presented here is limited to land planes having more more than 2000 kg (4400 lbs.) landing weight with nose wheel or tandem undercarriage and to landings on a dry hard-surface runway. It is assumed that only the main wheel brakes are applied. Maximum allowable landing weight is re- garded as standard, although it may be useful to conduct tests at other weights as well. The other variables, approach speed in particular, should be held as close to their standard values as possible. Only power-on approaches, characterized by a certain rate-of-descent or flight path angle, will be considered. Glide approaches for light aircraft are dealt with in Ref. 5. 8:9 LANDING MEASUREMENTS Prior to the time when formal landing 8:15 obstacle height, groundspeed at touch down, air phase time, air distance, ground distance, wind speed, air temperature and air pressure. Records must be kept of power settings, run- way state, brake disc temperature, and tire wear after landing. As in the case of take-off tests, a full time history of distance is not required; it is only necessary to measure the speed at two stations accurately, near the obstacle height and at touch down. Since the point at which the airplane is at obstacle height and the point of touch down may differ greatly for each test, the two recording cameras should cover a sufficiently large area to insure that records are obtained for the desired flight path regions. Landing tests should be divided into air- phase and ground run tests. Only a few ground runs (approximately three), utilizing maximum brake capacity at standard weight, should be performed to avoid excessive wear and consequent replacement of the tires. At least five air phase tests are necessary for the chosen approach technique which provides a given rate-of-descent or flight path angle. 8:10 BASIC LANDING EQUATIONS (a) Ground Run The equation for ground distance s (no wind, no slope) is developed in the same manner as the take-off ground run equation. The average deceleration d, giving the same overall speed-distance change from touch down to standstill as the actual variable deceleration, is given by g J = 1 [μg (W-C) + D-F] W B performance tests are conducted, the stan- dard approach technique should be esta- blished. The rates-of-descent of present-day airplanes in power-off glides are so high that the application of a considerable amount of power during final approach has become standard practice. This means that the type of approach is limited by predetermined values of rate-of- descent or flight path angle, and not by specific performance features of the airplane. Inas- much as instrument approaches require the longest total landing distances and determine the required landing runway length, it is obvious that this type of approach is of primary importance. The airplane should arrive at the obstacle height in a steady descent with a predetermined true airspeed, rate-of-descent or flight path angle. On the other hand, to obtain the shortest possible air distance, a different technique is required, the success of which depends mainly on the pilot's skill and his acceptance of risks. These factors tend to introduce scatter in the results. To insure that consistent approach data are obtained, the instrument type of approach is considered here. Again it should be empha- sized that for purposes of this analysis, there is no need to obtain exceptional performance data. The determination of the best standard approach within the operational limits of the airplane is the real aim of the tests. During the ground run, the use of excessive brake power should be avoided. In the past, it has been difficult to achieve optimum brake operation, and as of this time, automatic wheel brakes or any equipment providing direct control on brake power and tire condi- tions have not been made standard installa- tions. In this connection, it might prove useful to provide the pilot with an indication of horizontal deceleration as a guide to proper use of the brakes. During landing tests the following vari- ables must be measured: groundspeed at 8:33 where L, D and F occur at a mean speed V. 8:16 Keeping the lift and drag coefficients con- stant during the reduction process, an equa- tion similar to Eq. 8:4 can be derived Ast Δω Δσ gFAF AW + SI W Wd W م 8:34 The last term of Eq. 8:34 is usually ignored when the idling thrust is low, in which case the deceleration ratio d/do is roughly 1. Generally, the deceleration is a linear func- tion of V². For the case of constant decelera- tion the mean speed V has a limit value of 0.7 V₁ · (b) Air Phase The air phase s2 (no wind, no slope) is given by dE V2 - Vi+2gh $2 E2 g ds = 52 W d E₁ d 8:35 tion principally of the relatively low lift-drag ratio, whose value depends on ground effect and the normal acceleration required to curve the flight path. Weight and density corrections at constant lift coefficient involve only changes in the true approach and touch down speeds, thus changing only the total energy to be dissipated. Consequently, when weight and density corrections are the only ones to be made, the last term of Eg. 8:37 is zero. In practice, the influence of weight and density increments on flight path angle or rate-of- descent is frequently abolished by changing the power setting so that usually AW/W=AF/F. Therefore, weight and density corrections may often be omitted entirely. 8:11 LANDING PERFORMANCE REDUCTION (a) Wind Corrections For the ground run the same correction factor as given in section 8:6 may be applied. However, because the deceleration is generally constant, it may be assumed that the exponent p equals 2 so that The mean deceleration d at the mean speed V of 0.5 (V₁+V½) is = g W 2 Slw + Asi 8:36 Again keeping lift and drag coefficients constant, an equation similar to Eq. 8:9 can be derived, whence As2 = $2 v-vi Δό V²) (AW DO \V−vi + 2gh, :) σ 14-15 wd X V₁ Fw= SI. 8:38 where s₁₁+As gound distance in still air S| W ground distance with head- wind Vw ΔΕ AF AW W (+-) 8:37 V₁ -Vw ground speed at touch down. A headwind of magnitude Vw will shorten the air distance by an amount equal to the product of headwind velocity and the air phase time; i.e., When the net thrust acting during the air phase is low, the mean deceleration is a func- 8:17 As₂ = Vw 12 8:39 Again, the possible wind gradient might be accounted for by using an effective wind velocity in Eq. 8:39 equal to 1.1 times the actual surface headwind velocity (see section 8:6). (b) Slope Correction A positive runway slope B increases the effective deceleration during the ground run so that s₁ + As¡ = FB si where RS is rate-of-descent at obstacle height h above runway plane. (c) Lift Coefficient Corrections Usually it is not advisable to correct for lift coefficient at obstacle height and touch down. Rather, an attempt should be made to keep airspeed close to a predeter- mined value and to avoid excessive float after transition. This is feasible with airplanes having nose wheel or tandem landing gears. and Ad = - gß where ẞ = mean slope of that portion of the runway used for the test (uphill positive). As in section 8:6, we may write Asi SI or Asi $1 5 11 Ad gß 2gs,B v² 2gs, B FB-1 = vi 8:40 No slope correction to the air distance is applied when the obstacle height is related to touch down level. However, when the ob- stacle height is measured perpendicular to the runway plane with slope B, the correction As is (d) Landing Assistance The use of reverse thrust or a brake or drogue parachute requires separate analyses of those portions of the ground run during which the assistance is operative and a separate investigation of the effects of atmos- pheric conditions. Accurate measurement of effective change in deceleration and accurate establishment of the time of application are necessary. 8:12 SUMMARY OF LANDING EQUATIONS (a) Ground Run Measured values: The distances s s₁, V₁, Vw,o, B, W. S¡, V¡, and s2 appearing in the formulae are corrected for the previous effects in the given order. ASI- (1) Wind correction As2 $2 11 V2 RS B 2 -I = Fw - I W W 8:18 (See Fig. 8:1) (2) Slope correction ASI SI 2gs, B 29518 · FB-1 s₁ corrected for Vw. No CL, correction is required. Combination of 1 and 2 gives A52 v-vi /AW AW Ao $2 V2-V₁+2gh W σ SI corrected for wind and slope = Fw・ FB • SI test (when s Si test S₁) rections where (3) Weight, density and thrust cor- Asi SI AW Ao Fg [AW AF W סו | 6 + 251 wd w F S2 corrected for Vw Reduction equations for different landing techniques are given in Ref. 4 and in Chapter 6 of this manual. SI corrected for Vw and F = mean thrust (if any) at ▼ = 0.7 V (b) Air Phase Measured values: S2, VI, V2, Vw, W,σ, t2 rection (1) Wind correction As₂ = Vw t2 (2) Weight, density and thrust cor- None, except for glide approach 8:13 CONCLUDING REMARKS TO PART I In this part of the chapter, we have con- sidered the general problem of take-off and landing performance from the point of view of what might be considered "conven- tional operation" of airplanes. The cases of minimum length take-offs and landings either with or without obstacles in the flight path have not been considered in this chapter. However, the problem of minimum length take-offs is considered in Chapter 6 as well as in Part II of this chapter. Important specialized analyses such as for carrier-type take-offs and landings or the performance of vertical take-off aircraft wherein the stability and control, as well as the performance characteristics of the air- plane, enter the picture, are beyond the scope of this present work. However, it is hoped that such topics will be considered in future volumes. Finally, we note that, although the analysis given here is applicable to turboprop air- planes, as of this time little experience has been obtained with these machines, so that the most satisfactory method for correcting for power variation is not yet known. This is another topic warranting further con- sideration. 8:19 PART II THEORY 8:14 GENERAL PRINCIPLES The take-off maneuver is usually divided into three phases, which are illustrated in Fig. 8:3: (1) The ground run, from the start of roll to the point at which the chosen take- off speed is reached. (2) The transition to climbing flight. (3) The climb to some obstacle. Throughout the ground run the aircraft is assumed to remain close to, though not necessarily in contact with, the runway. The second and third phases often comprise a single maneuver. 8:15 THE GROUND RUN During this phase, the speed increases progressively through a number of important points which require definition. These speeds are defined in the order in which they usually occur on a typical multi-engined aircraft. Control Speed on Speed on the (a) Minimum Control Ground, VMCG This is the speed above which, if sudden failure of any engine occurs, the aircraft being on the ground, the pilot is able to regain control and thereafter maintain a straight path parallel to the one originally intended, without reducing thrust or retrimming. The lateral divergence of the path, and the control forces must be within defined limits. (b) Critical Speed, Vcrit If sudden failure of one engine occurs during the take-off run at this speed, the pilot is able either to stop, or to continue the take-off to the 50-foot point, in the same total distance. If power failure occurs earlier, it is more economical of distance to stop the aircraft, while if it occurs later it is better to continue the take-off. Vcrit must, therefore, be not less than VMCG. (c) The Stalling Speed, Vs This definition is largely self-explanatory. In its estimation no account is taken of the beneficial effects of the proximity of the ground but it is open to question whether or not the effect of engine thrust should be in- cluded. In all other respects the configura- tion is that appropriate to take-off. (d) Minimum Control Speed in the Air, VMCA This is the speed above which, if sudden failure of any engine occurs while the aircraft is away from the influence of the ground, the pilot is able, without reducing thrust or retrimming, to regain control and thereafter maintain straight, steady flight at the same airspeed. The control forces and angular divergence of the aircraft must lie within certain limits. This speed is, of necessity, greater than the stalling speed, though this is not implied by the definition. (e) Take-off Safety Speed, VTOS This is the speed below which the airspeed must not be allowed to fall after leaving the proximity of the ground. It must, therefore, exceed the stalling speed by an appropriate margin, and must, in addition, exceed VMCA by an agreed amount. The relation of these speeds to the various stages of the whole maneuver is shown (purely diagrammatically) in Fig. 8:3. 8:20 TRANSITION DISTANCE SAFETY MARGIN OVER STALL CONTROL SAFETY MARGIN 4 4 YCRIT. VS GETAWAY 8:21 Fig. 8:3 Take-off Maneuver Illustrating Significance of Various Phases and Speeds START GROUND RUN CONTROL SAFETY MARGIN VNCO —ACCELERATION AT FULL POWER STOP, IF ENGINE FAILS BEFORE THIS POINT VMCA VT.O.S CLIMB DISTANCE TRANSITION ENDS XGROUND LEVEL AIRCRAFT CAN TAKE-OFF TO 50 FEET WITH ONE ENGINE "DEAD" OR STOP, IN THIS DISTANCE CONTINUE TAKE-OFF IF ENGINE FAILS AFTER THIS POINT 50 FT The getaway point, i.e., the point at which the aircraft becomes airborne, can occur at any point after the speed Vs has been exceeded, provided that sufficient angle of attack can be achieved with the main wheels on the ground. No appreciable height may be gained, however, until the take-off safety speed, VTOS, has been exceeded. (f) Estimation of Ground-run Distance Consider the aircraft to have reached a speed V ft. per sec, in zero wind at a distance s feet from the start. The aircraft, of weight W lbs., is being accelerated by a forward thrust F lbs. The retarding forces are partly aerodynamic, arising from a total drag coefficient CD and partly frictional, due to the coefficient of rolling friction μ. The coeffi- cients CD and CL are assumed to be held constant throughout the ground run and it is assumed that ground effects have been con- sidered in evaluating these coefficients. 3/0 The equation of motion is therefore: 3 = F - S (W-CL CD · — pV²s - µ (W - CL · — pv²s) where s = forward acceleration, ft. per sec.² = d²s/dt² 2 g = acceleration due to gravity P = air density, slugs per cu. ft. 8:41 off speed, where Fo is the static thrust and k is a constant. W 310 Eq. 8:12 may then be rewritten: ï = (F。 − µW) - — pv²s (CD¯µCL+kFo/{{pS). With the following substitutions, Ŝ = V ૐ : ᏧᏙ I d(v²) ds 2 ds A = 2g (Fo/W-μ) PS 9 15 (CD - μCL + KFO / / / PS) B = 9 W 8:43 we can put Eq. 8:43 into the integrable form: d(v²) ds = A - BV². 8:44 Since we may be dealing with a take-off during which the thrust undergoes discontin- uous changes due to engine failure or the use of rockets, it will be as well to integrate Eq. 8:44 in the general form to give the distance travelled in accelerating from a speed V₁ to a higher speed V2. This distance is then given by I/B S= 109, A – BVi A - BV2/ A - BV² loge A-BV2 8:45 It may be noted that when B tends to zero, the expression for the distance becomes simply (vž – vi) S = A S = wing area, square feet t = time from start, seconds. The thrust T is assumed to decrease with speed according to the relation F = F。 (l-kV²) 8:42 over the speed range from rest to the take- 8:22 8:46 and in this form an approximation may be obtained for the effect of wind. The speeds V₁ and V₂ become ground speeds, and the con- stant A should be the mean acceleration, in ft. per sec. throughout the run. If we split the total drag coefficient Cp into profile and induced drag coefficients so that CD = CDe + KC², 8:47 then by differentiation of Eq. 8:45 with respect to C, we find that the lift coefficient for a minimum value of acceleration distance is given by CL = μ/2K · 8:48* Typical values of μ and K are 0.05 and 0.03 respectively, for a concrete runway and for an aircraft of aspect ratio of about 6, in the presence of the ground. The optimum value of C₁ during the run is, therefore, as high as 0.8, and would be even higher for a take-off from, say, a grass surface or other uneven surface. Shortest take-off runs will, therefore, be achieved where the lift coeffi- cient is brought as quickly as possible to a value approaching the optimum. Although this phase has been referred to as * Differentiation of Eq. 8:45 leads to the re- quirement that dB = 0 dCL together with another extraneous condition involving velocity. Discarding the extraneous condition and evaluating dB dC in terms of the definition of B leads to Eq.8:48. 8:23 the "ground" run, the last part, up to the take-off safety speed, can in fact be made with the wheels clear of the ground. This change may produce a very little difference in the total distance, for throughout this part of the run there may be very little load on the wheels, while the induced drag will be reduced by the favorable effect of the ground. To lift the aircraft clear of the ground at an early stage may therefore have little or no effect on the total ground run, and might even increase it. 8:16. THE TRANSITION AND CLIMB The airborne phase of the take-off maneu- ver has to be considered in two separate parts, or in one single stage, depending on the class of aircraft and the type of take-off required. As the take-off safety speed is reached, it is assumed that the pilot increases angle of attack and thereby applies an increment of normal acceleration, causing the aircraft to start to climb. When the angle of climb reaches the value at which the aircraft is to climb steadily, the lift coefficient is re- duced to the value appropriate to the climbing speed, and steady flight ensues. If the standard 50-foot height has not been attained at this stage, then the airborne phase falls naturally into two parts. If, however, the 50-foot point is passed before steady condi- tions can be achieved, the airborne phase is obviously one continuous maneuver. Our attention is therefore first directed to that part of the maneuver in which the angle of climb is being increased. *If we consider an airplane at any point in the transition flight path between take-off and steady climb, we have the equilibrium of forces shown in Fig. 8:4. * Editor's note: This section to Eq. 8:55 has been rewritten to provide the basis for the transition equations given in the original paper. Summing forces perpendicular and paral- lel to the flight path we have D + ma + W siny - T = 0 · h L = W cos y + CF 8:49 L Flight Path S (CG)- T ma D CF W Fig. 8:4 Forces During Transition 8:24 ハ ​Tangent_To_S_ X 8:50 To solve these equations we presume that initially the airplane is flying at a speed Vg and a lift coefficient CL. The pilot then instantaneously applies a lift coefficient in- crement ACL to start the curved flight path. The transition lift coefficient CL。+ACL is assumed constant during the maneuver as is the difference between the thrust and drag. Under these circumstances Since Va is not too different from Vg, this equation may be linearized by setting Va= Vg in the right-hand member of the equation, whence where dy vå A - B = | ds L = = 1/2 pv² (CL。 + ACL) S where Va is the equivalent speed at the point considered and W = Po POV CLOS where Vg is the equivalent getaway speed. Thus, CLO + ACL A = V& CLO v B go Similarly the second balance equation becomes I d (v²) 2 ds T-D + gy = 00 = 970 E 2 W ~ - (VO) (1 + ACL) The centrifugal force is given by WV2 wv2 CF = wv2 dy gR g dx d2h dx² (assuming y and dy/dx to be small). For small angles s~x and the first balance equation becomes W 11 CF W 4/3 where a。 is initial flight path acceleration ao and Yo is climb angle equivalent of ao or d(va) ds + 2 goy = 290 Y giving us the pair of simultaneous linear differential equations in Va and Y v² - B. dy = 1 A ds d(va) +29σy = 29σYO 29070 ds 8:51 or va v +1 els ACL v² dy dy. CLO ds об 8:25 These equations have solutions of the form and provide the characteristic equation 2go BA YN whence λ = ±¡ 2922 va for ACL << CLO 2 1+ ACL CLO 2 об Since has no real part and since the right-hand terms of the differential equation set are constants, the general solution to our equations may be written va = C₁ sin as + C₂ cos as + K₁ y = C3 sin as+ C4 cos as + K2 Υ where K, and K2 are the particular solutions and a = √2go/V Direct substitution gives 1 K₁ DI- V& CLO CLOTACL This gives K₂ = Yo The initial conditions are: Thus (Valo = Vg lylo = 0 d(va) ds dy ds C لى 11 2go Yo = √2 JI v v² Yo g ACL v CLOTACL ACL C3 = √2CLO CA Yo' v² va V₂ = V₁ {√zy sin (12000) + C + AC gos ACL CLO+ACL COS √2 gos + CLO (12)) CLOT ACL 8:52 gos Yo {1 - cos (√2 90s)} + ACL sin (VZ98) of Y = Yo For small angles we have dh = yds X · ƒ yds = } and n = X {Y。 [ ! - v² - Yo [3- Vo n = 0 S 1% √2 CLO v √2 gos ACL COS + gos va √2CLO ACL. V CLo 290 sin (2003)] + ACL. [ sin (√2000)) COS √2905 8:53 √290 8:54 8:26 By setting Y=Yo in Eq. 8:53, we find that the limiting value of s for transition is given by S = V3 √290 arctan (yo√2 CLOACL). 8:55 If the height gained, given by Eq. 8:54, in the distance s given by Eq. 8:55 exceeds the obstacle height, then Eq. 8:54 would be used to determine the airborne distance with the substitution h equal to obstacle height. If, on the other hand, steady conditions are reached at a height less than that of the obstacle, then we define the transition dis- tance as the difference between the actual distance to some point on the steady climb path and the distance in which this point would have been reached had the aircraft been able to climb straight off the ground at the steady climb angle Yo. Fig. 8:3 illustrates this definition. In this latter case the transition distance, ST, which is not the same as that given by Eq. 8:55 may then be written ST=f. V² /√2.go where the factor f is given by f = sin 8 ACL CLO (1 - cos 8) /√20 8 = arctan (√2% CLO/ACL). 8:56 Fig. 8:5 shows the variation of the factor f with yo for a range of values of ACL/CLO from 0.05 up to 0.8. It is clear that the value of ACL/CLO chosen has a marked effect on the transition distance, and the effect is equally important in Eq. 8:54. We can combine Eqs. 8:52 and 8:54 to give the total airborne distance to obstacle height in the form SA Yo va - va 2go - + h) 8:59 which could, of course, have been obtained quite simply by consideration of the changes in total energy occurring during the maneu- ver, with the assumption that variations in drag are small. Eq. 8:59 does not involve ACL directly but it does require accurate information on the change in speed, which appears as the difference of the squares of two large quanti- ties. This change in speed, given by Eq. 8:52, in turn depends on ACL. Where steady climb conditions are achieved after the obstacle has been passed, a simpler, though less accurate method may be used in estimating the distance to the height h. It is assumed that a mean equivalent lift increment ACL is applied, and that the flight path is an arc of a circle. If the root mean square airspeed over this path is Vm, corre- sponding to steady flight at a lift coefficient CLm, then we have Vi /Rgσ = ACL/CLm 8:60 where R is the radius of curvature of the path. The distance SA to an h of 50 feet is then given by SA 200 W/S PGACLI 8:57 The steady climb distance to h feet is then h/Yo, assuming Yo is small, and the total airborne distance to h feet becomes SA = Sth/Yo 2500 8:58 8:61 8:27 FACTOR "f" 0.9 0.8 0.7 0.6 0.5 0.4 0.3 1.0 من الله 0.05 0.10 0.20 0.30 0.40 0.60 0.80 0.2 0.1 0.1 0.2 X TRANSITION DISTANCE A = +. V 2 2 1√2.80 T V f • = TAKE-OFF SPEED, FT/SEC. Y = LONGITUDINAL ACCN AT TAKE-OFF 'g' UNITS ACE LIFT COEFFICIENT INCREMENT. CLE 0.3 = LIFT COEFFICIENT AT AIRSPEED Vg 0.4 0.5 Fig. 8:5 Effect of Normal Acceleration on Transition Distance 8:28 where W/S is the wing loading. C in Eqs. 8:60 and 8:61 includes the increase in lift arising from any increase in speed above the initial value Vg and is normally higher than ACL used earlier. Flight tests have shown that where the airborne phase is, in fact, a continuous maneuver, irrespective of whether the climb angle at the obstacle height is greater or less than the steady value, Eq. 8:61 is sufficiently accurate for the purpose of calculating the lift increment ACL. 8:17 PILOTING TECHNIQUE IN THE AIRBORNE PHASE From the foregoing it will be clear that the pilot has a far greater influence on this part of the take-off than he has on the ground run, and it is here that we must attempt to be precise in our requirements for pilot technique if repeatable results are to be ob- tained. The important parameter is the lift coef- ficient increment ACL or ACL and for predic- tion purposes we need to know how the maxi- mum increment used by the pilot varies with such factors as speed margin over the stall, acceleration or climb angle, etc. It has been found that a skilled test pilot, asked to achieve the shortest possible distance, can produce consistent results for AC, derived by the method of Eq. 8:61. The total lift coefficient, (CLm+AC), was found to be a linear function of (Vm/Vs)², reaching the expected value of CLmax at (Vm/Vs)² = 1, and falling with increasing speed. The reduction with speed is to be expected since at the higher speeds the time taken to apply the increment is proportionate- ly a larger fraction of the total time to reach 50 feet, thus reducing the mean effective in- crement. There is therefore an optimum speed at which the maximum value of ACL will be applied. Below this speed, proximity to the stall will limit the increment that can be used, while at higher speeds there will not be time to apply and use the available increment, This optimum applies only to the airborne phase, and the total distance from the start to the 50-foot point will be a minimum for take-off speeds much nearer the stall, Some typical flight results are shown in Fig. 8:6. The small scatter of the points, one for each take-off, is indicative of the accuracy with which this particular type of airborne path can be repeated. During the experiments from which these results were obtained, no particular safety limitations were imposed on the pilot apart from his own knowledge of what was a safe maneuver. In normal practice, and particu- larly for civil aircraft, safety considerations will generally limit the lift increment that may be used. In particular, the airspeed must never be allowed to fall below the take- off safety speed and from this it follows that the instantaneous climb angle ought not to exceed the steady climb value. Additional data from take-off measure- ments on civil and military transport air- craft have shown that the lift coefficient in- crements used are about half the values that would give the absolute minimum airborne distance, and hence the airborne distances involved will be roughly 50% greater than the minima. To determine analytically whether a par- ticular lift coefficient increment can safely be held until obstacle height is reached, the distance defined in Eq. 8:55 is substituted in Eq. 8:54 with the value of the height h. The solution of the resultant equation for Yo gives the longitudinal acceleration which will en- sure that the climb angle at the obstacle height is equal to the steady value. If the thrust available gives an accelera- tion less than this limiting value, the steady climb should be started before the obstacle height is reached. For example, at a take- off safety speed of 160 knots, a value of 8:29 8:30 LIFT COEFFICIENT TWIN-JET FIGHTER-MINIMUM TAKE-OFFS AT FULL THROTTLE 1.5 1.0 0.5 201 MAXIMUM LIFT COEFFICIENT TAKE-OFF CONFIGURATION TOTAL LIFT COEFFICIENT USED DURING TRANSITION LIFT COEFFICIENT FOR LEVEL FLIGHT LIFT COEFFICIENT INCREMENT 1.0 1.5 2.0 2.5 3.0 3.5 4.0 (V. Fig. 8:6 Lift Coefficients Used in Airborne Phase NOTE - LIFT COEFFICIENT INCREMENTS WERE CALCULATED ON ASSUMPTION THAT FLIGHT PATH WAS A CIRCULAR ARC VM ROOT MEAN SQUARE AIRSPEED DURING TRANSITION = STALLING SPEED, TAKE-OFF CONFIGURATION off at reduced power, or to close all throttles and come to rest, in the same overall distance. One Analytically, the critical speed can be estimated by the following process. engine is assumed to fail at a range of speeds. Up to this speed the ground run is made at full power; thereafter it is continued at reduced thrust to the obstacle height using the procedure already described. The total dis- tance thus obtained is plotted against the speed at which the engine is assumed to fail, as in Fig. 8:7. The distance required to stop the aircraft following engine failure is also estimated, using the method of Eq. 8:45 with the modifi- cations to the following sections on the esti- mation of ground run during landing. The total accelerate-stop distance is then plotted on the same graph as shown in Fig. 8:7. Two such (dotted) curves are drawn, indicating the effect of different amounts of braking---either wheel brakes or air drag. The points of intersection of these curves with that for the take-off distance give the corresponding critical speeds, and the dis- tances involved. In the the example illustrated, improved braking appears, at first sight, to have little effect on the distance involved, but the im- portant effect is the increase in critical speed. The take-off can be continued following engine failure only if the speed is above the minimum control speed VMCG. If, therefore, the critical speed is below VMCG, the take-off must be abandoned, even though, as shown in Fig. 8:7, a considerably greater distance is required. With improved braking, however, the critical speed may be raised above VMCG giving a useful reduction in the length of runway required. (b) Flight Test Measurements Measurements of the critical speed can usefully be combined with those for the deter- ACL/CL。, held at 0.2 until a 50-foot height is reached would require a longitudinal ac- celeration of about 0.1g to prevent the climb angle from being too steep. At 80 knots, the same increment would require an accelera- tion of 0.26g, so that light aircraft will nor- mally perform the airborne maneuver in two parts, for lack of sufficient acceleration. Similarly, using the distance given by Eq. 8:59 for no change in airspeed in Eq. 8:54 with h = 50 feet, we may determine the value of Yo for no drop in speed at a 50-foot point. Again, at a value of ACL/CL。 of 0.2, at 160 knots take-off speed an acceleration of 0.05g is required while at 80 knots, 0.1g is neces- sary. The requirement for no loss of air- speed is therefore easier to meet than that for no excessive climb angle. 8:18 APPLICATION TO FLIGHT TEST PROCEDURE The total distance to obstacle height may be measured under what are considered to be normal operating conditions, but it is often difficult to produce consistent results, par- ticularly for this airborne phase. It may be more reliable, therefore, to request the pilot to produce the shortest possible airborne dis- tance consistent with safety and hence to calculate the lift coefficient increment ACL by the method of Eq. 8:61 for high perform- ance aircraft, or Eqs. 8:56, 8:57, and 8:58 when the available thrust is relatively low. Then, using, say, half this value of ACL, the length of the airborne path can be cal- culated either as a single maneuver, if the acceleration is adequate, or in two separate parts if it is not. At least 3, preferably 4, take-offs should be made in each case, and the average of the best two used for reduction purposes. 8:19 CRITICAL SPEED DURING TAKE-OFF (a) Estimation If, on a multi-engined aircraft, one engine fails during take-off, at the critical speed, the pilot is able either to continue the take- 8:31 3,000 2,500 2,000 TOTAL TAKE-OFF DISTANCE TO 50FT. | TOTAL DISTANCE, YARDS 1,500 NORMAL DISTANCE TO 50 FT. (ALL ENGINES OPERATING) 1,000 ACCELERATE - STOP DISTANCE NORMAL BRAKES 500 25 50 ACCELERATE-STOP DISTANCE WITH IMPROVED BRAKING 75 CRITICAL SPEED NORMAL BRAKES YMCG TAKE-OFF SAFETY SPEED IMPROVED BRAKING | CRITICAL SPEED 100 125 150 SPEED, KNOTS, AT INSTANT OF ENGINE FAILURE Fig. 8:7 Critical Speed During Take-Off mination of the minimum control speed on the ground, VMCG. The additional complication is mainly in the ground equipment, rather than in the pilot's task. Sudden engine failure should be simulated as closely as possible, starting at a fairly high speed, the take-off being continued in each case, until a failure is simulated at a 8:32 speed at which the pilot is not able to keep the angular and lateral divergences of the aircraft within prescribed limits. The total distance obstacle height is obtained in each. case, and plotted against the speed at failure. Similarly, a series of abandoned take-offs should be made at gradually increasing failure speeds, the accelerate-stop distances being measured and plotted against the speed at failure. Then, even though the minimum control speed VMCG may be higher than the critical speed, extrapolation of the two curves towards each other will enable an intersection to be obtained, from which the critical speed may be derived. The greater the length of runway available for the simulation of aban- doned take-offs, the less will be the amount of extrapolation required. 8:20 THE LANDING MANEUVER The landing maneuver is, in many ways, the exact opposite of the take-off, not the least difference being the difficulty of achiev- ing precise, consistent results when a quan- titative study is undertaken. It may therefore be profitable to discuss in general terms, the reasons for this lack of precision, although it is obvious that in this case the pilot's skill and judgment plays a much larger part than it does during a take-off. The landing path can be divided into several phases as in the case of the take-off. Fig. 8:8 illustrates these phases which are defined as follows: (a) The approach, at a steady rate of descent which must, however, be reduced before touch down. The approach distance is that length of this steady path from obstacle altitude to the point where that path would intersect the ground. (b) The flare, started at some height above the gound, by an increase in angle of attack, with the intention of reducing the vertical velocity to as near zero as possible by the time ground level is reached. (c) The float, which follows the flare if zero vertical velocity is achieved before the wheels touch the gound. During this phase, speed decreases until the aircraft attains an angle of attack at which a touch down can be made. With a nose-wheel type undercarriage, the aircraft can usually touch down at the end of the flare, without floating. (d) The ground run, from the point of touch down until the aircraft comes to rest. 8:21 ESTIMATION OF LANDING DISTANCE It is usual to calculate (and to measure) the landing distance from the point where the aircraft is at some altitude h above the runway to the point where the aircraft comes to rest. While this procedure has the advantage of precision, it does not indicate the length of runway needed for a safe landing. It may be argued that, for practical purposes, the important distance is the length of runway used in reducing the aircraft speed to a value at which it may be said to be taxying. The starting point is therefore at the down-wind end of the runway, which may be at a height which can vary widely from one landing to the next, depending on the pilot's skill and judgment. The calculation will, in any case, follow closely the procedure used for the take-off distance, and the normal practice is to per- form the calculation for a starting height of 50 feet, and to multiply the result by a factor to allow for the inevitable inaccuracies in height and glide path angle in crossing the end of the runway. (a) The Approach Distance If his the height at the start of the approach path (normally 50 ft.) and y is the steady glide angle, then the approach distance is simply h/y when Y is a small angle, in radians. (b) The Flare Distance The comprehensive analysis of this portion of the maneuver has recently been published by Doenhoff and Jones (Ref. 18). In this paper, a study is made of the ideal landing flare. The aircraft is assumed to 8:33 START OF HOLD-OFF 8:34 h = 50 FT. APPROACH FLARE FLOAT GROUND RUN STOP START OF FLOAT TOUCH-DOWN AIM POINT OF INITIAL GLIDE Fig. 8:8 The Landing Maneuver start the flare at an optimum speed such that at the end of the flare the minimum possible distance has been covered and the speed has fallen to the stalling speed, with the flight path tangential to the runway surface. Because of the reduction in speed during the flare, there is therefore a minimum speed at the start if the flare is to end with the speed equal to the stalling speed. This effect has been discussed by Meredith (Ref. 19). Doenhoff and Jones consider the effects of the inevitable errors in the pilot's judgment of his height, position and speed on the approach, with reference to the length of the flare, and conclude that the most effective way of reducing the length of runway required would be to provide some means of fixing the point of the start of the approach glide. This point is commented upon later. An elementary analysis of the flare maneu- ver will serve to show the source of the variations in this distance in practice. We assume the flare to start from a steady glide at a speed Vo ft. per sec. and a glide angle Y radians. Assuming Y to be small, as usual, the rate of descent is therefore VY ft. per sec. The flare is accomplished by the use of a lift coefficient increment ▲С above the value CL required for steady flight at the speed Vo. A normal accelera- tion of g (ACL/CLO) ft. per sec? is therefore produced, and, to a first approximation, a circular flight path results. The vertical velocity VY is thus reduced to zero in a time tf seconds, where tf = VYCLo /gACL (approx.). The height lost in this time, which is therefore the height at which the hold-off should be started, is hf, where hf = VZt+ f = V²y²CLo/2gACL; f i.e., hf 11 (x² W) 19 /gp ACL feet where W/S is the wing loading, in lbs. per sq. ft. If the flare path is assumed to be a circular arc, then the flare distance, as defined in Fig. 8:8, is half the total distance occupied by the flare maneuver. Thus the flare distance, Sf, is given by Vit = = Sf= 2 v²yCLo/2gACL YW (TW)/9PACL feet. S Throughout this simple analysis we have ignored the loss in speed during the flare, and assumed that the initial speed is high enough for the lift coefficient increment ACL to be applied and maintained in safety. Actual- ly the speed loss is of the order of 5%. At a given approach speed, therefore, a strictly limited value of ACL is available. The flare distance is then proportional to the glide angle at the start of the flare, while the height at which the hold-off must be started varies as the square of the glide angle. Since the pilot is unlikely to make use of the full lift coefficient increment, for fear of stalling while still well above the runway, there inevitably will be variations in ACL which will appear as further variations in the flare distance. Although the airspeed does not appear directly in the above expressions for flare distance and hold-off height, it does, of course, affect directly the value of ACL to be used. The normal approach airspeed will be dictated by engine-cut safety requirements as well as by the requirement for an adequate margin over the stall for maneuvering, gusts, 8:35 etc. The approach airspeed will generally be near, if not actually less than, the minimum drag speed. An inadvertent loss in airspeed might therefore result in an excessive glide angle close to the ground, and at the same time reduce the lift increment that can be used in safety. In these circumstances, if the flare is to be completed in order to avoid damage to the undercarriage, the hold-off must be started at a much greater height than usual, and will take a correspondingly greater distance to complete. Piloting skill, as well as the aerodynamic characteristics of the aircraft, both play an important part in fixing the distance occupied by the flare and the type of touch down which ends it. (c) The Float and Ground Run An aircraft with a nose-wheel type under- carriage is able, by virtue of its relatively low ground angle of attack, to touch down at speeds well above the stall, and the float is eliminated. However, whether the actual touch down is made immediately at the end of the flare, or later, the distance involved in bringing the aircraft to rest, or reducing its speed to the taxying speed can be estimated by a method identical with that used for the ground during take-off. Eq. 8:45 gives the distance x required to change the speed from V₁ to V2 ft. per sec. in the form X = loge B A - BV12 A - BV₂² 2 B = aps (CD - HCL + KTO / / Ps). W To For landing run analysis we have A = 2g (To -H) 8:36 The thrust To normally becomes the idling thrust of the engines with throttles closed; if reversed thrust is produced, To will become negative. The coefficient of friction μ is now due to the braking action and must, of necessity, be arithmetically greater than To/W if To is positive. The constant A is therefore negative during the landing ground run. For the constant B, the lift and drag coef- ficients C and Cp are, of course, the appro- priate values considering ground effect and are again assumed constant throughout each stage of the run. The constant k accounts for the variation of the idling or reversed thrust with airspeed and will differ from that used for the take-off estimate. The drag coefficient CD will include the effect of such drag producing devices as parachutes or air brakes, if used. Since the initial speed V₁ will exceed the final speed V2, while B is normally positive (though not necessarily so, if the drag is low while the wheel brakes are particularly effective), the distance x will always be posi- tive and the expression can be used without modification. Obviously, if the run from the end of the flare to rest is made in stages over which the means of of deceleration are changed discontinuously (for example, by release of a parachute, or application of wheel brakes, or a change from a true float to a run along the ground) then any estimation of distance using the above expression would have to be made in similar stages. A major difficulty in making reliable esti- mates of this part of the landing run is in the choice of an appropriate value for the friction coefficient μ. Reliable information is very meager, but such as it is, it confirms that μ varies with speed, and that at high ground speeds very much reduced values of μ are appropriate. Where pilot-operated brakes are used, the pilot is unlikely to use much brake at the start of the run, for fear of producing a dangerous skid, and he may do little more than give an occasional "jab" at the brakes until the speed has fallen, and the wheel load increased to a point where the wheels will not skid. Even if the wheels have automatic brakes, the ground run must, therefore, be divided into two parts. Over the first part, a coeffi- cient of friction should be used which is between the unbraked value (usually 0.05) and the maximum attainable value. For the second part, full braking action is assumed and the braking force becomes constant, with slip occurring in the brakes themselves. The speed at which full brakes can be used may be estimated graphically on a diagram on which the friction coefficient μ is plotted against ground speed V, as in Fig. 8:9. The available value of μ is shown as a full line, while the value of μ required to prevent skid- ding is derived from the known maximum re- tarding force which the brakes can apply, divided by the instantaneous load on the braked wheels. In the example shown in Fig. 8:9, from the touch down speed down to about 80 knots, danger of skidding exists unless automatic brakes are used. Without such brakes, therefore, the usual technique is for the pilot to hold the nose wheel up and to rely on air drag until the brakes can be put full on, when the angle of attack is reduced to a min- imum to increase the wheel load. By differentiation of the expression for the deceleration distance it can be shown that this distance is a maximum when the lift coefficient used is equal to μ/2K, the value K being the induced drag factor (CD CD+K C²). = No compromise between the use of wheel brakes and air drag is therefore practicable, and the pilot should concentrate on making the maximum use of one or the other, but not both. 8:22 THE CHOICE OF THE LANDING APPROACH AIRSPEED An essential part of any landing tests or measurements is the choice of the minimum approach airspeed. Since this depends crit- ically on the stalling speed, it is assumed that this speed, or the minimum practicable flight speed if the aircraft has an unconven- tional stall, will have been established before- hand. The maximum lift coefficient increment to be used for maneuvering must still leave a margin over the maximum determined by the preliminary tests. The maneuvering requirements refer to corrections to the flight path in both the horizontal and vertical planes. The magnitude of these corrections depends mainly on the pilot's ability to detect errors in position and rate of change of position. We should expect that if the visual or other form of guidance is improved, not only will the accuracy of the final arrival be increased, but also that the magnitude of the corrections which the pilot applies will be reduced, although he may make such corrections more frequently. Maneuverability is, of course, not the only criterion on which the pilot will assess the minimum approach airspeed, and he will be influenced by the controllability of the aircraft as regards attitude, forward speed and rate of sink. The speed must also be above the minimum at which full throttle can be applied, with one engine inoperative, for an emergency overshoot. Control of the attitude of the aircraft, e.g., correction for wing dropping, will de- teriorate as the speed is reduced due to the more sluggish response to control move- ments. Further, on swept-wing aircraft the lateral stability derivatives change markedly with lift coefficient, and dutch-rolling may occur. If the aspect ratio is low, the attitude 8:37 8:38 FRICTION COEFFICIENT M 0.8 0.6 0.4 0.2 ނ AVAILABLE (VERY APPROXIMATE, ESPECIALLY ABOVE 80 KNOTS } 1 SPEED AT WHICH FULL BRAKES CAN BE APPLIED FB REQUIRED = WHERE: W-L FB W L = MAXIMUM BRAKING FORCE, LB. = AIRCRAFT WEIGHT, LB. = WING LIFT AT APPROPRIATE SPEED, LB. ROLLING FRICTION - NO_BRAKES 0.0 50 60 70 80 GROUND 90 RUNNING 100 110 SPEED, KNOTS Fig. 8:9 Use of Wheel Brakes During Ground Run 120 130 140 of the aircraft may become excessively nose- up making the view inadequate. Control of airspeed and rate of sink depend on the relation of the airspeed to the minimum drag speed, VMD. If the airspeed is above the minimum drag speed, the pilot is able to make a correction to his height by elevator movement only, and a glide path parallel to the original can be flown, at the original airspeed, without change in thrust. If the airspeed is below VMD, however, then a form of instability exists, in that if the airspeed changes in the course of making some correction, then that change will in- crease with time unless a change in thrust is made. Flight at speeds below VMD is pos- sible---deck landing approaches are fre- quently made at speeds in this region---but the maintenance of a steady approach requires constant adjustment of the power setting as well as the elevator. It is clear, therefore, that the minimum approach airspeed is not amenable to calcula- tion, and one can only indicate the general principles which govern the pilot in his choice. 8:23 SOURCES OF VARIATION IN TOTAL LANDING DISTANCE Of all flight test experiments, the measurement of total landing distance is probably the least precise. Even when all the test conditions are accounted for in the reduction of the results, large residual varia- tions must be expected, and a generous safety factor has to be applied in determining the minimum safe runway length. The fundamen- tal inability of the pilot to follow the ideal path in space and time is the main reason for this inconsistency. While part of this residual variation arises from the variation in the use of the brakes during the ground run, automatic brakes now allow the maximum use to be made of the brakes irrespective of piloting skill. Addi- tional decelerating devices, such as brake parachutes, further reduce the length of the ground run. Nevertheless, the point on the runway at which the aircraft is brought to rest will continue to be ill-defined until the pilot is given additional assistance in starting his approach glide from the optimum point in space. This assistance can, in fact, be given in a simple form, and need involve no additional airborne equipment. The starting point for the ideal glide will, presumably, be on the ex- tended center line of the runway, and this is fairly easy to indicate to the pilot except in bad visibility, when radio aids may be re- quired. In VFR conditions, however, markers indicating the position of the runway center line would suffice. In the vertical plane the pilot needs an indication of his position relative to the ideal path, and this can be provided, in VFR con- ditions, by two markers alongside the runway at the aim point of the ideal glide. These markers, separated by 150-200 feet and at different levels, are arranged to lie on, and thus to define, the path which the pilot's eye should follow. He then has a direct indication of his height error above or below the ideal path. This is essentially a short-range device, covering the final mile or less of the approach. Tests have shown that this form of indication is readily appreciated by the pilot, and con- sistent flight paths can be flown. The hold-off and flare will still be required unless the chosen glide angle is so shallow that the undercarriage can safely absorb the vertical component of the velocity. With simple equipment of this sort, the variations in height and glide angle at the edge of the airfield are appreciably reduced, and more consistent total landing distances will result. 8:39 REFERENCES 1. Lush, K. J., "Standardization of Take-off Performance Measurements for Aeroplanes," USAF Technical Note R-12, 1952. 2. "Review of Performance Techniques," A. & A.E.E. Discussion Memo 2 5821/PAH, 1952. 3. Herrington, R. M. and Schoemacher, P. E., "Flight Test Engineering Manual,” USAF Technical Report 6273, 1951. 4. Royal Aeronautical Society Data Sheets, RC 2/1 & 2, 1950. 5. Perkins, C. D. and Hage, R. E., “Airplane Performance, Stability and Control," John Wiley and Son, 1949. 6. Hamlin, Benson, "Flight Testing," MacMillan, 1946. 7. Hartman, E. P., "Considerations of the Take-off Problem," NACA TN 557, 1936. 8. Diehl, W.S., "The Calculation of the Take-off Run,” NACA TR 450, 1932. 9. John, G., “A Further Development in Calculating the Take-off to 50 Feet Distance of an Aeroplane," Aircraft Engineering, April, 1948. 10. Gates, S. B., "Notes on a Method of Analysis of Ground Run During Take-off," R. & M. 1820, 1937. 11. Wetmore, J. W., "The Transition Phase in the Take-off of an Airplane," NACA TR 626, 1938. 12. Ewans, J. R., Hufton, P. A., "Note on a Method of Calculating Take-off Distance," A. & A.E.E. B.A. Dep. Note Performance 20. 13. Garbell, M. A. and Young, W. M., "The Ground Run of Aircraft in Landing and Take-off," Garbell Aeron. Series 3, 1951. 14. Lovell, J. C., Lipson, S., "An Analysis of the Effect of Lift-drag Ratio and Stalling Speed on Landing Flare Characteristics," NACA TN 1930, 1949. 15. Ribnitz, W., "Umrechnungsverfahren fur Start und Landewege," GDC 10/12022, D. V. L. TB 3, 1939. 8:40 16. Multhopp, "The Problem of the Shortest Take-off Run," Translation GDC 16/57 T. 17. Quinn, J. J. & Lush, K. J., "Take-off Turbine-Jet-Propelled Aircraft, an Investigation into the Technique of Measurement and of Reduction to Standard Conditions," A. & A.E.E. Res/232, 1946. 18. Doenhoff and Jones, "An Analysis of the Power-off Landing Maneuver in Terms of the Capabilities of the Pilot and the Aerodynamic Characteristics of the Airplane," NACA Tech. Note 2967, August, 1953. 19. Meredith, "Note on the Minimum Speed from which the Direction of a Gliding Airplane can be Changed to a Horizontal Path for Landing," R. & M., No. 993, British A.R.C., 1925. 8:41 AGARD FLIGHT TEST MANUAL VOLUME I, CHAPTER 9 SPECIAL TESTS By D. 0. Dommasch Princeton University Section 9:1 W. E. Gray Royal Aeronautical Establishment United Kingdom Sections 9:2, 9:3, 9:4, and 9:5 J. Idrac Centre d'Essais en Vol, France Section 9:6 M. Guenod Centre d'Essais en Vol, France Section 9:7 J. Foch Centre d'Essais en Vol, France Section 9:8 ( VOLUME I, CHAPTER 9 CHAPTER CONTENTS Page TERMINOLOGY 9:1 INTRODUCTORY COMMENTS 9:1 9:2 INTRODUCTORY COMMENTS ON AIR FLOW VISUALIZATION 9:1 9:3 METHODS OF FLOW INVESTIGATION 9:2 9:4 FLIGHT TESTING 9:8 9:5 OTHER TECHNIQUES 9:10 9:6 CALIBRATION OF AN ANGLE OF ATTACK MEASURING SYSTEM 9:11 9:7 MEASUREMENT OF SIDESLIP ANGLE 9:12 9:8 AIRBRAKE EVALUATION 9:14 REFERENCES 9:21 ) TERMINOLOGY R. Angle of Attack CL Lift Coefficient F Engine Thrust m Airplane Mass 8 Gravitational Acceleration V True Flight Speed Vcal Calibrated Airspeed P Air Density 9 Dynamic Pressure M Mach Number Со Drag Coefficient ACO Drag Coefficient Increment Y Flight Path Angle with Horizontal 9c Compressible Dynamic Pressure Pi Indicated Pressure Difference B Sideslip Angle 9:1 INTRODUCTORY COMMENTS problems. Measurements of pressures and velocities have always been among the chief methods of study, but as provisions are seldom made in a given airplane design for such measurements, the long established method of studying the airflow visually with light streamers has been much used in flight test work. It is useful when conducting stability and control tests to know the values of the side- slip angle and the angle of attack with some degree of precision. Knowledge of these parameters is also of importance when evaluating the performance of the airplane armament as well as take-off and landing characteristics under critical operating con- ditions (such as carrier take-offs). For this reason several types of angle of attack and sideslip angle measuring devices are in present use, and in general, flight calibra- tion of these is required before dependence may be placed on their scale readings. The flight calibration of equipment such as this is classified here under the broad heading of "special tests. The use of streamers, or wool tufts, has been invaluable in studying the stalling of wings, and in showing where the flow at high speed breaks down or separates at junctions or behind shock waves and for these purposes it is still widely used. In addition to calibrations such as these, we also must conduct special tests to deter- mine the suitability and effectiveness of such items as dive brakes or dive flaps particularly in regard to the let-down char- acteristics of very clean jet aircraft. The steady increase in flight speeds and improved aerodynamic design cleanness has increased emphasis on skin friction drag, and has led to a closer study of boundary layer flow. While this flow can be studied by measuring pressures and velocities, these methods are clumsy when applied to a large area and their use at only one location leaves room for errors due to local surface imper- fections. These latter may change the bound- ary flow from the laminar to the turbulent type, with a six-fold to ten-fold increase in drag. For research work and for developmental flight testing, it is sometimes mandatory to investigate the actual local airflow conditions existing about a given airplane to establish the causes of buffeting, control malfunction- ing, flow breakaway and the like. The topics listed above, although they bear little direct relation to one another, have all been grouped together in this chapter under the heading of special tests. There are, of course, a number of other special tests which are conducted, and it is hoped in the future that this chapter will be expanded to include additional topics. It is for these reasons that the visual methods of studying the boundary layer over whole areas, devised in Britain and used in flight testing there for the past ten years, have largely superseded the older methods. The visual methods all rely on the basic difference between laminar and turbulent flow, viz., that the turbulence is accompanied inevitably by a rapid vertical interchange of air particles throughout the boundary layer and so to the surface. 9:2 INTRODUCTORY COMMENTS ON AIRFLOW VISUALIZATION Whether the action used is a chemical one or a physical evaporative one is a choice dictated by the test conditions that have to be met. In either case only a very modest amount of preparation is needed on the air- craft that is to be tested to change the sur- face finish on the test areas. This prepara- tion usually takes from one or two hours to one or wo days depending on the areas A clear understanding of what the air is doing as it passes over the surfaces of an aircraft is always helpful and sometimes vital in the investigation of aerodynamic 9:1 involved. These methods will greatly speed the work of the scientist. The present chap- ter will be directed mainly to a description of them. of stalling and of local breakaway of flow; it cannot be used to show turbulence in the boundary layer flow where the size of the disturbances is of a totally different order. (b) Chemical Methods A study of shock waves in high speed flight as distinct from the breakaway of flow due to their action has recently been carried out. The methods employed for this have made use of the change in the refractive index of the air at the shock wave as in the older wind- tunnel techniques. These methods have al- ready been widely described and will be dealt with briefly here. These methods (Ref. 1) of studying the boundary layer flow make use of the turbu- lence itself to indicate the turbulent areas. To do this, the wing surface is sensitized with a material that changes color on expo- sure to an active gas or to impurities in the atmosphere. The best results have been ob- tained on a white surface, which is sprayed before flight with a colorless sensitizing solution, the composition of which is chcsen to suit the concentration of gas to be used or likely to be encountered. 9:3 METHODS OF FLOW INVESTIGATION (a) Tufting This well-known method indicates the local direction of, and the directional steadi- ness of the flow and to a slight extent its velocity. It can be used to show the flow both on and at a distance from the surface and it relies on the relative lightness and flexibility of the material used. The passage of the aircraft through the air brings much more gas or impurity into contact with the surface where the flow is turbulent and so causes a darkening in these regions, the areas of laminar flow remaining white. The only active gas that has been used is chlorine in a concentration of about five parts per million. This is safe for short exposures and a flight of a few seconds through it produces a picture. If the material is too heavy, the tufts or streamers can give a false indication of un- steady flow by the oscillations they can ex- hibit in a steady stream, a feature that must always be guarded against. When the tufts are mounted on masts, precautions must be taken to avoid their getting entangled with the supports during violently unsteady flow. The result of taking such precautions may well be excessive tuft weight and stiffness and this also must be guarded against. An example of such a test result is given in Fig. 9:1. The sensitizing solution is com- posed of starch, potassium iodide, sodium thiosulphate and water, and an atmospheric relative humidity of over 50% is needed to give a reaction. The proportions used are as follows: Starch 1.0 gm. Potassium iodide 3.0 gms. Sodium thiosulphate 0.5 gm. Water 100 cc Properly arranged, however, tufts have the great merit of showing the flow behavior under changing flight conditions, during which time they can be photographed continuously. The chief use of this method is in the study The solution is sprayed on the surface as a fine "mist," the tiny droplets never being allowed to merge into a wet film. To avoid 9:2 E Fig. 9:1 Transition at 60% Chord on “King Cobra” Wing. Chemical Method at Low Altitude. this happening, several sprayings at short intervals are made until a total of three or four cubic centimeters have been sprayed on each square foot of surface. Two methods of producing a chlorinated atmosphere have been used: (1) by discharging gas into a high factory chimney and flying through the smoke from it, and The action of the chlorine or the atmo- spheric impurities is to oxidize the KI, which discolors the starch. The function of the sodium is to absorb the oxidizing agent to the limit of its capacity and so give a chemical "threshold," thus producing a better contrast between turbulent and laminar areas. (2) by releasing liquid chlorine from an aircraft simultaneously with smoke to mark its location; the liquid immediately vaporizes into a gas. The second method permits tests to be done at altitude although it has not been used at more than 7000 feet. For oxidation by atmospheric impurities, the sodium content is reduced by one-half to keep the flying time within reasonable bounds. Even so, the use of a town polluted atmosphere is almost essential. (c) Evaporative Methods (Dry and Wet) These methods were developed from the 9:3 chemical method, and rely on the differing rates of evaporation (instead of chemical action) in the laminar and turbulent areas of boundary layer flow. black. It becomes white again as it dries and this happens first where the boundary layer is turbulent. The dry or sublimation method (Ref. 2) uses the evaporative properties of slightly volatile solids to give a picture of the flow. The solid is chosen to suit the test conditions and is dissolved in a solvent that does not affect the wing surface. The surface should be black and it is sprayed with a solution in what is best described as a just-wet state, the solvent evaporating within a second or two of reaching the surface. When the method (Ref. 3) was first devel- oped it was thought that the wetting liquid had to have approximately the same refractive index as the white coating to get good results. Published descriptions of the method have named several liquids as suitable, most of which, if not all, must be sprayed on as they soften the surface. This is a serious dis- advantage in flight testing owing to the ad- herence of grit and dust. The coating thus formed has a "frosted" appearance and can be seen and photographed fairly easily. It is, however, "aerodynamic- ally rough" owing to its crystalline nature and it must be lightly rubbed down before flight. This is best done with the palm of the hand. An example of a test result using this method is given in Fig. 9:2. It has since been found that the refractive index can be very appreciably different and still give black and white pictures. This permits kerosene to be used in conjunction with a coating of china clay in diluted aero- plane dope* and allows it to be wiped on just before flight and any surplus removed. Al- though it is more likely to collect grit than the dry solid coatings, the “wer" surface can be dusted down just before take-off. Owing to the dry nature of the coating this method has the advantage for flight work of being less likely to collect dust and grit be- fore take-off and that a final removal of any grit is possible just before flight. It has the disadvantage of being very sensitive to changes of temperature, a factor which limits its use to low level testing. The variation of evaporative rate of kero- sene with temperature is somewhat less sensitive than with the volatile solids so that when this method is used and the surface wetted on the ground, tests can be made at a somewhat greater altitude. The solids that have proved most suitable have been napthalene in cold weather, ace- napthene, and azo-benzene under very hot or high-speed conditions. Petroleum ethers with boiling points between 80°C and 120°C C have proved to be suitable solvents, and a 5% solution (W/V) is convenient. A recent development of this method for use at very high altitudes employs what has been called a "wetting at altitude" technique. This overcomes the otherwise inherent diffi- culty of applying a coating to a wing in warm conditions at low altitudes and expecting it to evaporate in the intense cold of the upper air. The wetting liquid is therefore emitted as a spray from a second aircraft and the test wing is wetted by this spray at the operating height by flying close behind the tanker aircraft. (d) The Wet or "China Clay'' Method This method employs the evaporative properties of liquids and uses a porous white coating on the wing over a black undercoat. This has the advantage of giving pictures in black and white, for the white surface becomes transparent when wet and therefore appears *200 gm. fine kaolin, 250 cc clear dope and 750 cc thinners. 9:4 Fig. 9:2 Transition on a "Vampire" Wing. Sublimation (Dry Evaporation) Method, at Low Altitude. 9:5 This development gives complete freedom of choice of liquid, and as the picture dries out on the test wing it is photographed by the tanker aircraft. Several tests at various speeds can be done on one flight even with changes of altitude. An example of such a test result is given in Fig. 9:3. where flow was laminar during the first test are still available to record a new "front," dried out during the second test. The first result, of course, would be photographed in flight. This procedure has not yet been demonstrated as no case has arisen that offered a possible saving of effort. (e) Deposition Method This method is the only visual one for the study of the boundary layer that enables a series of tests to be carried out during one flight, although in the cases of the dry evapo- ration and the foregoing "wetted before flight” method, it may be possible to get two results per flight. light. This should be pos if the transition “front' is known to move forward on both the top and bottom surfaces with change of test speed. The wet areas The deposit of smoke particles in regions of turbulent boundary layer flow has been used occasionally in ground tests on rotating airscrews. The deposition is due to the velocity normal to the surface given to the particles by the turbulence enabling them to adhere by impact. The effect was first Fig. 9:3 Transition at 50% Chord on''Vampire" Wing. M = 0.7 at 35,000 feet. Wetting at Altitude (Wet Evaporation) Method. 9:6 noticed in high-speed tunnel tests where city air was being drawn over models and later on airscrews used in polluted atmo- spheres and even on a glider after prolonged towing. Some tests have been made to assess the practicability of the method as a flight technique, for the action is mechanical in its nature and therefore unaffected by tempera- ture or humidity to any extent. The conclu- sion reached was that the amount of smoke required would be excessive and piloting difficulties of seeing well enough to keep in dense smoke might arise. Moreover, only one test per flight could be made, probably involving two aircraft. Therefore, the method has not been pursued further. density at a shock wave. Both employ light passing in a spanwise or more or less span- wise direction near the wing surface. One of the methods is virtually the same as has been used for some years in wind tunnels with a refinement for indicating the spanwise position of the shock waves re- corded. It gives a photograph of the chord- wise position and inclination to the surface of a shock wave and the distance it extends from the wing surface. The field covered in one test, however, is rather limited both in depth and in chordwise extent. The other method records on a moving film the chordwise and spanwise positions of shock waves at a chosen distance from the wing surface. It is much less limited in a chordwise direction, but does not indicate depth or slope of the waves. A full descrip- tion of these techniques has already been made public (Ref. 4). (f) Optical Techniques These have been used successfully in flight to get photographic records of shock waves and their positions. Two techniques have been used, both making use of the re- fractive effect of the sudden change in air An example of the first method is shown in Fig. 9:4 with shock waves extending almost to the wing surface. AleFLOW Hlasa TunZIY S2 WING SURFACE 1 Fig. 9:4 Shock Waves at M = 0.77, Extending Almost to Wing Surface (the Several Waves are at Different Spanwise Position). Actual Depth of Photograph is 2.5 Inches. 9:7 9:4 FLIGHT TESTING The application of the foregoing methods of air flow visualization to flight test work is covered in this section with special reference to the limitations, difficulties, and choice of techniques that have been en- countered or that might arise in pushing the methods to their limits. (a) Tufts to Show Flow Separation cumstances may often dictate the choice, as may also the extent of the areas to be tested, If the location is windy and sandy, then the dry evaporation method is best because there is less contamination by grit. If tests have to be done in a fly-laden atmosphere, then again the dry evaporation is best because the surface can most easily be protected before take-off (see par. 9:5(c)). The operating factor of elapsed time from preparation of the test surface till take-off can be important in hot sunny conditions as the black surfaces needed for dry evapora- tion absorb heat rapidly and the coating evap- orates from the upper surface. Hangar protection is required for all spraying if there is more than a light wind, otherwise the uniformity of coating suffers too much. A local wind-screen may be used for dry evaporative coatings but hangar coverage is nearly always required for chemical method spraying so as to avoid dust while the spray is drying. (1) Stalling Experience in the use of this method is widespread. It is perhaps only necessary to emphasize here the need to avoid the use of material that is too heavy for the job in hand. The risk is that the tufts may appear to show unsteady flow which is in fact not there. Some compromise between suitability and dura- bility is usually required but the choice should lean towards a fleeting truth rather than an enduring untruth and the bother of renewing tufts periodically or even frequently should be faced. Tuft behavior is usually recorded satis- factorily by one or more fairly high-speed motion picture cameras mounted on the test aircraft. This may be supplemented by the pilot's impressions, when obtainable. The wet evaporation method has the ad- vantage of easy application under windy con- ditions immediately before take-off (if nec- essary). This lessens the risk of collecting dust on the surface which would otherwise require a final wiping of the test surface. The chemical method coating cannot be wiped over after application but it can usually be omitted from the forward part of a wing. a (2) High Speed Separation The observations on stalling again apply, but the durability problem is more acute. Nylon tufts usually stand up well to the tremendous battering imposed on them in a breakaway region at high speed. An important factor in the use of the evap- oration methods for low altitude tests is that the duration of the flying at the test speed or condition must be quite considerably longer than the duration of the unsteady conditions of take-off, climb, acceleration, etc. If the test speed is very high this problem is eased slightly since the stagnation temperature (which is a major factor) rises rapidly with speed. (b) Boundary Layer Observations (Chemical and Evaporative Methods) (1) Low Altitude Steady Conditions (2) High Altitude, Steady Conditions Here the choice is wide, as all the methods will work under these conditions. Local cir- Here there is no choice; the only possible alternative to wet evaporation is a modified 9:8 1 chemical method which permits it to be used in almost dry air; this has only been tested in a laboratory (see par. 9:5(b)). The wet evaporation must be of the "wetted at altitude" kind, since no liquid capable of evaporating at 30,000 or 40,000 feet in any acceptable time would remain unevaporated during the climb. Evaporated solids at 30,000 feet disperse at about 1/200 of the ground-level rate and, although liquids are appreciably better, the falling off in rate with height is still prohibitive. If both aircraft can fly at the test speed required and even wet at that speed, then a fairly quick drying liquid is used and a series of pictures of the drying is recorded. Several tests can thus be done in rapid succession. While the liquid is drying and the surface is changing from black to white in the turbulent regions, the film of liquid not embedded in the porous coating in the laminar areas is continually moving rear- ward into the turbulent areas. This results in a dribbly picture in the intial stages but this "overflow” finally dries away to give a clean transition line or "front." The most satisfactory wetting arrange- ment so far tried is to emit liquid at a rate of about 10 gallons per minute from the wing tip of the cooperating aircraft. This keeps it as far away from the jet engine efflux as possible and so avoids drying the test wing accidentally during the wetting process. The test aircraft is flown some 100 feet behind with the test wing in the spray of liquid. Owing to the angle of attack, the lower surface gets wet first. Tests have been confined to wings outboard of the engines, although kerosene spray has accidentally gone into jet engine intakes on some occasions without noticeable effects. If the tanker aircraft cannot attain the speed required of the test aircraft, the routine is more complicated. A slower drying liquid must be used and the test run lengthened so as to give time to accelerate after wetting and to decelerate for photo- graphy while permitting the full-speed run to dominate the record. The two aircraft keep within immediate reach of each other by orbiting with the tanker on the inside. The liquids that have been used for wetting are aircraft kerosene at heights of about 20,000 feet and stagnation temperatures of about -20°C and "white spirit" (kerosene, boiling point about 180°C) at 35,000 feet and stagnation temperatures of about -40°C. In this more difficult adaptation the drying time must be nicely judged so that at the end of the test run at full speed there is no free film of liquid on the wing. If there is, it mi. blot out the real answer during the decelera. tion. There is, fortunately, a good margin of time between the stages when the picture has become a good “clean” one and when it becomes too faint to be photographed. (3) Transient Conditions 3 If the pilot or crew of the test aircraft cannot see the top wing surface (if the top is wet, the bottom must be wet) the ''tanker" pilot will have to verify whether the test sur- face has been wetted, It has been found best to equip the "tanker" with two cameras and sights, one looking obliquely upward and forward and the other down and sideways. The pilot can thus "fly formation" on the test aircraft and photograph both surfaces. When it is necessary for research pur- poses to know the state of the boundary layer flow under flight conditions which can only be held for a very short duration, then the original chemical method is the only work- able visual technique. It has given satisfac- tory results when the duration of the test condition was only two or three seconds and visual records have been obtained with ex- posures of much less than one second. 9:9 conditions to ease the lighting problems. These results are possible because there is practically no chemical action except while in the active gas and because the concentra- tion of dilution of the gas can be controlled to suit the experiment. These methods require a moderate amount of installation, and parts of the equipment must of necessity be housed in fairings external to a wing. Although one of the tech- niques involves a traveling light source, it is not unduly complicated and has been used very successfully. 9:5 OTHER TECHNIQUES The type of test with which this method can cope is where the aircraft has to be dived to maintain the test conditions (in which case pressure readings become un- reliable) or where, for instance, a wing's range of angle of attack for laminar flow has to be established in free air for comparison with its corresponding range when tested in a low-turbulence wind tunnel. A test of this kind can involve flying at negative incidence down a curved path. Such tests have been done successfully. The marking, aligning, and timing of such experiments requires precision flying. (a) For Separation Two methods that might be useful in flight to determine an area of breakaway flow under special conditions such as where camera positions for tuft photography are difficult, are worth keeping in mind. Since with this method the gas is released into the laying aircraft's wake, it is essential to break up the wing-tip vortex system by flying with the wing flaps lowered. If this is not done, the trail of gas and smoke must be left for at least five minutes before the test aircraft flies lengthwise through it because in this time the trail may be bent by atmo- spheric disturbances. It is helpful to lay the gas trail in a stable layer of air, i.e., where there is a temperature inversion. One has been used in wind tunnel tests and consists of feeding a very small quantity of active gas into the suspected region of break- away. This discolors the surface areas of the breakaway, the region having been sensitized as described in section 9:3. The second method, and there is some flight evidence that it will work, is to dis- charge smoke into the breakaway region in adequate quantity to mark the area involved. Either method can be used without affecting the actual flow breakaway. (b) For Boundary Layer Observations Tests under transient conditions have not so far been required at high altitudes and the chemical method has not been used over 7000 feet. Drier air conditions at high altitude might call for a modification of the chemical action described under "Chemical Methods" in section 9:3, and such a modification is referred to in par. 9:5(b). A chemical method (that has been kept in reserve in case chemical technique tests should be needed in very dry air conditions) has been devised and tested under laboratory conditions. The moisture content needed in the air is only a trace, instead of the 50 relative humidity required for the chemical action described in section 9:3. (c) Shock Wave Observations In carrying out flight tests with the optical chniques described in section 9:3,"Optical Techniques," it is not necessary to do the flying under dusk conditions although the initial development was done under such The active gas used is again chlorine and the sensitive coating on the test surface is of a nature that requires much more effort 9:10 to remove and renew for each test as opposed to washing the starch coating away with a mild scouring powder. On a swept wing the inboard end of a cover must be protected by an overlapping strip to prevent it lifting and, in fact, end strips are always useful in preventing a cross-wind re- moving a cover while the aircraft is taxying out to take off. (c) Protection of Surfaces Before Transition Tests This is sometimes a problem in itself especially with regard to insect contamina- tion near the leading edge of a wing. 9:6 CALIBRATION OF AN ANGLE OF ATTACK MEASURING SYSTEM (a) Description of the Method Dust and grit may be wiped from surfaces treated for dry or wet evaporation tests and this is best done at the take-off point on the runway. Surfaces chemically sprayed may have the forward part of the wing left un- treated if the flow there is known to be laminar and so reduce the risk of turbulence wedges due to grit. This also permits a covering to be used to keep flies off. Because angle of attack is custom arily measured by a vane-type of pressure pickup type sensing device, the instrument readings are influenced by local flow conditions. Thus, the relation between the angle sensed by the instrument and the actual angle of attack of the airplane must be determined by test calibration. Flies collected near the leading edge during take-off and early climb can com- pletely spoil a test. Testing within two hours after sunrise is usually a sufficient safe- guard in summer in a temperate climate but tests in the heat of the day call for a cover that can be shed at altitude. The calibration procedure usually em- ployed consists of first placing the airplane in stabilized horizontal : flight under calm air conditions. In these circumstances the relative wind velocity is horizontal and obvi- ously the weight acts vertically. Therefore, the angle between the airplane's reference axis and the axis of a vertical level in the airplane provides a measure of the actual angle of attack (the angle of attack of the reference line being 90° minus the above- mentioned angle). For small test panels, paper has been used which rips in flight along the leading edge. A more satisfactory method has been to cover the forward surface with two sheets of thin waterproof cloth held on after being wetted with water. The upper cover extends around the leading edge to a point just behind the stagnation point of the wing section at take-off and is shed in flight by increasing the angle of attack to near the stall, when the stagnation point moves back and the forward moving air lifts the edge. The lower cover has its front edge protected from the airstream by the top cover overlapping it slightly. These "fly away" covers usually . depart together. The lower cover should extend to about 20% wing chord if possible but 10 or 12% is sufficient for the top cover to give complete fly protection, The instrument utilized for calibration in France is the vertical level SF IM (J32), attached to one of the standard recorders, All or A20. The inclination of the recorder axis with respect to the longitudinal airplane reference axis is measured on the ground by means of an artillery level. An auxiliary instrument, the adjustable collimeter (SF IM- U-60), is often provided to aid the pilot to maintain stable flight. It is important that the speed be held constant during tests and that vertical atmo- spheric motions be negligible; however, it is possible to correct for deviation of the test flight path from the horizontal, since 9:11 (d) Conduct of Tests and Data Reduction* this flight path angle may be computed from readings of speed and altitude. The angle of attack is then given as the difference of the measured angle and the flight path angle. For cach stabilized run deter- minc: (b) Summary of Purpose of the Tests The airplane attitude from the vertical level The pressure difference dpi Because angle of attack is generally measured in terms of some related parameter, the purpose of the calibration is to relate the actual angle of attack to the measured parameter. The dynamic pressure (compres- sible) Generally, in France, the parameter measured is the pressure difference dpi (furnished by properly placed pickups) divided by the compressible dynamic pressure, qc. This case is the only one considered here since procedures for others may easily be determined. From these data, determine the angle of attack "a". If the vertical speed is zero, the angle is at once determined from the level reading; otherwise, where a: adoY actual angle of attack of reference line (c) Test Equipment Qo = observed angle (1) Required Equipment y: flight path angle Twin pickup angle of attack head Differential pressure gaged for measuring dpi Differential pressure gage for measuring 9c: and tan y: w/V, where w : vertical speed and V : true flight path speed. Knowing a , a curve of dp:/9c: f (a) is prepared for each configuration considered. From the foregoing, curves of a= f (Vcal) or Ci= f(a) may be deduced for each con- figuration at given weights and altitudes. A vertical level A sensitive altimeter 9:7 MEASUREMENT OF SIDESLIP ANGLE An artillery level (2) Desired Equipment Ambient temperature pickup The sideslip angle is defined as the angle between the longitudinal fuselage axis and the flight path as viewed from above the flight path. As pointed out in section 9:1, this angle, like the angle of attack, must be measured in connection with the investigation of stability characteristics and aircraft armament effectiveness. Equipment for determining en- gine operating conditions Thrust meter Manifold pressure torquemeter. gage or Tachometer or revolution counter * Editor's note: Sections on pilot technique and on the duties of the observer have been deleted for the sake of brevity. 9:12 1 Theoretically, the simplest way to meas- ure sideslip angle would be first to calibrate a sideslip detector in a wind tunnel and then mount this same detector on a boom long enough to place it outside the region of disturbed flow created by the airplane. It is, however, impracticable to use a boom long enough to accomplish this purpose and accordingly, other means must be employed than described above. mally located on a boom; however, the differential pressure type may just as well be built into the nose of the airplane. Since the calibration depends on the pickup loca- tion, the location must be specified precisely if usable results are to be obtained. In particular, the location of the support axis with respect to the plane of symmetry is of prime importance. For low speed, light aircraft, it is possible to calibrate the sideslip indicator in flight using a straight road or other similar ground reference to establish a compass heading with the pilot flying along the road at a con- stant sideslip angle. The numerous and ob- vious practical difficulties associated with this method render it useless for tests of high-speed airplanes. In flight, the sideslip angle is measured using a sighting collimeter previously aligned to the aircraft's plane of symmetry. Using this instrument, the angle between the smoke trace and the airplane's longitudinal plane of symmetry may be directly measured and the results recorded using a camera or other means. For high-speed and other aircraft, a theoretically and practically satisfactory method of calibrating the sideslip indicated is to lay a straight smoke trail in the atmo- sphere as a base reference. An airplane equipped with smoke generators follows a straight course and the pilot of the airplane to be calibrated follows the smoke trail at an altitude approximately ten feet below the smoke at a stabilized sideslip angle. As an aid to the pilot in establishing the steady sideslip angle, a thread may be affixed to the windshield (if it has a flat section) and several marks may be made on either side of the thread to provide angular reference. If desired (as in con- nection with stability analysis), a position gyro may be installed to record the angle of bank (if any) or in multiplace airplanes, this same information may be recorded by an observer reading the position of the "ball" of the ball-bank indicator. Similarly, control position indicators may be installed, again to provide information on stability characteristics. (a) Trace Plane Equipment The trace airplane must have performance comparable to the test vehicle and must be equipped with a smoke-laying device. It is not necessary that the smoke used remain visible over great distances; however, the trace should remain distinct for a distance of about 1500 feet aft of the trace airplane. No other special equipment is required for the trace airplane. The normal instruments required for speed and altitude measurement should be installed and, of course, a means provided for recording the indications of the side- slip detector. I (c) Test Configuration and Range of Tests (b) Test Plane Equipment The test configurations to be investigated are determined by the reason for conducting sideslip investigations in the first place. If the tests are conducted in conjunction with stability investigations, the configuration, as well as range of sideslip angles, speeds and altitudes is determined by the extent of Sideslip detectors in common use are of either the differential static pressure or vane type. The vane type detector is nor- 9:13 the desired stability checks. On the other hand, if the sideslip detector is ultimately to furnish information to a gun-sight computer, it may be possible to limit the tests to, say, the "cruise configuration." In conducting the data runs, both the trace airplane and the test airplane are stabilized at the required speed and altitude with the test airplane about 600 to 1200 feet behind the trace plane and at an altitude of about ten feet below the smoke trace. The trace plane must maintain a rigorously constant heading during the tests to provide a straight smoke trail. Normally, the sideslip calibration should not be sensitive to the c.g. location. More- over, if a differential pressure type pickup is employed, the indication should be a function only of the sideslip angle B at low speed. The pilot of the test airplane, using rudder control, adjusts the heading of his airplane to the desired angle aligning the collimeter with the smoke emission point on the trace airplane and, crossing controls, banks the airplane to follow the smoke trail. At higher speeds, Mach number effects will alter the pressure coefficient distribu- tion. For high-speed airplanes, calibrations should first be conducted at low altitude over the range of speeds which the airplane is capable of attaining. High altitude checks should then be made to determine the effect of Mach number. When stabilized flight is achieved, data may be recorded. This operation is repeated for each desired sideslip angle, speed and altitude. The range of sideslip angles to be in- vestigated depends on the purpose of the calibration. Frequently calibrations extend- ing over 6° of sideslip will be sufficient. (e) Test Limitation - Zero Sideslip Angle Checks (d) Conduct of Tests The principal practical difficulty encoun- tered in conducting these tests is that after the pilot of the test airplane has succeeded in nullifying his drift with respect to the smoke, the sideslip angle as read on the collimeter is no longer exactly the desired value. This difficulty can be overcome by using photo recording and then plotting the data to obtain the calibration for the desired sideslip angles. Prior to obtain ing actual data, certain preliminary checks should be made: (1) The system operation should be ground checked. (2) It should be determined that the nose boom if used does not vibrate exces- sively in the speed range to be checked. If vibration is encountered, the natural fre- quency may be lowered by adding ballast along the boom. (3) The range of instrumentation should be investigated to make sure it is adequate to cover the range of pressures and/or angular travels encountered. It is apparent that tests must be conducted in calm air to insure that the smoke trail is not distorted or bent by wind currents. 9:8 AIRBRAKE EVALUATION (4) The required engine speeds at the stabilized flight speeds and altitudes to be tested should be determined to avoid excessive time being devoted to achieving stabilized flight. Airbrakes are the mechanism which allows the pilot to increase the airplane drag and are particularly important on low drag air- planes. 9:14 If v is the descent velocity, then v: V siny, so that the preceding becomes This discussion considers the nature of tests required for the evaluation of airbrake characteristics from the performance stand- point (variation of flight path angle at con- stant airspeed, deceleration value, etc.) and from the handling qualities standpoint (mo- ments due to operation of the brake, resulting vibrations, etc.). Av: ACD = ? 2 mg 9:4 The increased Aco produced by the air- brakes permits the pilot to use them as a control" of deceleration or of flight path angle during a descent. We shall see that the classic flight equations permit us to relate the drag change ACO, the deceleration and the flight path inclination one to the other. From this equation we see that, provided ACO is not a function of Mach number or angle of attack, it is possible (for a given configuration) to determine ACo by merely measuring the change in descent rate due to the airbrakes at one airspeed. If, on the other hand, it is desired to meas- ure the influence of some parameter (for example M) on ACO , measurement of the descent rates at the same angle of attack but at various Mach numbers permits us to obtain the desired results easily. (a) Variation of Descent Angle at Constant Airspeed (b) Deceleration in Level Constant Altitude Flight Let F be the net thrust of the engine (insufficient in this case to maintain altitude at the speed V), and let YI and Y2 be the descent angles with dive brakes retracted and extended respectively. Then with brakes retracted*, summation of forces along the flight path gives If we assume that prior to opening the airbrakes, the speed is stabilized at the value Vi , then F + mg-sin ya co sve F: { vi?cos. 9:5 9:1 and with brakes extended If we presume that the airbrakes open instantaneously and if the pilot has kept the airplane at constant altitude, the instan- taneous deceleration due to the airbrakes may be computed from comparison of Eq.9:5 with F+mg.sin ya 2 (Co+Aco) sve 9:2 and since the flight speeds are the same in both cases, 2 dv F-moto Eco + ACO) SVI dt 9:6 mg (sin 72-sin y) = cosv? , y2 whence 9:3 - OV dt 2 р 2m ACo sve 4C09. *Editor's note: Assuming the thrust line to be aligned with the flight path. 9:7 9:15 Let us consider Fig. 9:5, which presents curves of airplane drag with airbrakes re- tracted (CI), airbrakes open (C2) and thrust available (C3) as functions of airspeed V for one given altitude. Therefore, for a given airplane, the in- stantaneous deceleration will be inversely proportional to the mass; proportional to ACO, which is determined essentially by the shape and arrangement of the airbrakes (but sometimes also by the angle of the attack and Mach number); proportional to the dynamic pressure; or at constant altitude, proportional to the square of the speed and at constant speed proportional to the density p. If the airplane is in steady level flight at point Aj and the airbrakes are extended, the deceleration is given by Since Eq. 9:7 may be written dvi dt - 4,8 AB . 9:9 dy YPM? dt 2 dV YpMSACD 9:8 If the pilot maintains level flight (constant altitude) following extension of the airbrakes, the deceleration at the speed V is where y: specific heat ratio p : ambient pressure dv dt . AB 9:10 it follows that a deceleration started at a given Mach number (due to the opening of the dive brakes) decreases with increasing altitude, where A and B are the points on the curves C3 (not CI) and C2 where the abscissa has a value of V. These conclusions permit us, among other things, to calculate the effectiveness of air- brakes under conditions different from those of the tests. However, it is well to exercise caution for the following reasons: (1) the opening of the airbrakes is not instantaneous and therefore, the speed is no longer V when the drag coefficient achieves a value Co + ACD. Since AB is a function only of Vi , the differential equation (9:10) can be solved by graphical methods and it is possible to determine the decrease in level flight speed at the end of a given time, We conclude these remarks on level flight deceleration by pointing out that in certain cases the deceleration in finite time may vary in the opposite fashion from the instantaneous deceleration (obtained when the airbrakes are first opened). (2) Co may depend on Mach number and angle of attack. (3) Finally, the instantaneous decel- eration is not necessarily the best criterion for airbrake effectiveness. It will often be more informative to determine the decrease in speed over a finite time interval (for example, 5 or 10 seconds) and this deter- mination, as we shall see, is not as simple as that of dv/dt. Consider Fig. 9:6, which illustrates the case of an airplane where Co increases rapid- ly above a certain speed (Mach number), and where Aco due to airbrakes is a constant (curves C, and C2). We shall compare the decelerations obtained starting at cruising speed V below the drag rise to those obtained starting at the maximum level speed vi, which is in the drag rise region. some 9:16 THRUST AND DRAG C2 А B A2 C3 V2 v V V AIR SPEED Fig. 9:5 In the first case, dV - d = 4; Bi dV A, B, where Ai Bi > A, B, > A, BI since for constant ACO and at the end of a given time (say 10 seconds), the speed will have fallen to V3; but in the second case, A1 B; - AB 6). 9:17 Now, from Fig. 9:6, the speed cannot decrease below Vi regardless of the length of time the brakes are used; thus, if (V' V ) < (V V, ), although we have an initially greater deceleration at V,' than at V , the time taken to effect a given speed change may be larger at Vi than at V. To put this another way, reasoning analogous to that which we have just considered can help the test engineer to establish a test program and to in- terpret the various results obtained from it. However, care should be used in extra- polation less fallacious conclusions be ob- tained. For this reason, in the following paragraphs we shall consider the testing of airbrakes under the most varied conditions of utilization. In these circumstances, should we con- clude that the airbrake effectiveness in- creases or decreases with speed? THRUST ه ای AND DRAG C AL c's A B C3 A 1 | V₂ Vz Y V. Vå vi 빌 ​AIR SPEED, V Fig. 9:6 9:18 (c) Detailed Tests Quantities to be Measured and Instruments Required case of a single seat plane, a ground ob- server may assist the pilot using radio communication, for example, by counting the seconds following airbrake deflection, the pilot being responsible for reading the airspeed indicator at 5 and 10 seconds. (1) Altitude should be measured to +1 mb using conventional altimeters or barographs. (1) Loading (2) Airspeed should be measured with a conventional airspeed indicator. (3) Longitudinal deceleration com- puted either from the slope of the curve of true speed versus time or read from a longi- tudinal accelerometer accurate to +0.02g (difficult to obtain) range + 0.2g to -0.8g. The loading of the airplane influences the decelerations and the descent angles and, in practice, an average weight is used for each test condition. It is assumed a priori, that the airplane center of gravity location is unimportant to the tests. (4) Load factor n measured by an accelerometer with a range of -lg to +4g. (2) General Test Conditions* (5) Surface load(s) should be meas- ured within 1 kg. (6) Control position should be meas- ured to an accuracy of 1%- It is important that precise control of the airbrakes be available at maximum deflection to avoid yaw at certain speeds. Since most experience with airbrakes has been gained on high performance combat type airplanes, the comments here apply principally to these. The tests required for a transport plane would certainly be of more limited scope. If unusual characteristics are found, it may be necessary to install longitudinal position measuring gyros, vibration pickups and flow visualization systems such as de- scribed earlier in the chapter. To be of maximum utility, airbrakes should be capable of safe extension throughout the entire airplane operating range of speeds and Mach numbers. A sequence of tests demonstrating this might comprise the following: (7) Instruments Required by the Pilot to Conduct the Tests. For proper per- formance of the tests, the normal panel instruments are generally sufficient provided they are correctly calibrated. However, if not already installed, a pilot's airbrake position indicator with markings at the ex- treme positions is very useful as an auxiliary item. a. Clean configuration - low alti- tude at maximum allowable calibrated air- speed. b. Clean configuration high altitude at maximum allowable Mach number. c. Clean configuration - low alti- tude at (1) calibrated airspeed corresponding to maximum level flight airspeed and (2) to calibrated airspeed corresponding to the speed for best energy climb. (d) Abbreviated Tests It is possible to conduct a brief evaluation of the airbrakes without employing recording equipment using a knee pad, the panel instru- ments and the aid of an observer. In the *Editor's note: This section has been abbre- viated from the original in that specific numerical values have been deleted. 9:19 . d. Clean configuration high altitude at calibrated airspeed corresponding to level flight Vmax. b. It is of utmost importance in the case of the flights of tests c., d. and e., that the pilot establish and maintain level flight as closely as possible. From the tests we may then establish curves of true and calibrated airspeed and longitudinal decel- eration as functions of time. - ap- e. Clean configuration proach speed at reduced thrust. f. Clean configuration - combat dive at reduced thrust at various speeds and flight path angles (angle held constant during descent). (3) Handling Qualities Q g. Clean configuration - descent from high to low altitude at reduced thrust - comparison of descents with and without airbrakes, The measurement of the control forces (or force change) due to opening of the airbrakes should be investigated to determine that the brakes produce no unsatisfactory changes in "feel"'*, which limit their use. In cases where no force change is involved, the air- brakes should also be deflected with the primary controls free to permit measurement of load factor variation and change in trim altitude. . h. Clean configuration let down part flaps - thrust adjusted to hold a given speed or glide path angle. i. Landing configuration -other- wise same as h. A sharp diving moment at the instant of brake extension is generally dangerous, whereas a nose up moment is troublesome if it is of any magnitude, (e) Suitability Tests (1) Usable Limits During extension of the airbrakes, note should be made of: In general it should be established initially that: a. Vibration a. The airbrakes are usable at the maximum Mach number and maximum cali- brated airspeed without the appearance of dangerous phenomena. b. Assymetric effects due to unequal opening time or interaction with a primary control surface. c. Destabilizing effects. b. The airbrakes give the re- quired deceleration or glide path character- istics. (4) Utilization The pilot should report on: (2) Performance Characteristics a. The advantages of using part flap. b. Time of operation a. Increase of flight path angle - The increase in flight path angle due to airbrakes for a given initial speed should be investigated for conditions f., 8., h. and i., listed above. In some instances it is also necessary to investigate the variation, due to airbrakes, of airspeed for a fixed flight path angle. *We are here considering the case of air- brakes, not pull out flaps" which were used toward the end of the last war to produce stalling moments in a dive. 9:20 c. The requirement for a positive position indicator. e. The possible uses during operational flight. d. The control location. f. The use during a let down. REFERENCES 1. Gray, W. e., "A Chemical Method of Indicating Transition in The Boundary Layer," R.A.E. Technical Note, No, Aero 1466, July, 1944. 2. Pringle, G. E. and Main-Smith, J. D., "Boundary Layer Transition Indicated by Sublima- tion,” R.A.E. Technical Note No. Aero 1652, June, 1945. 3. Richards, E. J. and Burstall, F.H., “The 'China Clay' Method of Indicating Transition," R & M 2126, August, 1945. 4. Lamplough, F. E., “Shock-wave Shadow Photography in Tunnel and in Flight,” Aircraft Engineering, April, 1951. 9:21 AGARD FLIGHT TEST MANUAL VOLUME I APPENDICES AGARD FLIGHT TEST MANUAL VOLUME I, CHAPTER 10 PERFORMANCE TESTING OF HELICOPTERS Ву Ralph B. Lightfoot Sikorsky Aircraft Division of United Aircraft Corporation VOLUME I, CHAPTER 10 CHAPTER CONTENTS Page SUMMARY TERMINOLOGY 10:1 INTRODUCTORY COMMENTS 10:1 10:2 FLIGHT TEST PROGRAM 10:2 10:3 GROUND RESONANCE 10:4 10:4 CONTROL AND CENTER OF GRAVITY LIMITS 10:4 10:5 AIRSPEED CALIBRATION 10:6 10:6 LEVEL FLIGHT PERFORMANCE MEASUREMENT 10:7 10:7 CLIMB PERFORMANCE DETERMINATION 10:12 10:8 DESCENT PERFORMANCE 10:17 10:9 TAKE-OFF AND LANDING PERFORMANCE 10:18 10:10 MISCELLANEOUS TESTS 10:23 10:11 CALCULATION METHODS 10:26 10:12 BLADE STALL LIMITATIONS 10:30 CONCLUDING REMARKS 10:13 10:31 REFERENCES 10:32 SUMMARY The performance of a helicopter may be determined by methods similar to those used on conventional aircraft. Particular differences exist in evaluating vertical climb performance and establishing values of profile and parasite drag. Methods have been established to cor- relate power, gross weight, rotor speed, forward speed, altitude, and temperature by accounting for transmission efficiency, changes in profile drag with mean lift and Mach number, and parasite drag with speed. The resultant agreement with the original aerodynamic analysis falls within the accuracy of flight test measurement. Ab- solute values of efficiency, parasite and profile drags must be deter- mined with torquemeters and full scale wind tunnel tests. Otherwise, relative values may be assigned each parameter to render satisfac- tory performance analysis. TERMINOLOGY Mechanical Efficiency Tip Speed Ratio V Flight Path Velocity VNE Never Exceed Speed VH Horizontal Velocity Vy Vertical or Climb Velocity VRB Retreating Blade Velocity BHP Engine Brake Horsepower for Power Curves HAV Power Available to Main Rotor HP Shaft Power to Main Rotor Pi Induced Power PO Profile Power Hp Parasite Power Pc Excess Power for Climb W Gross Weight R Blade Radius с Blade Chord B Tip Loss Factor b Number of Blades P Air Density V Induced Velocity vo Induced Velocity in Hovering Density Ratio oe Rotor Solidity Ratio N Rotor Speed TERMINOLOGY (Continued) 12 Rotor Angular Velocity CLm Mean Rotor Lift Coefficient CT Rotor Thrust Coefficient CDE Parasite Drag Coefficient Inflow Factor u Velocity of Airflow through Rotor ca Torque Coefficient с Damping of the Oleo Struts M Effective Mass of Helicopter Minus Blades n Number of Blades or Maneuver Load Factor E Mass of One Blade C& Blade Damping Rate 1 Mass Moment of Inertia about Drag Hinge E Фр Natural Frequency of Fuselage on Landing Gear A3 1/2 [n.m/(M + nm)] · ml/1 . 1 Distance from Drag Hinge to Center of Mass of Blade р s/wp C/(M + nm) W P р λφ Co/lwp K Empirical Factor v Azimuth Angle in Rotor Plane 83 Offset Angle of Pitch Control A Ground Effect Parameter h Altitude 10:1 INTRODUCTORY COMMENTS inconsistent data and provides a sound basis for establishing performance under any con- dition desired. Numerous methods have been proposed to evaluate helicopter performance. Frequently flight tests have not borne out performance predictions, and therefore some doubt has been cast on the validity of basic helicopter theories. Flight tests have been conducted over a period of years, constantly striving to improve observation and piloting technique, as well as vibration characteristics, instru- mentation, procedures, and analysis. To obtain good flight test results, several cardinal rules apply. Data should be re- corded only during stabilized flight condi- tions. Pilots must be sufficiently trained in performance testing and helicopter opera- tion to hold desired conditions long enough to establish steady state conditions. The vibration levels of the aircraft and the con- trols must not hamper operation. Airspeed indicator installation errors should be mini- mized. Zero airspeed must be carefully obtained for reliable hovering and ground effect data. Rotor speed should produce optimum performance considering vibration, stall, Mach number, and profile drag char- acteristics. As already pointed out, experi- ence indicates that energy methods with proper factors offer the best method of analyzing performance tests. The principal items of performance to be tested are as follows: It has been found that all of these factors have been responsible for the discrepancies noted. Observers have recorded data during unstabilized flight conditions. Pilots have not been sufficiently trained as regards per- formance testing methods and helicopter operation to hold desired conditions long enough to establish steady state conditions. The vibration levels of the aircraft and the controls have hampered operation. Airspeed indicator and altimeter installations have demonstrated extreme calibration variations arising from both dynamic and static pressure errors. Zero airspeed has been difficult to attain for reliable hovering and ground effect data. Rotor speed has not always agreed with design values for the production of op- timum performance because of vibration, stall, Mach number, or profile drag charac- teristics. Attempts to resolve test data into force components have become unwieldy, whereas energy methods required empirical factors for the determination of efficiency, induced, and profile powers. Experience now indi- cates that energy methods used with proper correction factors offer the best method of analyzing performance tests. This system of analysis of flight test data has been used extensively to demonstrate the performance and airworthiness of all Sikorsky helicopters to the certifying and procuring agencies of the U. S. Government. By reversing the procedures of design performance analysis, it is possible to re- duce all test data to basic parameters (Ref. 1). Fairing these data erases the (1) Ground resonance (2) Control and c. g. limits (3) Airspeed calibration (4) Level flight (5) Best rate-of-climb (6) Vertical climb (7) Best rate-of-descent (8) Take-off distance (9) Landing distance (10) Allowable crosswind (11) Fuel consumption (12) Cooling (13) Critical altitude (14) Carburetor heat rise (15) Carbon monoxide (16) Fuel system operation (17) Torque distribution (18) Efficiency (19) Profile drag (20) Parasite drag (21) Stability and control (22) Miscellaneous equipment operation (23) Blade stall and Mach number effects 10:1 These items may be tested in a program similar to that outlined below. b. Ground resonance. c. Flight measurements of magnitude of various exciting frequencies. d. Stress and motion survey in flight. 10:2 FLIGHT TEST PROGRAM (a) Personnel Required (2) Cooling: Engine, accessories, transmission (1) Pilots (2) Mechanics (3) Test engineers (4) Weight engineers (5) Photographers (6) Government inspector (as needed) (b) Apparatus (1) Temperature recorder (2) Pressure recorder (3) Photopanel (4) Oscillograph (5) Theodolitic camera (6) Standard movie camera (7) Speed course (8) Pitot static trailing bomb (9) Carbon monoxide indicator (10) Weighted cord with yarn tufts a. Hovering, max and min RPM at rated power at sea level (overload to hold within ground effect). b. Hovering at rated power at hovering ceiling, normal weight. C. Vertical climb, rated power, normal weight. d. Climb at best rate-of-climb speed to service ceiling. e. High speed, max RPM, 30 minutes. f. Cruising speed, max and min RPM, 30 minutes. (3) Airspeed calibration (c) Weight and Balance (1) Basic aircraft without test equip- ment. a. Longitudinal c. g. location. b. Vertical c. g. location. C. Lateral c. g. location. Pressure survey with pitot static bomb. b. Speed course 30 mph to max speed. c. Pitot static bomb O to max speed, level flight, autorota- tion, and climb. (4) Climb, sawtooth (2) Test aircraft with equipment in- stalled. Weight of each item of test equipment. b. Arm of each item of test equipment. a. Rated power; max and min RPM normal weight; 0, 50, 100 mph. b. Rated power; optimum RPM; normal weight; 0, 10, 20, 30, 40, 50, 60, 70, 80, 90, 100 mph. c. Rated power, optimum RPM, minimum weight, О and best speed. d. Rated power, optimum RPM, max overload weight, 0 and best speed. (d) Flight Tests (1) Vibration a. Survey of vibration charac- teristics (aircraft suspended to simulate flight). 10:2 C. Maneuvers a. (5) Autorotation, sawtooth Descent with no power at speeds indicated in (4) above. b. Descent with no power at five rotor speeds from max to min. (6) Vertical climb 1. 360 degree turn left and right at 0 and 20 mph. 2. Left and right turns at 50%, 75%, and 100% max level flight speed. 3. Acceleration from min to max speed. 4. Deceleration from max to min speed. 5. Pullout from dive at 0, 50%, 75%, and 100% max speed. 6. Pullout from dive and turn at 75% max speed. 7. Max pitching, rolling, and yawing angular ac- celeration at 0, 50%, 75%, max speed. a. Hover in ground effect: wheel clearance 0, 5 ft., 10 ft., D-10 ft., D-5 ft., (D = rotor diameter), rated power, RPM, normal weight. b. Hover out of ground effect: rated power and RPM, nor- mal weight, at ten altitudes up to hovering ceiling 100, 500, 1000, 1500, 2000, 2500, 3000, 3500, 4000, 4500, 5000 ft. C. Hover in ground effect wheel clearance: repeat a at mini- mum weight; 0, 5, 10, 15, D-10, D-5, D ft. d. Hover out of ground effect: repeat b at minimum weight. e. Vertical climb at normal and minimum weight. . (9) Airspeed and RPM limitations, max forward c. g. a. Max rotor speed in autoro- tation, max speed of flare (reduce pitch below min to meet design value). b. Rated power (full throttle) 2000, 5000, 8000 ft. at five rotor speeds from max to min RPM, Note forward speed for greatest tolerable roughness. Correlate with structural flight test sub- stantiation. (7) Level flight a. Power required at sea level, 0 to Vmax at max and min RPM. b. Power required at critical altitude, min to max speed, max and min RPM. C. Power required at 75% of service ceiling, min to max speed, max to min RPM. (8) Controllability (10) Take-off and landing, crosswind tests with theodolitic camera a. Rated power, optimum RPM, accelerate to and climb at 20, 35, 50 mph. b. Repeat a at 95%, 90%, 85%, 80%, 75% power. At best sinking speed in ap- proach, simulate power fail- ure and subsequent landing from various altitudes to de- termine lowest for safety. d. Repeat at 60 mph, 30 mph, 20 mph, 0 mph. C. a. Max aft c. g., max and min RPM; stick position, min to max speed. b. Max forward c. g., max and min RPM; stick position, 20 mph (backward) to max speed. 10:3 e. In ground effect, fly sideways into and with wind right and left, up to max controllable speed. head is desirable to prevent excessive chord- wise stresses at the root of the blade. How- ever, a soft acting gear is usually required to absorb the shock loads of landing. These limitations should be explored and demon- strated to be of such magnitude that all characteristics are achieved, i.e., no ground resonance, low blade stresses, adequate strength of landing gear. (11) Miscellaneous a. Carbon monoxide. b. Heating and ventilating ratio, C. Maximum or critical inclina- tion of fuselage for fuel sys- tem. d. Carburetor heat rise. 10:3 GROUND RESONANCE Ground resonance is a violent unstable motion of the aircraft while in contact with the ground, usually when partially airborne. It is caused by a motion of the blade in the plane of rotation coupled with a rocking motion of the aircraft as a whole. The tires, landing gear, and pylon structure act as a spring with a rate which is incompatible with that of the natural frequency of the blade about a real or effective drag hinge in the plane of rotation. The vibration characteristics may be safely examined if the aircraft is first re- strained by at least four cables attached to the highest stationary point on the main rotor pylon. Two cables on each side may be looped around steel stakes driven in the ground at considerable distance from the helicopter to provide maximum horizontal restraint to the rotor head. A ground crew should stand ready to snub out any vibration which may develop by pulling the loose ends of the cable very quickly if any evidence of resonance appears. Resonance should be checked throughout the complete range of RPM and power conditions before the re- straining cables are released. Excitation may be introduced by periodical application of the control column in circular motions in the direction of rotation at progressive rates from approximately 0.25 to 1.5 times rotor speed, Ground resonance is prevented by rigid blades of proper natural frequency or by a combination of dampers on the blade acting in the plane of rotation and on the landing gear usually serving also as oleo struts. Proper damping is determined as a function of the product of the two damper rates. In Ref. 3 it is noted that for most heli- copters the ground resonance will be sta- bilized when the following inequality holds In operation, the resonance character- istics should be checked during take-off and landing at zero speed, as well as during power-off landings at high forward speeds, as indicated on Flight Test Plan (d)(1)b and (d)(10)c and d. Under all conditions, any oscillation which may be introduced should be damped or immediately made avoidable by a change of power conditions. However, no resonance should occur at any operating condition such as during a rev-up to take-off speed. Ap:λφ > 13 Ź (A-3) 10:1 10:4 CONTROL AND CENTER OF GRAVITY LIMITS Initial tests may be made to check the limits experimentally. Adjustable orifices in the landing gear and the dampers may be set at various positions, and tests conducted for resonance. Low damping in the rotor Mechanical and aerodynamic character- istics of the rotorcraft will impose limitations 10:4 stalling speed of the retreating blade where the actual velocity at the blade element is equivalent to the rotational speed diminished by the forward speed SR-V = VRB: -: on the forward speed, rotor speed, gross weight, altitude, power, and the location of the center of gravity (Ref. 5). Mechanically, the limitations will be imposed because of the inclination of the aircraft and the amount of control required will be limited by the allowable displacement of the control system. Thus, with an aft location of the center of gravity, more forward displacement of the control system is required to attain a given speed. At high power or low rotor speed conditions, more control is required. In general LE C ¿ PSV? and 24 V = ✓ CLPS Thus, the stalling speed If a stalled condition is encountered on the retreating blade, more pitch and control displacement will be required to attain a given speed, The stalled condition may cause undesirable vibration before maximum con- trol or power is attained. The situation may be relieved by decreased pitch and power, by increased rotor speed or a decrease of forward speed, gross weight, or altitude. These relations should be established as indicated in Fig. 10:1 to insure a satis- factory vibration level, VRS 2L va : NR-V CLmax PS and the maximum forward speed in knots for blade stall is 2L V : NR- CL maxes Detailed analysis of these results may lead to consideration of Mach number in- fluence on maximum lift coefficient. How- ever, a practical correlation of data may be made by empirical determination of the NR (30) (1.69) KnW Rcb / leo 140 NR 16.13 KnW Rcb Por 120 10:2 100 2000 FT , M.P.H. 5000 FT Ve 8000 FT. BO where 1.69 converts ft./sec. to knots, Kis the correlation factor determined from flight tests for the particular helicopter, and n is the load factor. As a matter of interest, it has been noted that the magnitude of Vrs or the stalling speed of the blade section, is of the order of 190 knots for a disc loading of three pounds per square foot and a solidity ratio of 0.05. 60 40 160 170 180 190 200 210 ROTOR R.P.M. Test data are obtained under conditions outlined in the Flight Test Program, (d)(9)b. Starting at a relatively low forward speed, Fig. 10:1 Rotor and Airspeed Limitation 10:5 require the least correction of observed data. Since the dynamic pressure at low speeds is given by - į ove, pitch and throttle should be adjusted to maxi- mum positions consistent with the pitch- throttle synchronization and the power rating of the engine to obtain the desired rotor speed. Under the initial condition, a climb will ensue. More forward speed should be achieved by control stick displacement, noting maximum level flight speed. The control stick should be displaced more forward to obtain as high a speed as possible consistent with the structural limitations imposed by the stress analysis or the maximum allow- able stresses revealed by fatigue studies based on the stress survey in Flight Plan item (d)(1)d. we know that 2 (0,-) V and Vcal : Ve 2 2 (07-D) : Po Frequently before these limits are ob- tained, the vibration level of the control stick or the aircraft will become intolerable for pilot comfort or even uncontrollable. This speed does not usually exceed 115% of the maximum level flight speed. The allowable never exceed speed may be taken as 90% of the maximum speed demonstrated (to provide a margin of safety and comfort). Detail study of blade stall should also consider the influence of local Mach number. Within the limits of full throttle operation, a preliminary but practical determination of the never exceed speed appears to be defined by the relation in Eq. 10:2 with K determined ex- perimentally as discussed above. It is seen that proper registration of both static p and total p, pressures is needed for correct airspeed indication. It is also noted that some error may be tolerated in both pressures provided one error compensates the other. A series of total head tubes may be lo- cated in likely positions, and coupled with the static reading of a trailing pitot static bomb to register airspeed on corresponding indicators. Similarly,- several static pres- sures may be picked up by tubes mounted flush to the skin of the fuselage, or other assumed good locations. These pressures . coupled with the total head pressure of the trailing bomb may be used to actuate several airspeed indicators. 10:5 AIRSPEED CALIBRATION The conditions for proper airspeed indica- tion are that the pitot head register the total head due to speed alone and that the static head register a pressure equal to the static pressure infinitely far away from the air- craft. In a rotary wing aircraft, the whole ship is in a slipstream generated by the rotors so that the areas affected only by forward speed are limited. Also, the pres- sure produced by the downwash varies be- neath the disc. Exact locations for the pitot head and static heads are peculiar to each type of ship, and should be determined by extensive survey to arrive at locations that The several indicator readings may be compared with the master indicator (the one connected to the pitot static bomb only) at all airspeeds and altitudes, including climb, autorotation, glide, and yawed flight. While it may not be possible to find locations where separate total head and static pressure are accurately determined, it is usually possible to find a static source location which will compensate for the error of total head regis- tration. Similarly yawed flight with a single static tube may produce errors which can 10:6 where BHP = engine power be corrected by dual static openings on either side of the fuselage. η : efficiency of power transmis- sion accounting for losses to cooling fans, gear fans, gear boxes, torque compensators, etc. Ho profile power to overcome profile drag of rotor blades Tests in level flight at speeds over 30 mph may be conducted over a speed course. Calibration for low speeds, in climb and in autorotation may be determined using a trail- ing pitot static bomb. The altimeter or static pressure error caused by compensa- tion should be checked with and without the rotor turning, in the ground effect and by flying at various speeds at known elevations such as in line with a radio tower and the horizon. The rate-of-climb indication may be checked by noting the sense of indication when changing from level flight to climb or glide. The errors should not exceed those allowed for conventional aircraft. HI : induced power to sunport hel- icopter HPp - parasite power to overcome resistance to forward flight HPc - excess power for climb Pacc : accelerating power for turns and maneuvers 10:6 LEVEL FLIGHT PERFORMANCE MEASUREMENT For unaccelerated level flight - Pacc - Hc=0 and 781P = Hot Pi + Po Reliable power-required data can only be obtained under extremely smooth air con- ditions with exceptionally good piloting tech- nique and with test conditions held long enough to insure stabilized flight (Ref. 8). Data should be obtained at extreme gross weight conditions, several widely separated altitudes, throughout the allowable RPM range, and with the center of gravity located at the extreme and intermediate positions. Tests may also be conducted to determine the optimum yaw angle, using a calibrated yaw head mounted on a boom beyond the rotor slipstream and fuselage interference effects. Profile, induced, and parasite power are discussed in the following paragraphs. (a) Profile Power, HP. ) For a blade element of span dr with re- sultant velocity normal to the blade span axis and in the plane of the rotor disc, Ut dPo = { cod dr vi ui ) 2 . Ut Although several ways of analyzing per- formance data and correlating the various items of performance are theoretically avail- able, numerous measuring difficulties have led to our use of an energy system of analysis. In this system, the power developed by the engine is accounted for by: where cy is the profile drag coefficient and U, is Up : Nr + x R siny. BHP : HO+HP+ 10 + 1Pc+ Hacc η 10:3 The average profile power is obtained by integrating over the radius and around the 10:7 azimuth circle and dividing by 2n; thus, for b blade For the ideal rotor the induced velocity in hovering is 27 R Pore("/" { cod or nex . dr 12 2 2 TPR2 3 + 3 urR sin y + 34? VR? siny Jardy. 3 R' sin? y Because of the high rotor tip pressure gradient losses exist which may be accounted for by inserting a tip loss factor B in the above equation, thus + TR Now, bc : 0 + R, and for average values e of де and T 2 TpRB? ਭਾ Cd, we have 2 TT Po eR Pools 1" Х In forward flight the induced velocity is reduced due to the greater mass flow rate. This problem is considered in various ref- erences and Fig. 10:2 is reproduced from Ref. 18, showing a plot of v/vo against V/vo where V is the net rotorcraft speed. R + KR* sin v + Žu? Rʻsin? + H'R*siny dy Fig. 10:2 shows the induced flow ratios for several values of flight path inclination y, where or се Рcd R$ 12 4 Rºn -1 Vy Y = tan VH . + [12 t +37] ) Ce 1 e Pcd "RRS TT These data apply to cases where the tip speed ratio 4 is less than 0.15. 01 .ر3 + 1) ( *) 8 Bennett, in Ref. 19, suggests that includ- ing the effects of radial flow changes the constant 3 to a value of about 4.65; thus, converting to horsepower If the helicopter configuration includes a fixed airfoil section intended to produce lift and unload the rotor in forward flight, the gross weight, W, must be reduced by the amount of lift received from the lifting sur- face, when substituting W for T in the above relations. ce PCG RØR (1 + 4.654?). 4400 10:4 For the case of no auxiliary lift (b) Induced Power, Pi HP; - 550 W W V 2R2 B2p Vz From the simple momentum theory, the induced power is 10:5 Ty - 550 where v/vo may be obtained from Fig. 10:2 or empirical data. where T = rotor thrust v : induced velocity. We shall now consider the nature of the tip loss factor B for use in Eq. 10:5. Sissigh, 10:8 in Ref. 27, suggests that B - 1 chord at 70% R 1.5 R For a good blade design flight tests have shown that values of B = 0.98 are attain- able; however, rectangular untwisted blades may have values of B as low as 0.90. rotor hub, and all nonlifting components of the helicopter. It may be determined basi- cally from drag estimates of all components of the helicopter and from wind tunnel tests of the fuselage without the rotor blades. Flight test data may then be used to adjust the relationship empirically, also serving to consider scale effect and such other losses not properly accounted for in the theoretical treatment. (c) Parasite Power, Hp The parasite power, Pp, is the power re- quired to overcome the drag of the fuselage, From hovering test data, the induced power, Hi and profile power, Ho may be determined. (See section 10:6(d).) Then the 1.0 0.8 0.6 v/ V. Y=0° 0.4 Y = 2,0° Y=4.0° 0.2 o 4 8 10 2 6 V/v. Fig. 10:2 Induced Velocity in Forward Flight (from Ref. 18) ) 10:9 parasite power is Pp : 7 BIP - H-HR Under test conditions, the value measured is an "equivalent" value; i.e., This is accomplished as follows: From hovering test data we may compute the induced power IP, and at the same time we may directly measure the total power nBIP n BH. Thus, the profile power in hover is IP p = IP pe repe vero IPO = 7BHP - HP and since 4 : 0 during hover, we have From several level flight performance runs at various speeds, gross weights, rotor speeds, and altitudes, the parasite power may be determined and plotted versus for- ward speed. It may be convenient to express the drag in coefficient form, as Ho CeCOPTR$123 . 4400 201 CDF whence Da R? V? cd 440012 CEPTORS ρπΩ - and since Pp: (D4V)/550 11001PV 용 ​PUR? V3 10:6 Cof - Typical Cor data are presented in Fig. 10:3. Now, co is a function of the rotor thrust coefficient or alternatively, mean rotor lift coefficient Clm, so that tests should be conducted throughout a range of gross weight, engine speeds, and altitudes to obtain cd as a function of CT or Clm: In Fig. 10:4 are shown typical test data vs Cum Cum is defined below. · . (d) Data Correlation It was mentioned in section 10:6(c) that hovering tests could be used as a basis for establishing profile power characteristics. of co 016 .05 con .014 .04 012 .03 PARASITE DRAG COEFFICIENT, D DO 010 02 008 .01 .006 O 20 60 80 100 2 LO Vy , M.P.H. 5 CLME MEAN Fig. 10:3 Parasite Drag in Level Flight Fig. 10:4 Rotor Profile Drag Coefficient as a Function of Mean Lift Coefficient 10:10 and A mean rotor lift coefficient may be defined by the equation 2T .BR U drdy ST Re (278° + oH BROD a Te vi drove (2pp BR TT 2 2 3 T: - . Dr Slang) cup cum dr Oy ( с dy also 627 R sin v "**; ) kaina vi drd y = VPRO hence: where B is the tip loss factor. In forward flight no thrust is provided by the reverse flow region of the retreating blade, that is over a circular region of radius O to - R sin y ; thus, with azimuth 0 degrees being forward, we may rewrite the above integral as BR T: 27 CLm. TB'R BR 2CTORY ce TTKBR) 2 R? 27 3 븘 ​6 % 99 ficuCum drony 10:7 Let 27 - R sin y uCumcardy. K: "Olm ر4 B B 2 블 ​2 H 97 For an average chord c, we have then 2 TT BR ܗܬܘ KI?CT). de T: pbcc 21 115.7*03 drev -"";ara 10:8 27.-HR sin y u Fig. 10:5 presents calculations for K at various tip speed ratios for B: 0.98. . T may be expressed as 3.3 TECT PTSR * and since 3.1 bc 1 • TR 2. we have FACTOR "K" 2016 (AR) 2.7 CLm 2 BR is""Dierov- ."S*** sin parou y U drdy 2 2.5 Now 2.3 . .2 N .3 .5 Ur-Br + H NR sin v u$ =(p2 + 2HPR sin y + ki? R? sin?u) Fig. 10:5 Lift Coefficient Factor vs Tip Speed Ratio 3 10:11 As a final check and means of fairing data, all performance information may be reduced to profile drag form, i.e., since rate of turning may affect climb performance, but the usual reduction methods and rec- tilinear climbs appear to provide reasonable data. H: 781P - HP - HP - HPC - and (a) Climbs at Best Rate-of-Climb Speed cd . 4400 (nBIP-HP: - PPP-HPC) σε πρ Ω' R(1 + 4.65μ2) 10:9 With this value plotted against the cor- responding value of The airspeeds for best rate-of-climb at various altitudes are established by the con- ventional sawtooth procedure described ear- lier in this volume. Following the establish- ment of the climb schedule or schedules, continuous climbs to altitude are made (at least two at near maximum gross weight and two at near minimum weight). Readings of altitude, time, and manifold pressure should be recorded every thirty seconds while the airspeed, outside and car- buretor air temperatures, and engine RPM should be recorded periodically. CLM : K 2 Ct e all test data may be well faired to yield basic parameters from which rational per- formance characteristics may be determined for any gross weight, altitude, forward speed, rotor speed, temperature, humidity, etc. While most subsequent calculations will conveniently utilize this cd vs CLm relation- ship, it may be expedient to express the drag in terms of Cų, i.e. The flight test data, corrected for instru- ment calibration, should be entered on the calculation tables. These data are reduced as indicated below. The rate-of-climb is taken as the rate of change of pressure alti- tude and used as the rate-of-climb at the intermediate altitude. It is corrected to true rate-of-climb at standard altitude by the multiplying factor, Co = 8,+8, C++ 8, c The actual techniques of data reduction are presented in section 10:11, which serves to clarify specific procedures. Absolute Outside Air Temperature Absolute Standard Air Temperature 10:7 CLIMB PERFORMANCE DETERMINATION Horsepower is derived from the engine manufacturer's power curve and corrected to actual power at standard altitude by the multiplying factor, Absolute Standard Air Temperature Climb tests should be conducted both under forward speed conditions and in pure vertical flight. JAbs.Carb.Air Temp. x Abs.Out.Air Temp. Tests are conducted at extreme gross weight conditions and with variations in speed for sawtooth climbs as indicated in Refs. 9, 10, and 11. The effect of rotor speed should be studied as well as the effect of c. g. location. Temperature, humidity, and All data are corrected to actual power using power curves, and these values are plotted for each climb. Due to variations of weight and power 10:12 found as follows: throughout the tests, all data must be re- duced to the same gross weight, and rated power at the density altitude of the test. The flight test data corrected to standard altitude are plotted for each climb as func- tions of altitude. The true rate-of-climb values versus altitude are faired using the altitude versus time curve to determine the slope. The rate-of-climb at each altitude where data are recorded should be converted into observed excess thrust horsepower by the formula: Di = Co; s v? Coi is the induced drag coefficient, P is the density of the air in lb. sec. 2/ft. , S is the area of the rotor disc = TR? sq. ft. and V is the true airspeed in ft./sec. The induced drag coefficient is deter- mined by the analogical relation 2 CO; Cis 7D²B2 Obs. Excess : Test Weight x True Rate-of-Climb THP 33,000 where D is the rotor diameter = 2BR, B : tip loss factor : 0.98, and CL is the lift coeffi- cient of the rotor disc defined again by analogy by L where the test weight is determined by com- puting average fuel consumptions and actual fuel consumed during the flight to correct the initial gross weight. q P 2 2 을 ​S v The power developed during the climbs may be somewhat below or above that avail- able at standard altitude. The rotor effi- ciency may be assumed to be 80%. There- fore, the thrust horsepower not used that could have been available for climb is found by: with L being the lift in pounds corresponding to the gross weight, and V again being the true airspeed in ft./sec. By substituting in the above equation for Di, we find that 2 w? D; = TP O v2 82 Die rede TIP Not Used 0.80 (Rated Installed LP - Actual Test Hp ). Multiplying this by Po/ Po the equation becomes 2 W2 To find the available excess thrust horse- power for climb under standard conditions, the thrust horsepower not used is added to the observed excess thrust horsepower. 0, " ; " oʻgʻv? ) The rate-of-climb at the desired gross weight is obtained as follows. The induced horsepower is determined by a formula based on analogy with fixed wing aircraft, i.e. where P Po is the ratio of the density of air at any altitude to the density of the air at sea level, subsequently designated by o. Simplifying this by substituting known values and changing V from ft./sec, to mph, the induced drag becomes . Di 2.97 Di x V 375 HP: 124.5 · W2 o².v².o B? B 10:10 where V is the true airspeed in miles per hour, and D; is the induced drag in pounds (Note that o and oe are not the same quan- tity.) 10:13 Substituting this expression for induced drag, the induced horsepower equation for B: 0.98 is: approximation and then adjusted by level flight test data. Hovering test data deter- mine an average value of mean drag co- efficient at a mean lift coefficient. Level flight at the same lift coefficient allows evaluation of parasite drag from the relation 124.5 · w? 0.353 · W2 PP . 375.02.V.o.8? D2.V.0 10:11 Copa 1100 PT RP v3 (78HP - HP-HO! 10:12 where W is the desired gross weight, D is the rotor diameter, V is the true speed in mph, and o is the density ratio, From the energy equation, The change in excess power is: Profile Power Po, ก BHP - (HP; + P + PC) A Excess Power : HP; (corr.'vt.) - HP (test wt.) and 7 where BHP is the rated power at the altitude, is 0.80 to allow for losses due to cooling, transmission, torque compensation, and non-uniformity of airflow. Excess Power : Avail. Ex. TIP (test wt.)- A Excess Power if the test weight is less than the correct weight; ( A Excess Power is added if the test weight is greater than the corrected weight.) cd Solving for in Eq. 10:4, 4400 PO OPTNR (I + 4.65 ?) cd . The standard rate-of-climb for the cor- rected gross weight can now be found by: Ex. Power x 33,000 SR-O-C: W This equation yields values of cd which are plotted against mean lift coefficient Clm, the mean lift coefficient again being determined by the relation The SR-O-C curves are faired for the high and low weights and from these faired curves are obtained approximate horsepower required for climb by means of the formula CLm 2ÇI . к १८. where the factor K is presented in Fig. 10:5. IPC, २ SR-O-CW 33,000 C As before Induced power is found by CT : W ρπΩ?R : Thrust Coefficient, 124.5 W2 Pi LO 375 D' B? Vo o Solidity Ř po dr 7,0607 : Ratio Parasite power is found by s'RS 10:13 CD, в°рт 1100 CDf • parasite drag coefficient, and may be determined from wind tunnel tests as a first As noted before, values of cd are also calculated from hovering performance and a faired cd vs Cum curve obtained as in Fig. 10:4. 10:14 (b) Vertical Climbs and Hovering Using A, we find that Hi PP i WA W 550 V2 RPBP ſ 10:15 Hovering is a special case of the vertical climb wherein the vertical speed is zero. From the simple momentum theory, the thrust produced by a rotor in hovering or vertical flight is T : Mv, when Mis the mass flow rate of air through the rotor and v, is the total change in velocity imparted to the air moving from rest above the rotor. To determine profile power, tests must be conducted out of ground effect, and then we have IP = RIPA -Hico . The mass flow rate M is wheren M = "R? Bº up; is a factor accounting for trans- mission losses, cooling losses, etc.; n has an average value of 0.80, thus where u is the average axial flow through the rotor. Po: 8HX-HICO Hrea Since V, • 2 v and v: U -V, we have and Ireq is the engine power requirement of the test as measured by a torquemeter or determined from engine charts. T = 2TR 8 u (u-VIP; U . in hovering, V= 0 and therefore Knowing Po, we find cg from Eq. 10:6, i.e., 4400 120 Cd : u? - ਵਲੋਂ T 27 R2 BPP e PT NRS се The induced power in hover is simply Tu, therefore Thus, cd can be determined as a function of Clm from hovering tests out of ground effect. P R = Tv T 2 RBP 10:14 Hovering tests may now be conducted in ground effect and 1 = 1; Hic determined as a function of altitude. Normally, ground effect is measurably present to an altitude equal to the rotor diameter. In Ref. 21, it is suggested that and IP i - W W 550 V2R? BP A = ) where T: W since we are in equilibrium. where D = rotor diameter and h= height of 3/4 radius station of coned rotor above ground. Induced power is less in ground effect than in pure hover, and ground effect may be expressed in terms of a factor A defined as HP: A: HP. i which is less than or equal to one. l; is the induced power far from the ground. This factor should be checked by test on each helicopter. Practically speaking, zero speed performance of the helicopter is dif- ficult to measure directly because of large effects introduced by small deviations from the trim hover condition. 10:15 Care must be exercised to avoid small sinking speeds which include entry into the vortex ring flow state (Ref. 12). Under these conditions, large amounts of power may be expended without regaining a true hover- ing attitude. Several procedures may cir- cumvent this dilemma. Operation in the early morning, about sunrise, frequently provides still air conditions, Anemometers on top of buildings turn very slowly and wind socks droop, indicating wind speeds less than five mph up to altitudes of several hundred feet. Under these conditions ground and building references serve very well to indicate zero airspeed. eliminates the need for exact deterinination of zero airspeed. The test conditions are intended to simulate the worst conditions anticipated and to give substantial variation of thrust coefficient. The standard correc- tions for horsepower and temperature should be made as shown previously in section 10:7(a). In vertical flight with nonzero airspeed T: 2TRB ulu-ve and ст : T PZP - R 2 ulu-V) Bº R? Letting 3 At altitudes reference may be made to a cloud of smoke or to supersensitive indica- tions of a swiveling trailing bomb. A long light plumb line with yarns attached and with a light weight on the end will indicate any slight relative wind by movement of the yarns in a direction opposite to the flight path. Practically, it is sufficiently reliable to use the service airspeed indicator pending accurate results of the calibration done on Flight Plan (d)(3)c. . 12R CT: 282,2- 21182 NR or CT V=NR - Rate-of-climb in fps. B? Experience has indicated that true hover- ing may be attained starting from a level flight attitude of about 40 mph. Power is added gradually, followed by a reduction of airspeed to retain constant altitude. The process is repeated with small increments of power and decrements of speed until zero airspeed is obtained. V = 6012 Ria et -) : Rate-of-climb in fpm. In Ref. 29 it is shown that approximately (translating to our notation) co: CTA + ст. + че са 8 and therefore At several weight conditions, measure- ments should be made of the power required to hover with the wheels several inches, 5 ft., 10 ft., 15 ft., 20 ft., 25 ft., and 30 ft., etc., off the ground under practically zero wind conditions. These flights are considered as being in the ground effect. Flights should also be made to determine the maximum power required to hover at substantially zero airspeed at several altitudes out of the ground effect. All powers should be taken using several engine speeds. [co-ed] ¢ * Ć Ceca 8 = so that V = 60 N R X CT 2 X B4 ਵਿo - 4) - ] In zero airspeed autorotation, the sinking speed may approach 40 fps, but it will require several feet loss of altitude to attain. Up to approximately 10 ft., the sinking speed will not cause structural failure of the aircraft. From 50 to 75 ft., vertical landings may be effected without injury to personnel and only minor damage to the aircraft. Continuing the example, from altitudes between 100 and 300 ft., it would be difficult to make a safe vertical landing, power off. The aircraft may be placarded against hover- ing in this region. At the higher altitudes, some forward speed can be attained and a safe landing effected. Between 300 and 400 ft. altitude, quick pilot reaction is required, while above 400 ft., the landing can be made quite easily. The distance required to land over a 50-ft. obstacle can also be measured satisfactorily with the theodolitic camera equipment. It is desirable to determine the effects of gross weight, speed, altitude, and temperature. As noted above, for some helicopters there appears to be an altitude interval from which a safe autorotative landing cannot be made from a hovering attitude in the event of power failure. Some vertical drop is re- quired to attain control and forward speed. This minimum altitude may be determined as shown on Flight Plan (d)(10)c and d by a succession of flights leaving it to the dis- cretion of an experienced test pilot as to whether he is able to nose down, attain a reasonable glide speed, flare out and roll the ship to a stop without striking a tail rotor blade, tail skid, or tail cone on the ground, For the case cited above, landing proce- dures have been established for power fail- ures at the very low altitudes and at the higher altitudes. When close to the ground, although there is very little time available, the main rotor pitch should be increased as rapidly as possible. At the higher altitudes, pitch should be decreased quickly to increase the rotor speed. The stick should be moved forward to drop the nose and a forward speed of approximately 40 mph acquired to hold the desired rate of descent. When the ground clearance is approximately 35 ft., a flare should be executed by pulling back on the stick to decrease speed to approximately 35 mph. The stick may then be moved for- ward to lower the nose to a level position and to eliminate the possibility of dragging the tail. The succeeding descent to the ground may be controlled by pulling up on the main pitch control to permit a normal landing at a contact speed of approximately 20 mph. The 50 mph approach is usually obtained first. An altitude such as 500 ft. may be selected as the starting point. If the landing is successful and the pilot does not have to apply power at the last instant to avoid a crash, a lower altitude may be attempted, until the lowest reasonable altitude is reached. This program is repeated at vari- ous forward speeds. With large departures from the 50 mph approach speed, more alti- tude is usually required to effect a safe landing following a simulated power failure. The landing gear characteristics will also limit the maximum height from which the helicopter may be dropped following a power failure. Correspondingly, these minimum and maximum altitudes will vary with speed, tending to converge at a minimum approach speed as indicated in Fig. 10:11. For hovering at intermediate altitudes, i.e., 10 to 400 ft. (covered by aircraft plac- ard), a satisfactory procedure is difficult to recommend. At the lower altitudes it might be wise to decrease pitch slightly and then increase pitch rapidly to a maximum. At higher altitudes, it would be helpful to gain forward speed; if an airspeed of 40 mph The maximum rotor speed desired for a satisfactory flare should be determined, one which will not exceed design criteria. The accelerations experienced by various com- ponents of the aircraft should be demonstrated 10:22 to fall below design values for all landings. With multi-engine helicopters, many of the above problems are avoided. while the aircraft is tied down, it is possible to reproduce the numerous power and air- flow conditions experienced in flight, so that prolonged tests are unnecessary. See Chap- ter 6 for data reduction procedures. 10:10 MISCELLANEOUS TESTS (a) Allowable Crosswind (c) Cooling The auxiliary rotor capacity and fuselage characteristics will allow hovering in a considerable crosswind. The magnitude of this wind should be determined so that opera- tion will not be attempted under dangerous conditions. It has been found satisfactory to determine crosswind limits by flying side- ways in view of the theodolite camera equip- ment and noting the speed at which directional control is lost. Experience has indicated that critical cooling usually occurs at zero airspeed and maximum power. Tests may be conducted at hovering ceiling, but the same result may be obtained more easily by overloading the aircraft to lower the hovering ceiling into the ground effect. Basic cooling problems may be studied and corrected under these conditions. However, the adequate cooling of the engine and transmission should be demonstrated in hovering, climb, cruising, and high speed conditions with maximum and minimum RPM (Ref. 15). (b) Fuel Consumption (d) Critical Altitude The variable pitch and RPM possible on a helicopter facilitates the determination of specific fuel consumption. With accurate fuel flow meters installed, it is possible to determine the fuel consumption with a series of ground, hovering, and overload hovering flight tests. Under these conditions, or 500 Critical altitude (Ref. 11) is the highest altitude at which the engine manufacturer's rated power at rated RPM may be obtained under standard conditions of temperature and pressure on a particular installation. Forward facing carburetor airscoops, pres- sure rise of cooling fan, carburetor duct friction losses, temperature rise, etc., may alter the supercharging effect on critical altitude. Forward speed will increase the desirable ram effect while heat transfer through the ducts will decrease the critical altitude, oo AL TITUDE AT POWER FAILURE 200 AVOID 100 The net effect of these factors may be determined by noting the manifold pressure in a climb at full throttle above the critical altitude indicated by the engine manufacturer. The effects of forward speed may be checked by a series of level flight runs at maximum and minimum speeds using full throttle at several altitudes, approximately 1000 ft. apart near the indicated critical altitude. The manifold pressure is corrected to standard temperature conditions at pressure altitude 20 40 60 80 100 Vo , M.P.H. Fig. 10:11 Maximum and Minimum Altitude vs. Speed for Safe Power-Off Landing 10:23 by an impeller chart as presented in Chapter 6, considering the unavoidable carburetor heat rise. The corrected manifold pressure is then converted to power available. It is this power that should be used in correcting performance data to full power conditions, the chart will determine the CO concentra- tion. Such a kit is available from the Mines Safety Appliance Company, Pittsburgh, Penn- sylvania, U. S. A. (g) Fuel System Operation (e) Carburetor Heat Rise Carburetor heat rise is the temperature difference between carburetor air tempera- ture and outside air temperature. The amount of carburetor heat rise available should be determined with variations in outside air temperature and power. The value should be checked in level flight at various speeds to include the effects of ram and in climb. The full cold air value should be deter- mined as well as the ram produced by the cooling fan. The ram developed by a cooling fan may be sufficient to counteract the car- buretor heat rise normally encountered in the full cold position. This effect will be ac- counted for in the critical altitude study described above. Depending on the particular fuel system installation and the quantity of fuel in the tanks at a given instant, it is possible to interrupt the flow of fuel to the carburetor before the tanks are exhausted if the ship is put into some unconventional attitude. It is necessary to study all possible combina- tions of ship attitude and fuel quantity to determine the amount of fuel that may be left in the tanks before engine stoppage will occur. Critical attitudes may be determined in flight; for example, with a forward c. g. the fuselage will nose down the most and possibly uncover the sump of a flat fuel tank. The aircraft is correspondingly oriented on the ground, and the engine run until the fuel pressure warning light comes on or the engine stops, The fuel remaining in the system is then drained and measured. This residual fuel is deducted from the nominal tank capacity and classified as part of the empty weight. (1) Carbon Monoxide Depending upon the configuration of the aircraft, tests should be conducted while hovering in little or no wind, in forward, backward, and sideward motion, having the doors and windows open and closed. Tests should also be conducted in cruise, high speed, and autorotation conditions unless it is apparent that any condition is not critical. (h) Torque Distribution The CO content should not exceed 0.005% (Ref. 14). The CO content is measured by using a small glass tube filled with two chemicals that react and change color when exposed to a CO concentration. The glass tube is open to the atmosphere at one end, and the other end is inserted into a rubber squeeze bulb. A sample of air is drawn into the bulb over the chemicals. The tube is then compared to a standard color chart which is part of the kit. Proper matching of the colors of the chemicals in the tubes with For academic purposes it is desirable to know the distribution of power to the various components of the aircraft. For satisfactory performance analysis, however, it is also necessary to know where the power developed by the engine goes. It is desirable to have torque meters to determine the power de- livered by the engine and that absorbed by the main rotor, auxiliary rotor, transmission, and cooling fan. Torque may be measured with wire strain gages indicating directly on a galvanometer calibrated to read torque. A permanent record may also be obtained on an oscillo- graph, but at a slight loss in accuracy of 10:24 no load registration. The strain gages should be mounted in pairs on opposite sides of any shaft in question with the axes of the gages at 45 degrees to the axis of the shaft to form a complete Wheatstone bridge for accuracy and temperature compensation. The im- pressed voltage on the gages and the re- sponse to load passes through a slip ring arrangement. meters are provided to determine the power delivered to the main rotor. Without the torque meter, it will be necessary to assume a transmission efficiency, obtain a first approximation of profile drag from hovering tests, determine parasite drag on the basis of the initial profile drag values, and com- plete the lift drag polar from climb, and high altitude test data. Other torque meters which might be em- ployed are specially designed shafts for magnetic strain gages or hydraulic restraint of plane pinions in planetary transmissions. The hydraulic pressure should be converted to terms of torque. Approximate values are: Wind tunnel tests of an actual sample of the rotor blade at flight Reynolds number and Mach number will further serve to es- tablish the relationship of section lift and drag coefficients (Ref. 16). Reduction of data to mean rotor blade lift and profile drag coefficients offers a very convenient method of fairing out test data obtained at various altitudes, weights, RPM, speeds, Main rotor .80 - .90 and powers. Auxiliary rotor .07-0 (j) Parasite Drag Transmission .02 - .05 Cooling fan .05 - .10 The most satisfactory information ob- tained on parasite drag has been determined by full scale wind tunnel tests on the air- craft with the rotor removed. With sufficient flight test data, a relative value may be determined which renders satisfactory per- formance analysis. Hovering and level flight tests are conducted at the same mean rotor blade lift coefficient. The torque information makes it possible to determine the efficiency of the transmis- sion system so that a factor (7) may be de- termined for various tip speed ratios (ulex- pressing the percent of power developed by the engine which is delivered to the main rotor. With little or no knowledge of the power distribution, an efficiency of approxi- mately 80% may be assumed. Any discrep- ancy will then manifest itself as profile or parasite drag, but has little effect on the final performance data. The first approximation of profile power obtained in hovering tests, when parasite drag is obviously zero, together with induced power, may be deducted from forward flight data to yield a parasite drag value. It will usually be found that parasite drag coefficient varies with speed and is often higher than anticipated at high speed (see Fig. 10:13). (i) Profile Drag (k) Stability and Control With sufficiently reliable flight test data, it is possible to ascribe relative values to profile drag which, although not absolute, render fairly accurate means of adjusting flight test data for variations of gross weight, altitude, etc. More nearly absolute values of profile drag may be obtained if torque Present stability and control require- ments for conventional aircraft have been relaxed for helicopter application due to the high degree of controllability present and the special pilot training required. For the present, certain elements of stability and 10:25 devices, it is desirable to investigate the stability characteristics and rate them for flying quality. Subsequently it may then be possible to incorporate an automatic pilot. control may be checked. At high speed, forward stick motion should increase speed and back stick should decrease speed. If lateral motion is also required, no reversal should be present. During the high speed dives there should be no tendency for the nose to drop. It is desirable for the rotor- craft to remain level laterally and for the nose to rise with up-gusts when extreme forward or fixed control is applied. During pullouts from dives, only a moderate roll should be allowed, and rotor speed should not exceed prescribed limits during entry or pullout from dive. (1) Miscellaneous Equipment 1 The optional equipment provided for a helicopter should also be subject to examina- tion if it involves the safety and operation of the aircraft or personnel. Thus, a hoist should be checked for capacity, operating characteristics, and effect on performance and control characteristics while in opera- tion. Dual and servo controls should comply with the usual requirements. Litters for rescue work should not endanger the occu- pants by the possibility of carbon monoxide fumes nor require extraordinary technique if they are to be loaded while in flight. Floats should be checked for ground resonance characteristics and consideration given to their power-off landing characteristics. Insufficient experience with stable heli- copters is available to permit dictating the type or degree of stability desired. Static stick position stability as noted above is desirable, i.e., forward stick for forward motion. It may be desirable to introduce a stable stick force gradient, which returns the stick to the trimmed position if left free. However, slight instability here is preferred to any large sacrifice of maneuverability. For instance, since the pilot must direct all his attention to operation when rapidly changing from hovering to high speed, etc., it is detrimental to fight large stick forces while doing so. The opposing inertia forces inherently resist enough without adding more load. For a long range cruising operation, this may not be true, and a small stable stick force gradient is desirable, probably of the order of 0.1 lb./mph, 10:11 CALCULATION METHODS Lateral stick forces may be of similar magnitude. The time of oscillation for the stick-fixed condition should be of sufficiently long period to be readily controllable. 20-second period is acceptable. With an un- stable stick force gradient, the period of the stick-free dynamic instability may be as low as 10 seconds, because the pilot is always in control. Flight test data are reduced to standard conditions by methods outlined in earlier chapters of this volume to the extent of determining gross weight, actual power, true speed, true rate-of-climb, and density altitude. For changes in gross weight of less than 2%, the test data may be scaled up to a desired weight by inverse proportion. For changes in gross weight in excess of 5%, induced power, must be corrected using the ratio (W/W, X (Ref. 17). For changes of gross weight in excess of 5%, the variation of mean lift coefficient will require consid- eration of a new profile drag. For climb test data less than 500 ft./min., all factors usually must be considered. This, then, requires a fairly reliable evaluation of the relation between mean lift coefficient and mean profile drag coefficient. Test data and performance results tend to substantiate the assumptions of uniform induced flow over the effective radius, al- lowing the use of momentum analysis. Twist If stable stick forces of sufficiently high degree are introduced, the difficulty of dy- namic stick-free instability will largely dis- appear. Until continued operating experience is gained with the several possible stabilizing 10:26 where the factor K is given in Fig. 10:5 derived from Ref. 20. and taper may be accounted for in the ex- pressions for tip loss and effective solidity and by use of an equivalent blade angle at 3/4 radius. The variation of induced power with speed and angle of climb may be ex- pressed by the ratio of Fig. 10:2, taken from Ref. 10. The profile power is effective over the whole radius and includes radial losses as suggested by Ref. 19. The parasite power is determined from level flight or full scale wind tunnel tests. The climbing power is that excess power which produces climb. The total power dispersed may then be accounted for by Eq. 10:3. A polar curve may be constructed by plotting Co vs Clm as in Fig. 10:4. Fairing a curve through the scattered points elimi- nates the vagaries of test data accumulated under a variety of conditions of weight, speed, altitude, power, temperature, and piloting techniques. It may be helpful in fairing data for CDF to first fair the data of equivalent parasite power HP pe plotted against Ve, equivalent airspeed. n 7 BIP = 'Po + HP; + Ppt PC + + + c where The empirical values of cd and Cop, Figs. 10:3 and 10:4, when used in the basic power equation, Eq. 10:3, will give properly cor- related performance results. Fig. 10:12 indicates the agreement of original test data to be expected with empirical calculations in climb tests, while Fig. 10:13 shows the agreement in level flight. n : transmission efficiency : BIP: actual brake horsepower delivered by engine Hi : induced W power VOTRBF W induced velocity in VoVZTRAB? hovering (see Eq.10:14) ਨ V W 550 = to V C Special consideration must be given to vertical climb performance. Although Pp: 0 and IP : , 7 BHP - HP - Ho, the in- duced power in vertical climb is reduced because of the improved operating conditions. In ground effect which exists up to a rotor diameter, the induced power is also, as ex- plained earlier, reduced approximately but should be determined for each helicopter. For take-off without a prepared runway surface at altitude or overload conditions, the value of 1 is usually taken for a 10 ft. wheel clearance. . B : .92 for straight untwisted blade .98 for tapered twisted blade Wy : variable with flight path speed V and y from Fig. 10:2. angle - tan - ROC below of climb 60 V 30 mph IPO profile de pod R'n' -(1 + 4.6543 power 4400 PP parasite power 1100 Pc climbing, ROC · W power 33,000 ROC = rate-of-climb in ft./min. The 4400(BHP-IP: LPP IPC) profile ce TPR (1+4.65427 Co Parevo Then corrected vertical climb is ex- pressed in fpm as: . zippe 3 550(Hayitpe! AW W 2TAR B2) ROC = 60 Vc = 60 : drag, cd • ( ) {550 $) ਦਾ ਸ W (Pav- The mean lift coefficient, Cum : K oe from Eq. 10:17. 10:27 = .013 at 300 ft. (.016 at 5700 ft.) 'd . sea (.014 at 2600 ft.) For example, W : 5000 lbs., R = 25.25 ft., RPM = 189, BHP = 500 HP at level (400 HP at 5700 ft.), B = .97, de = .05, coning angle (B.) = 6 degrees = sin' 0.1, height of helicopter = 14.1 ft., disc area TR? : 2000 sq. ft., and tip speed S R = 500 ft./sec. . from Fig. 10:4 = oe PT RP (NR) Рп HO с d' 4400 Find hovering ceiling for 10 ft. wheel clearance above ground. Assume two alti - tude's, -300 ft. (P = .0024) and 5700 ft. (P = .002). (.05)(.013)(.0024)(2000)(125,000,000) 4400 89 HP at -300 ft. (91 HP at 5700 ft.) h : 10 + 14.1 + (.75)(25.2)(0.1) = 26 ft. = P - av .8 (500) = 400 HP at -300 ft. (320 HP at 5700 ft.) D 50.4 ft. : h/D : .516 D. (h/b)} , 0.8 550 (400 - 89 5000 -.8 (40 5000 (40096.97.2 .0024 л w 5000 - - ROC = 60 1.0024 0024) CT- PT R’ (12 R)? (2000)6.0024)(25,000) ) 5000 550(400-89) .00417 at -300 ft. (.005 at 5700 ft.) 1270 ft./min. at -300 ft. - (240 ft./min. at 5700 ft.) (3.29](2) CT . C (3.29)(2)(.00417), .05 m The rate of change of climb .55 at -300 ft. 1270 - 240 5700 + 300 (.66 at 5700 ft., .6 at 2600 ft.) 172 ft./min./1000 ft. 450 20000 o TEST POINTS, SEA LEVEL • TEST POINTS, 10,000 FT. O TEST POINTS 400 16000 SEA LEVEL 4500 LB 12000 350 ALTITUDE, FT. B.H.P. O B000 300 5000 LB. 4000 250 -10,000 FT o 400 O 1600 200 2000 O 20 40 60 80 100 800 1200 ROC, FT./MIN. V , M.P.H. 1 Fig. 10:12 Climb Performance at Best ROC Speed Fig. 10:13 Power Required in Level Flight 10:28 By extrapolation Δh : 240 172 x 1000 = 1400 ft. expected. For example and continuing the case quoted above: = .008 from Fig. 10:3 at 60 mph (88 ft./sec.) Ср The altitude at which the helicopter may hover with its wheels ten feet off the ground is 5700 + 1400 = 7100 ft. CD, por v PTR? f 1100 P Po : 3 Effects of Temperature and Humidity (.008).0024)(2000)(88) 1100 24 HP at -300 ft. и е V 12R 88 500 : .175; 4.654?: .143 It is realized that increased temperature reduces the power developed by the engine and increases the effective altitude of opera- tion for power required. The effect of humidity has been established by the engine manufacturers. Test data obtained at car- buretor temperatures above 75°F and rela- tive humidity above 75% should be used with caution. Every effort should be made to conduct performance tests under atmospheric conditions which avoid these limits. о с РТ 2 R” (SR)". (1+4.65 m?) HO e d 4400 : 89 (1.143) = 102 HP at -300 ft. : ſa W 2TR BP = ve 5000 (4000)(.97 8.0024) : 23.5 ft./sec. To evaluate climb performance at tem- peratures other than that established for standard conditions, it is satisfactory to work from a density basis (Ref. 22). Climb performance is reduced to standard con- ditions of density altitude (Ref. 23). There- fore, for any pressure and temperature the aircraft will experience a certain density altitude, and develops the climb corres- ponding to that density altitude if the engine experiences a temperature normally encoun- tered at that density altitude. V!. 88 23.5 . = 3.75; v/. : .25 from Fig. 10:12. W . Pi N v/v 550 OV2TR BP 5000 - 550 (.25)(23.5) = 53 HP Pc Hav - HOPPp Pi 3 Since this is not the case, the engine power available is subject to a correction (Ref. 24). At that pressure and temperature, a certain power may be developed. The difference between this power and that which would be developed at that density altitude under standard conditions is corrected for trans- mission efficiency and subtracted from the excess power which would have produced the climb at that density altitude under standard conditions. : (.8)(500) -102 -24 -53 = 221 HP ROC 221 x 23,000 5000 . = 1460 ft./min. at -300 ft. (approx. sea level) The corrected excess power converted into climb yields the performance to be At sea level pressure altitude with 100°F 10:29 Hc = (.8) (285) = 102 (.0015) (.028) 7.0024)(.013) : outside air temperature and a 20°F car- buretor heat rise, the density altitude is 2600 ft. where P: .0022. At 2600 ft. under standard conditions, the temperature is 50°F and the engine would develop 450 HP. The excess power for climb may be obtained from the above values accounting for density changes and for increased pro- file drag coefficient noted above. -246.0015) -53 7.0024) 0024 = 9 HP. .0015 ROC 9 x 33,000 5000 : 59 ft./min. - He 7.0022 014 (.8) (450) - 102 1.0024/ 1013 . 1200 - 59 15,000 - 2600 1141 12,400 : 92.2 ft./min./1000 ft. - -24 .0022 -53 ,0024 .0024 .0022 : 182 HP Δh . 50 x 1000 = 640 ft. 92.2 ROC . 182 x 33,000 5000 : 1200 ft./min. at 2600 ft. : By extrapolation ceiling = 15,640 ft. under standard conditions. 10:12 BLADE STALL LIMITATIONS 8 This rate-of-climb can be obtained only when standard conditions prevail at the carburetor corresponding to a pressure al- titude of 2600 ft. and 50°F. But at sea level and 100° + 20°: 120°F the engine can develop actual HP : indicated HP ASAT/ACAT : 500 (460 + 59)/(460 120) = 473 BHP or 23 BHP more than that required for the same performance density altitude under standard conditions. at Unlike a conventional aircraft, rotor blade stall occurs at high forward speed, but actually it is due to the relatively low speed and high angle of attack of the retreating blade (Ref. 25). From this conception, the same factors which affect stalling on a con- ventional aircraft hold true on a helicopter. With a given rotor speed, as forward speed is increased, the retreating blade operates at lower relative speeds, and higher lift coefficients to maintain equilibrium, until the airflow is stalled over the tip (Ref. 26). If transmission efficiency is taken as 80%, the excess power developed for climb is : HP НР. CT + .80BHP: 182+(.8)(23)= 182 + 18 = 200 hp. The climb at 100°F at sea level becomes: ROCI 200 x 33,000 5000 For a given forward speed, as rotor speed is reduced, the same effect is pro- duced. If the gross weight is increased, generally higher lift coefficients will be required precipitating an earlier stall. As altitude is increased, the lower air density will require operation at higher lift coeffi- cients making the stall appear earlier. The first appearance of a stall may be incon- sequential, nor does it spread rapidly to cause trouble as is the case with the con- ventional aircraft. • 1320 ft./min. instead of the 1460 ft./min, obtained under standard conditions at sea level. Continuing the example for standard conditions, at 15,000 ft. P = 0015, Clm =[(55)6.0024))/(.0015)- 88, cd = .028, and engine power : 285 HP. The profile drag increases as the stall is approached, requiring more profile and 10:30 total power. Depending upon the aerody- namic characteristics of the airfoil section in the stalled conditions, the aircraft or the controls will become rough due to the periodic loss of lift and change of aerodynamic twisting moment on the blade. With flexible blades of constant pitching moment, the roughness signs may not be apparent, but an unusually large increment of power and control will be required for increased forward speed. altitude, and center of gravity location are dependent variables and their relationship should be established, At some sacrifice of performance, the minimum RPM, and extreme aft c. g. may be selected and the maximum speed determined at several alti- tudes. Further increase in speed beyond these conditions is not warranted, but safe opera- tion at design gross weight should be possible: Climb performance may be checked up to 75% of the absolute ceiling. However, the blades may stall at 85% without indications from the low altitude test. The test data should be checked up to service ceiling to detect any decrease in slope of the rate of climb vs. altitude curve, or by calculation, using Bailey's methods of Ref. 20, to insure maximum lift coefficient of the tip of the retreating blade does not exceed Clmax 1.6 for smooth airfoil contours and CLmax 1.2 for poor airfoil contours. (a) Up to the “Dive Test Speed,” (b) Down to minimum rotor speed, (c) Up to operating ceiling, and 10:13 CONCLUDING REMARKS (d) With extreme aft location of the center gravity. During (a) the helicopter should be flown at maximum rated rotor speed in level flight up to maximum speed, then nosed over to attain "Dive Test Speed". The test may be repeated at 50% rated BHP, and again in autorotation. The foregoing test and analysis procedure provides a means of establishing the heli- copter possibilities and limitations. While the evaluation of efficiency, induced, para- site, and profile power values may not be absolute, the parameters serve as a medium of correlation for test data obtained under conditions other than standard. The prin- cipal items of performance may be presented on a chart similar to those shown in this chapter. With power on, the minimum rotor speed, 9 10:31 REFERENCES 1. Lightfoot, R. B., "CAA Testing of Helicopter," presented at the Annual Meeting of the Institute of the Aeronautical Sciences, January, 1947. 2. CAA Flight Engineering Report No. 17, “The Apparent Effect of Pilot Technique and Atmospheric Disturbances upon the measured Rate of Climb of an Airplane." 3. Deutsch, M. I., "Ground Vibrations of Helicopters,” Journal of the Aeronautical Sciences, May, 1946, Vol. 13, No. 5. 4. Horvay, G., “Vibrations of Helicopters on the Ground,” Journal of the Aeronautical Sciences, November, 1946. 5. Meyers, Gary, C. Jr., “Flight Measurements of Helicopter Blade Motion with a Com- parison Between Theoretical and Experimental Results,” NACA Technical Note 265, April, 1947. 6. Specification MIL-1-6115 A. 7. Civil Air Regulations, Part 03, “Airplane Airworthiness-Normal, Utility, Acrobatic, and Restricted Purpose Categories." 8. CAA Flight Engineering Report 11, “Methods for Determining Airplane Cruising Range,” August 29, 1944. 9. CAA Flight Engineering Report No. 3, “Airplane Climb Performance." 10. Beerer, J. G., “The Reduction of Flight Test Performance Data to Standard Air Conditions by the Temperature Altitude Method,” Journal of the Aeronautical Sciences, October, 1946. 11. Hamlin, B., “Flight Testing - Conventional and Jet Propelled Airplanes,” The Mac- millan Company, New York, 1946. 12. Glauert, H., “The Elements of Airfoil and Airscrew Theory," The Macmillan Company, New York, 1943. 13. CAA Flight Engineering Report 4, "CAA Equipment for Recording Airplane Take-Off and Landing Characteristics." 14. Civil Air Regulations, Part 06, “Rotorcraft Airworthiness." 15. Specification AA-T-62. 16. Tetervin, N., “Airfoil Section Data from Tests of Ten Practical Construction Sections of Helicopter Rotor Blades,” submitted by Sikorsky Aircraft Division of United Aircraft Corporation, NACA M. R. for AAF, September 6, 1944. 17. CAA Flight Engineering Report 10, “Effect of Airplane Weight Upon Rate of Climb," September 15, 1943. 10:32 18. NACA Advance Restricted Report 15E1D dated June 1945 by Coleman, R. P., Feingold, A. V., and Stenyen, C. W., "Evalutation of the Induced Velocity Field of an Idealized Helicopter Rotor." 19. Bennett, J. A. J., "Rotary Wing Theory," Aircraft Engineering article taken from “Aircraft Engineering," January through August, 1940. 20. Bailey, F. J., "A Simplified Theoretical Method of Determining the Characteristics of a lifting Rotor in Forward Flight," NACA T.R. No. 716. 21. Knight, M. and Hefner, R. A., "Analysis of Ground Effect on the Lifting Airscrew,' NACA T. N. No. 835, December, 1941. 22. CAA Flight Engineering Report 12, "The Effect of Air Temperature Upon the Rate of Climb of an Airplane Equipped with a Constant Speed Propeller," December 1, 1943. 23. CAA Flight Engineering Report 13, “Altitude and Its Effect on Airplane Performance,” June 13, 1944. 24. CAA Flight Engineering Report 17, “The Effect of Power Upon the Calculated Airplane Climb Performance," November 1, 1945, 25. Gustafson, F. B., and Gessow, A., “Effect of Blade Stalling on the Efficiency of a Helicopter Rotor as Measured in Flight,” Proceedings of Third Annual Forum of American Helicopter Society, March 1947, or NACA T. N. No. 1250, April, 1947. 26. > Gustafson, F. B., and Gessow, A., “The Effect of Rotor-Tip Speed on Helicopter Hovering Performance and Maximum Forward Speed,” NACA Wartime Report ARR No. 16A16, March, 1946. 27. Sissigh, G., "Contribution to the Aerodynamics of Rotating-Wing Aircraft," Technical Memorandum 921, NACA, December, 1939. 28. Wheatley, John B., "Aerodynamic Analysis of the Autogiro Rotor with a Comparison Between Calculated and Experimental Results,"Report No. 487, NACA, January 16, 1934. 29. Dommasch, D. O., "Elements of Propeller and Helicopter Aerodynamics," Pitman, 1953. $ 10:33 1 1 i 1 AGARD FLIGHT TEST MANUAL VOLUME I, CHAPTER 11 THE EFFECT OF THE GROUND ON THE PERFORMANCE OF A HELICOPTER By I. C. Cheeseman and J. D. L. Gregory Aeroplane & Armament Experimental Establishment United Kingdom C VOLUME I, CHAPTER 11 CHAPTER CONTENTS Page TERMINOLOGY 11:1 INTRODUCTORY COMMENTS 11:1 11:2 THEORETICAL ANALYSIS FOR A SINGLE ROTOR HELICOPTER 11:1 (a) General 11:1 (b) Ground Effect at Zero Airspeed 11:1 (c) Ground Effect in Forward Flight 11:2 (d) Power Required in the Ground Cushion in Hovering and in Forward Flight at Constant Weight 11:2 (e) The Effect of a Change of Blade Loading on the Ground Effect at Constant Power 11:3 11:3 THEORETICAL ANALYSIS FOR TANDEM ROTOR HELICOPTERS 11:5 11:4 FLIGHT TESTS 11:6 (a) General 11:6 (b) Tests at Constant Weight 11:7 (c) Tests at Constant Power 11:7 COMPARISON OF THEORY AND EXPERIMENT 11:5 11:7 REFERENCES 11:9 TERMINOLOGY A Disc Area b Number of Rotor Blades с Chord at 0.7 Radius CT T/TPR (12R)?, Thrust Coefficient Ст/с Blade Loading Coefficient E Ratio of Effective Power at the Rotor to Total Engine Power G Vertical Separation of Rotors on Tandem Rotor Helicopter k Constant Defined in Section 11:2 P Engine Power PR Power Required to Rotate the Rotor R Rotor Radius S Horizontal Separation of Rotor Centers on Tandem Rotor Helicopters T Rotor Thrust T. Rotor Thrust away from the Ground Tg Rotor Thrust in the Ground Cushion u Total Flow Normal to the Rotor Disc UCO Value of u away from the Ground ug Value of u in Ground Cushion u u'oo Ratio of Normal Velocities through Rear and Front Rotors of Tandem Rotor Helicopter ८ Velocity of Flow Induced by the Rotor Normal to the Rotor Disc Voo Value of v away from the Ground vg g Value of v on Ground Cushion T VT 2 TOR? TERMINOLOGY (Continued) Aircraft Speed > N 6 Height of Rotor General Angle Measured from Mean Direction of Flow when in Forward Flight Defined in Section 11:2(C) η Nondimensional Constant Defined in Section 11:2 Tan" V/v = Mean Direction of Flow under the Rotor 0. Mean Blade Collective Pitch Angle Inflow Ratio : u/R N Р Air Density o bc/1.4R+ Solidity Taken at 0.7 Radius 12 Rotor Angular Velocity 1 11:1 INTRODUCTORY COMMENTS more nearly resembles the flow from a source, and so the method of Betz in which the sink is replaced by a source is used in the zero airspeed case. This model is not sufficiently detailed to show the effect of forward speed and a development has been made which is discussed below. (b) Ground Effect at Zero Airspeed The rotor energy equation at zero air- speed may be written The performance of a helicopter is al- tered by its proximity to the ground. In general the handling qualities of single rotor helicopters are not altered, but some effects have been experienced on tandem rotor machines. The ability to operate close to the ground at zero ground speed in different wind speeds is important in certain opera- tional roles. An approximate theoretical method of estimating the performance of helicopters at any height and airspeed has been devised and compared with experimental results. The theory has also been extended to cover the case of tandem rotor helicopters and this extension is now being checked by flight tests. Because the effect of the ground is de- pendent on the induced flow at the rotor, disc loading is an important parameter when comparing helicopters in the same flight condition, A secondary parameter is blade loading. EP - PR: Tv 11:1 where EP is the effective power at the rotor, PR is the power required to rotate the rotor, T the thrust, and v the induced velocity. Keeping EP - PR constant, the thrust in- side the ground cushion Tg is related to the thrust outside the ground cushion To , by O Tg Too Voo Vg 11:2 11:2 THEORETICAL ANALYSIS FOR A SINGLE ROTOR HELICOPTER (a) General where vg and Vos are the induced velocities inside and outside the ground cushion re- spectively. If 8 Voo is the change in induced velocity due to the presence of the ground, then vg : VO 8 vo and hence 2 To Top 1 SVO VOO 1 - 11:3 Proximity to the ground changes the induced flow through the rotor and this may be mathematically represented by the method of images. To apply this method, a model of the flow beneath the lifting rotor has to be developed. For various purposes several such models have been devised (Refs. 1, 2, and 3) but in all cases the induced flow builds up to the theoretical value of twice the in- duced flow at infinity. If these models are used the value of the ground effect so ob- tained is large when compared to experi- mental values (Ref. 4). A report by Betz (Ref. 5) suggested that the flow around a rotor could be approxi- mated by a three-dimensional hydrodynamic sink. This indicated that the ground effect for a hovering helicopter decreased with de- crease in height. The flow beneath the rotor To calculate 8 Vo , a three-dimensional source having a strength of Avco/4 is assumed, A being the rotor disc area. This value is chosen so that the source has the same flow per unit time as the induced flow of the rotor. Hence the image source pro- . duces an upflow at the rotor of AVDO . SVO ž 16 TT 11:4 € 11:1 Thus by analogy to Eq. 11:2 at constant effective power, where Z is the height of the rotor above the ground. Substituting from Eq. 11:4 into Eq. 11:3, the ground effect at zero airspeed is then expressed by Tg um ug ܢ4-, Too δυ - 11:7 Too 2 R (2) Hence 16 11:5 To 1 Too From this equation the rapid fall-off in ground effect with height is easily seen. 1- ito ( A '+ (c) Ground Effect in Forward Flight 11:8 The simple model used above is not sufficiently exact to show any change of ground effect with forward speed. This is because the directional properties of the flow beneath a lifting rotor have a consid- erable bearing on the ground effect. This expression is plotted in Fig. 11:1 from which it is seen that the ground effect as well as decreasing with height falls off very rapidly with forward speed. Since it is assumed that the disc inclination to the horizontal is zero, a simple relationship exists between V/V, and V/V, Using the experimental data of Ref. 7, the figures are presented in the more usual form with V/VT, instead of Viv, as the nondimensional velocity parameter. a An attempt to allow for this effect has been made by arbitrarily assuming a source with a strength of the form (k Au/47)cos where a is the polar angle measured from the line of symmetry, n? 1 and k is an arbitrary constant, The line of symmetry is assumed to be along the resultant flow direction beneath the rotor. (d) Power Required in the Ground Cush- ion in Hovering and in Forward Flight at Constant Weight In many applications data on the power 1.2 To obtain the strength of the source, the flow over a unit sphere is determined. From this it is found that n must be an even in- teger, the lowest possible value, namely n = 2, being assumed. The total flow is then adjusted by means of the constant k to give equivalence in the case of V = 0 with Eq. 11:4. To/T.1 O Assuming that in low speed level flight the rotor disc inclination to the horizontal is small, then the change in induced velocity su due to the image source can be written VY =0.5 VG sc 1.0 = 2.0 SV . δυ : Au cos20 16 z T 1.0 0.5 1.0 1.5 2.0 2.5 3.0 Z/R 11:6 where 8 = tan (V/v). Fig. 11:1 Variation of Thrust with Height at a Given Airspeed 11:2 required to maintain a helicopter at constant weight in the ground cushion is needed. This is calculated from Eq. 11:8. is therefore made by starting from the blade element thrust formula Too • Ź pobc SPR* (013-12) . 2 11:9 Consider a helicopter of weight W flying at speed V outside the ground cushion with an effective rotor power of (EP-PR). If this helicopter is now moved into the ground cushion the same power will maintain a weight of (Tg/T.) W where Tg/Tc is given by Eq. 11:8. Let (EP-PRO be the effective rotor power required to maintain a weight of (Tg/TO) W at a speed V outside the ground cushion. Then (EP-PR)g/(EP-PR), represents the effect of the ground cushion on power at constant weight. where a is the lift curve slope of the blade section, b the number of blades with mean chord c, the rotor angular velocity, o the mean collective pitch, and the inflow ratio. Since only slow speeds are being considered the terms in have been neg- lected in Eq. 11:9. Typical curves of (EP-PR),/(EP-PRIO are shown in Fig. 11:2. Taking the change in induced flow from Eq. 11:6 an expression for the ground effect is deduced as follows. We assume To is given by the same expression as to except that dg replaces 1. Then (e) The Effect of a Change of Blade Loading on the Ground Effect at Constant Power - To-To- pabc n°RY (1-1) ?. From Eq. 11:8 In Ref. 6 a variation of ground effect with the blade loading parameter CTO is found. An estimate of the effect of vari- ation of the ground effect with blade loading 2 ug TO R ::- (2) * (1-4 ok * U DO Thus V.0.6 | Pabc R3 X 2 : 1+ $ 0.8 Nie (EP-PR'S (EP-PRICE -048 Too 16 () * (1+() 3) ਚ) 1.0 We may express ^ (Ref. 7) as $ + 2.0 Too WA 1 TV2 TPR (NR)? : AVI NR 1.0 0.5 1.0 1.5 2.0 VINT Fig. 11:2 Variation of Power with Forward Speed at a Given Height where is a nondimensional parameter depending on V. 11:3 Then with For any Ст = TO TT PR?(2R)2 G : bc This expression indicates that the ground effect varies with (CT) /o rather than Co as suggested in Ref. 6. particular helicoptero is a constant, and the blade loading parameter is varied at constant disc loading by varying the RPM, Typical curves have been plotted in Fig. 11:3. From these it is seen that an increase in CT/ Co leads to a decrease in the ground effect. In general however the blade loadings of the helicopters have been in the neighbor - hood of 0.1 and it has been found, for per- formance estimates, that blade loading can be neglected and Eq. 11:8 can be used. OTR Tg .it Too 0.25 navo vom 16((71) 71+ SYCAMORE Copy 0.0589 0.0692 0.0777 0.0912 W+ 4600 5400 287 4600 250 5400 250 2 287 0 al 1.3 = 0.0589 0.0692 00777 0.0912 1.2 11 0.5 R To/TO 1.1 Z = 1.0 R FOR ALL 얼 ​VALUES AT = 2.0 1.0 0.4 0.8 1.2 1.6 2.0 VIVA Fig. 11:3 Variation of Tg/T with Speed and Blade Loading at a Given Height 11:4 11:3 THEORETICAL ANALYSIS FOR TANDEM ROTOR HELICOPTERS case is replaced by a source and an image source placed symmetrically beneath itself. In this way any combination of gap and stagger between the two rotors can be treated. Following through the analysis for constant power the expressions for the change of thrust on front and rear rotors deduced are, The simple model developed in 11:2(c) can be used in connection with the tandem rotor configuration. Each rotor in this ola = 0.2 , Ş= 2.0 = 0 Ř= 0.5 , = 1.8 1.25 1.25 NIE 0.5 -1.0 5 U OD UM 1.20 1.20 1.15 1.15 U OD =0 Z R 0.5, = 1.0 To/TC To/TC uoo 2 -0.5, R 름 ​che .ها هنا JO-1.8 $ -1.0, = 1.8 1.10 R 1.10 U... 근 ​R 1.0, = 1.0 uc Z R = 1.0, -1.0, 1.05 . -2.00 름 ​1.05 11 O: 1.0 U. U - 1.8 & unge = 1.8 -2.0, Uge: 1.0 < -2.0,4 Nulac u : 1.0 u'. - 2.0, "=1.8 VOO 1.00 O 2 1 2 R 1.00 o 1 v FRONT ROTOR VT REAR ROTOR >> Fig. 11:4 The Ground Effect on a Tandem Rotor Helicopter 11:5 uce {Tool front = 1 + 16 cos 20 (Z/R)2 u'on cos2(6+Y)cos y Z { + < ***** 2R) In these expressions the value of u uw is the ratio of the normal induced velocity at the front to that at the rear rotor and is introduced to allow for the interference of one rotor with the flow through the other. From Fig. 11:4 in which u". /um takes the value of 1, and 1.8 it is seen that this effect is small and in general this ratio can be taken to be unity. cos? + u'on c. u" 2. NICE + GR sto s Tool =1+ reor 係​) k (+ ( cos? (0-4)Cosy G12 ZR) 2R) + For tandem rotor helicopters, since the ground effect on the two rotors is different, there will also be a pitching moment asso- ciated with the change of thrust, For a helicopter with zero gap this results in a nose-down pitching moment. where cos?0 = 1+ (WUZcos? (8 +4): I > 0 V/)? V/ i + v + [s/122+6)] *{v^v}{22 S 2Z + G This effect has been noticed by the pilots during a cushion take-off as a sudden in- crease in nose-down attitude necessitating a backward cyclic control movement. This effect is modified by fuselage attitude, Figs. 11:4 and 11:5 being drawn for G/R = 0.2 and 0 respectively. The method of cal- culation given here, however, enables estimates of the change of trim due to the ground effect, with any combination of gap and stagger and fuselage attitude, to be determined. 22 + G cosy {s* + (22+ G)2 } GPS 1.125 REAR ROTOR THRUST 11:4 FLIGHT TESTS (a) General 1.100 FRONT ROTOR THRUST T/T 1.075 There are two principal ways of con- ducting flight tests, namely, at either constant weight or constant power. The former test method is preferred if a torquemeter or some other accurate power measuring device is available. Care should be taken not to operate in turbulent conditions and in particular, regions of thermal activity should be avoided. = 0 1.050 pero - 름 ​: 2.0 0.8 R 1.025 O 0.4 0.8 1.2 1.6 V/V, The test techniques described are ap- plicable to both tandem and single rotor helicopters. For tandem rotor machines the fuselage attitude should be maintained con- stant during each test by use of the dif- ferential collective pitch trimmer. The Fig. 11:5 The Ground Effect on a Tandem Rotor Helicopter 11:6 experiments should be repeated for a range of fuselage attitudes. (b) Tests at Constant Weight in the ground cushion. The helicopter is then hovered as near to the ground as possible. While the helicopter is main- tained at constant power the ballast is progessively jettisoned and the new height measured photographically. The ballast is loaded in a manner that keeps the heli- copter center of gravity as constant as possible. By plotting all up weight of the helicopter against height, the value of To is giveh by the asymptote to the curve. The helicopter is hovered at zero ground speed as near to the ground as possible. By increasing the power in small steps the helicopter is hovered at different heights which are measured photographically with a ground mounted camera. The power required at each height is obtained either from torquemeter readings or, if a cali- brated engine is installed, from boost and RPM readings. In this case the normal allowances for tail rotor and gear box power losses have to be made. For some helicopters the disposable load is insufficient to cover the height of the ground cushion, in which case, after the initial test, the helicopter is reballasted and hovered at a new power setting at height just below that reached in the pre- vious test. In this case the power required is adjusted to the value used in the final trip by an appropriate weight correction. The experiment is then repeated with the helicopter in forward flight, the airspeed and wind gradient being obtained as pre- viously suggested. Unless the tests are made in still wind conditions the wind speed at various heights above the ground must be determined. This can be conveniently done by flying the heli- copter, with and against the wind at a constant indicated airspeed, past the ground mounted camera. From the photo- graphic record the ground speed is deter- mined and knowing the aircraft instrument position error the true wind speed at any height is determined. An anemometer is also mounted about six feet above the ground to provide surface wind data. ! In this test the blade loading of the helicopter varies, but in general this variation is small and can be neglected. An estimate of the variation is obtained from 11:2(e). 11:5 COMPARISON OF THEORY AND EXPERIMENT In order to cover the range of forward speeds the helicopter is flown past the ground camera and the photographic record used to determine both height and ground speed. The results obtained are plotted as (EP-PR)g/(EP-PR). against V and Z/R. If the effect of the blade loading coeffi- cient is required the tests should be repeated at the same weight but different rotor RPM. In Fig. 11:6 the theoretical curves ob- tained from 11:2(c) have been plotted with a series of experimental points obtained by the methods described in 11:4 on a variety of single rotor helicopters. It is seen that agreement is good for Z/R greater than 0.6 but that the theory over- estimates the ground effect for lower values of ZR. (c) Tests at Constant Power For this test the helicopter is loaded with jettisonable ballast to the maximum gross weight at which it will just hover Insufficient tests have as yet been made on tandem rotor helicopters to enable a comparison between theory and experiment to be obtained. 11:7 1.2 EXPERIMENTAL RESULTS 0.4 V -=0 V, : 0.5 V, To/Too 1.1 > is is : 1.0 >1> ! 0.5 V : 2.0 V, >> 1. ` = 3.0 1.0 -0-0-0 : 2.0 BI V 1.0 0.5 1.0 1.5 2.0 2.5 3.0 Z/R Fig. 11:6 Variation of Thrust with Height at a Given Airspeed 11:8 REFERENCES 1. Castles and DeLeeuw, "'The Normal Component of the Induced Velocity in the Vicinity of a Lifting Rotor and Some Examples of its Application," NACA TN 2912, March 1953. 2. . Knight and Hefner, "Static Thrust Analysis of the Lifting Airscrew," NACA TN 626, December 1937. 3. Mangler, “Calculation of the Induced Velocity Field of a Rotor," Unpublished British M.O.S. Report. 4. Knight and Hefner, "Analysis of the Ground Effect on the Lifting Airscrew," NACA TN 835, December 1941. 5. Betz, “The Ground Effect on Lifting Propellers," NACA TN 836, April 1937. 6. Zbrozek, "Ground Effect on the Lifting Rotor," British R&M 2347, July 1947. 7. Oliver, “The Low Speed Performance of a Helicopter," British A. R. C. Current Paper 122, May 1952. 11:9 AGARD FLIGHT TEST MANUAL VOLUME I, CHAPTER 12 THE TRANSITION PERFORMANCE OF A HELICOPTER FOLLOWING A SUDDEN LOSS OF POWER Ву I, C. Cheeseman and G. F. Langdon Aeroplane and Armament Experimental Establishment United Kingdom VOLUME I, CHAPTER 12 CHAPTER CONTENTS Page TERMINOLOGY 12:1 INTRODUCTORY COMMENTS 12:1 12:2 THEORETICAL ANALYSIS FOR SINGLE-ENGINE HELICOPTERS 12:1 (a) Engine Failure during Vertical Flight outside the Ground Cushion 12:3 (b) Engine Failure in Vertical Flight inside the Ground Cushion 12:3 (c) Engine Failure in Forward Flight 12:5 12:3 MULTI-ENGINE HELICOPTERS 12:6 12:4 TEST TECHNIQUES 12:7 (a) Single-Engine Helicopters 12:7 (b) Multi-Engine Helicopters 12:8 REFERENCES 12:10 TERMINOLOGY Slope of Blade Lift Coefficient Curve a Forward Acceleration A Power Required to Hover b Number of Blades B Power Available with One Engine Inoperative с Rotor Blade Chord at 0.7 Radius Ср Blade Profile Drag Coefficient at the Mean Effective Lift Coefficient Fuselage Drag DI V2 f T 2 TP R? v? f مها IVE Vertical Flight Coefficients F T 2 TP RU? F, + F Acceleration Due to Gravity 60 h CD 8 H Height Loss Hi á Co DCPR i Rotor Disc Incidence to Flight Path Positive Upward (rotor angle of attack) J Rotor Moment of Inertia k uf AR K 号​+》 ) F 2 m Helicopter Mass TERMINOLOGY (Continued) R Rotor Radius t Time T Rotor Thrust u Resultant Airflow Normal to the Rotor V Induced Flow through the Rotor V Flight Speed C V V x' z Components of the Flight Speed in Direction Ox, OZ V. ( } mg 127 PR2 X,Z, Space Coordinates; x, Horizontal Forward; z, Vertical Downward P. Engine Power at Sea Level n Number of Engines E Efficiency Factor a Rotor Disc Attitude to the Horizontal, Positive Upward х A Constant Blade Pitch do 0 d. mR" H 外 ​NR P Air Density PO Air Density at Sea Level 91 T Ω Rotor Speed 12:1 INTRODUCTORY COMMENTS For safe operation of a helicopter it is important to know the height and speed bands in which an engine failure may result in a crash landing. For single-engine heli- copters engine failure is automatically fol- lowed by a landing, but for multi-engine machines, the ability to continue flight in various situations, such as during take-off and landing from restricted sites, must be assessed. Methods of estimating the per- formance of a helicopter following sudden loss of power have been developed in order to gain some preliminary knowledge of the performance of the helicopter prior to flight trials and also to enable a comparison to be drawn between different helicopter designs. pilot can utilize the forward speed to im- prove the gliding angle, and to maintain the rotor speed. However, during the flare out and landing the forward speed should be not more than 10 knots for a wheeled undercarriage in order to safeguard the crew when operating over rough terrain. There are thus regions in the height-speed plane in which an engine failure may easily result in a crash landing, a typical curve being shown in Fig. 12:1. In attempting to estimate the transition performance of a helicopter it is difficult to allow for the wide variety of possible pilot's control movements and the resulting response of the helicopter. A simplifying assumption has been made, namely that for each stage of the maneuver the pilot main- tains a constant disc attitude to the horizontal and constant collective pitch. 12:2 THEORETICAL ANALYSIS FOR SINGLE-ENGINE HELICOPTERS For axes fixed in space with OX hori- zontal forward and OZ vertical downward and assuming the fuselage drag to be D, ve and the rotor drag to have the ap- proximate form H, VR, the equations of motion are dVx m : - Tsina -D, VVx-H, V S R cos a dt Following total power failure there are three main cases to be considered. If power failure occurs near the ground in vertical flight, the pilot of existing types of helicopter has insufficient time to make any major control changes. . Hence, the helicopter drops back to the ground, the rate of descent being restricted by the util- ization of some of the kinetic energy stored in the rotor. The maximum safe height from which a landing can be made is deter- mined by the energy absorbing properties of the undercarriage. 12:1 m ? dVz dt mg - T cos Q - D, VV2 + H, VSR sina 12:2 d12 J . Tu Ω πρσ R5 8 con? dt 12:3 and in If engine failure occurs in vertical flight when well clear of the ground, the pilot is assumed to reduce the collective pitch in an attempt to maintain the rotor speed at the expense of increasing the rate of descent and then to make a flare out and landing in which the rotor energy is used to reduce the rate of descent to a value which can be absorbed by the undercarriage. It can be shown as in Ref. 1, that T/(2TP R? V?) and u/V are functions of i thus enabling the equations to be solved for any compatible set of initial conditions. Since the relationship for T/(2TPR? V?) and u/V in terms of i and is known only graphically in the low speed region (Ref. 2) the solution is obtained by a step-by-step integration process. In forward flight there are cases cor- responding to the above limits but here the Normally the rotor drag terms in Eqs. 12:1 and 12:2 can be neglected. 12:1 500 400 AVOID FLIGHT IN SHADED AREAS 300 HEIGHT, FEET 200 100 o 50 100 150 AIR SPEED, KNOTS Fig. 12:1 Typical Unsafe Height Band 2.0 VORTEX RING PROPELLER STATE 1.0 WORKING STATE O WINDMILL BRAKE -1.0 STATE 4:v (240R-++ &•v/( 2 } 2 πρR τ Este -2.0 -2.0 -1.0 O 1.0 Fig. 12:2 Vertical Flight Coefficients 12:2 (a) Engine Failure during Vertical Flight outside the Ground Cushion a step increase in the collective pitch setting. A point is finally reached at which the rate of descent is reduced to a figure which can be absorbed by the undercarriage but still keeping the rotor RPM inside its lower limit. The equations of motion for this case take the simpler form since a and Vy are zero. The equations can be written: %E . mg-T dt 12:4 . d12 J C Tu 2 PORS con. 8 By adding together the height lost in the three stages of the calculation an estimate of the safety height is found. So far no allowance has been made for the ground effect. This is a permissible assumption during the first two stages of the calculation, but must be considered during the flare out. Curves showing the variation of thrust with height are given elsewhere in this manual and are also given in Ref. 3. The modified thrust value is fed into the step- by-step calculation. 12:5 In vertical flight there is a unique re- lationship between T/(2*2 R? V?) and u/V since i is a constant and this is normally expressed in terms of the rotor coeffi- cients f, and F. The empirical connection between these coefficients is shown in Fig. 12:2. Using the blade element formula the following relationship is derived: (b) Engine Failure in Vertical Flight inside the Ground Crishion ***** ह 을 ​8 3 ] 12:6 where y = (a,/8), and is used with Eqs. 12:4 and 12:5 and Fig. 12:2 to determine the solution for a collective pitch setting 8. The calculation of this height boundary is most important because of the difficulty of obtaining the results experimentally without severely damaging the test heli- copter. Some experiments were carried out in Great Britain using obsolete heli- copters (Ref. 4) and these confirmed the rapid increase in rate of descent with height at which engine failure occurred. These equations are not exact since no allowance is made for the rotor operating in its own downwash. While it is mathe- matically possible to make such an allowance there is little information on the value of this variable. The solution of these equations is still of use since it represents the case of a near vertical descent which is what is nor. mally achieved in practice. On existing types of helicopters it has been found that the time available between engine failure and touch down is probably too short for the pilot to make any control movements. The motion is therefore cal- culated using Eqs. 12:4, 12:5, and 12:6, assuming that the collective pitch remains at its initial value. Allowance is made for the ground effect as indicated above, Following engine failure it is generally assumed that no control action is taken for a period corresponding to the pilot's re- action time and for the adjustment of the controls to become effective. Following this time (assumed to be of the order of two seconds) the collective pitch is assumed to have its autorotative value. Flare outs are calculated at different parts of the transition to steady autorotation by assuming On the evidence of the Hoverfly tests this method overestimates the safe height, An approximate method has been given in Ref. 4 but this underestimates the height, and therefore is not recommended for final estimation but is useful for comparative performance estimates. To simplify the equations it is assumed that pf, is a constant 12:3 2 2 30 250 NORMAL MAX. 200 NORMAL MIN. 20 OBSERVED R.P.M. BLADE PITCH, ROTOR SPEED, 150 O 2 3 15 ROTOR SPEED, RADIANS/SEC. ACCURATE NUMERICAL SOLUTION 10 10 5 O 2 3 APPROXIMATE NUMERICAL SOLUTION 30 20 ACCELERATION, FT./SEC.? 1 TIME FROM ENGINE CUT, SECONDS 10 O Fig. 12:4 Rotor Speed at Constant Pitch o 2 3 20 15 10 RATE OF DESCENT, FT./SEC. 5 15 O 2 APPROXIMATE NUMERICAL SOLUTION 3 15 RATE OF DESCENT, FT./SEC. 10 OBSERVED 10 HEIGHT LOST, FT. 5 5 5 3 2 TIME FROM ENGINE CUT, SECONDS ACCURATE NUMERICAL SOLUTION Fig. 12:3 Performance of a Single Rotor Helicopter with Pitch Maintained at Hovering Value until Touch Down (engine failure near the ground) 1 TIME FROM ENGINE CUT, SECONDS Fig. 12:5 Rate of Descent at Constant Pitch 12:4 k, determined from the hovering conditions. The equations can then be solved to give so -k Novor? 12= , V: К по +1 KNOT+1 The ground effect may be allowed for by estimating a mean value from the curves of Ref. 3 and using this to modify the value of k. Comparison of these results with the flight tests of Ref. 4 is shown in Figs. 12:3, 12:4, and 12:5. T 12:7 where (c) Engine Failure in Forward Flight Kamp к (ht 2 Cases similar to those considered in earlier paragraphs of this section exist in forward flight. There is now the added complication that the pilot's control move- ments are not uniquely defined. In Ref. 1 an investigation into the effect of control movements on the height lost was made. and the other symbols are as defined in the accompanying table. RATE OF DESCENT, FEET/SEC. FORWARD SPEED, FEET/SEC. ܘ̄ ܘ̄ ܘ̄ ܘ o RATE OF DESCENT, FEET/SEC. FORWARD SPEED, FEET/SEC. 2 4 FOR DISC ATTITUDE Q: -5° TIME, SEC. 8 1 9 O 01 20 30 130 o o o ā ģ 26 27 28 29 a: +7° TIME , SEC. 2 4 6 HEIGHT LOST, FEET ROTOR SPEED, RADIANS/SEC. 50 100 150 200 26 27 28 29 o o 8 o 50 001 100 200 26 27 28 29 ROTOR SPEED, HEIGHT LOST, FEET 2 TIME, SEC. 4 6 RADIANS/SEC. 8 o 영 ​ở RATE OF DESCENT, FEETI SEC. FORWARD SPEED, FEET/SEC. Fig. 12:6 Estimated Transition Performance, Effect of Control Movements 12:5 Estimates were made for a helicopter as- sumed to be flying initially at 112.6 ft./sec. Three different disc attitudes (a : -5 de- grees, 0 degrees, and +7 degrees)were as- sumed and the performance calculated from Eqs. 12:1, 12:2, and 12:3. If an engine fails while the helicopter is flying outside this speed range the speed must be increased or decreased if the flight is to continue and while this is done some loss of height is inevitable, This height loss determines the take-off and landing technique to be used when operating from restricted sites and the minimum size of the landing ground required for safety if an engine fails at any point on the flight path. These results are shown in Fig. 12:6. From these it is seen that the steady con- ditions are achieved more rapidly and with a smaller height loss by adjusting the control so that the helicopter approaches its best steady gliding speed. To complete the cal- culation it is necessary to calculate the height lost in making a flare out, the end condition being a rate of descent which can be absorbed by the undercarriage and a forward speed less than 10 knots. Fig. 12:7, due to Hafner (Ref. 6), shows how the single engine performance affects the take-off technique. The behavior of the helicopter after one engine has failed can be calculated by in- tegration of the equations of motion if the control actions are specified. 12:3 MULTI-ENGINE HELICOPTERS These equations equations can written, (Ref. 5), conveniently be The case of a helicopter with more than one engine is rather different from that of a single-engine machine as complete power failure is unlikely to occur. Even in this case, however, it may still be possible to make an autorotative landing. dV m og -H A:T sin 2 - D, VVx-H, VSR cos a d dVz = (mg - Tcos a)-0, VV2 + H, VSR sin a a + dt In some cases the aircraft will not be able to hover with one engine inoperative although it may be able to climb within some restricted range of forward speed. .J d2 dt EPon по [] - - PCOSRºs? P peces NORMAL TAKE-OFF PATH TAKE-OFF PATH IF ONE ENGINE FAILS ABOVE CRITICAL HEIGHT. ܠܕ€ २० RETURN PATH IF ONE ENGINE FAILS BELOW CRITICAL HEIGHT. Fig. 12:7 Take-Off Technique for Restricted Sites 12:6 The work needed to accelerate the heli- copter is The last two terms, representing the induced and parasitic torque and blade profile torque respectively, are taken as the sum of their values for each lifting rotor in the case of multi-rotor machines. 2 mys. mx (-2) mv ' . B A The work done by gravity is mgH, and the work done by the engine is B XB (-) 87, . These equations must be integrated by some step-by-step method which is labo- rious. The accuracy of prediction depends on the degree of approximation to which the control movements are estimated. As in the single-engine case, it is usual to assume that after an arbitrary pilot reaction time of two seconds the collective pitch is reduced to the value appropriate to steady flight with one engine failed and that during the transition to the final forward speed the disc attitude a is maintained constant. Since the sum of these various work items must be zero X H: 29 (-2) (4 m2 +1) . A B А. ܘܬܐ) An alternative approach based on energy balance has been made in Ref. 6. Consider the case of a helicopter which can climb with one engine failed but which has insufficient power to hover in this con- dition. Let the power required to hover be A and the power available with one engine out be B. When one engine fails while the aircraft is hovering the pilot is assumed to take action to accelerate the helicopter to climb-away speed. Then assuming that the power required varies linearly with speed up to the speed at which height can be main- tained on one engine, power required = A ( 1 - V/X) where X is a constant which can be found from the power curve for the aircraft. A test program carried out out using a single-engine helicopter fitted with a device which enabled power to be reduced suddenly (Ref. 7) showed that the energy method is simple and quick and gives a fair estimate of the order of height loss to be expected. Step-by-step integration of the equations of motion entails considerably more work but makes possible the comparison of dif- ferent control techniques. 12:4 TEST TECHNIQUES (a) Single-Engine Helicopters Thus B : 1 - Å ; t. (%) (-2). If the forward acceleration is a constant, a, then the total work required for sustenta- tion during the time t, in which the helicopter reaches the speed V, is (1) Simulated engine failure near the ground. An estimate of the height from above which a safe landing using rotor energy is not possible can be made as part of the handling assessment of the aircraft. It is obviously not possible to determine accu- rately this height experimentally for each new type because of the risk of damaging the helicopter, however a series of tests using obsolete aircraft has been carried out (Ref. 4) with the results indicated in section 12:2 (b). B2 S'ali-*) at dt : АХ 20 (-). 12:7 (2) Simulated engine failure at alti- tude. A fair idea of the height required to establish steady autorotation can be obtained from visual observations of the rotor speed indicator and altimeter during sudden transi- tions to autorotation from powered flight at various speeds. To make a complete in- vestigation an auto-observer or continuous trace recorder must be fitted to record control positions, accelerations, speeds, al- titude, and attitude. Pilot reaction time may have a critical effect on the behavior of helicopters with low inertia rotor systems and the effect of a delay before corrective action is taken should be investigated. Tran- sitions in which the horizontal speed is both increased and decreased from the initial flight speed should be made and the optimum technique at each speed determined. The first stage is to establish the steady rate of climb or descent with one engine throttled back and disengaged from the rotors. To do this, partial climbs or descents are carried out over as much of the forward speed range as possible, using standard per- formance methods. It is useful to carry out this test with the operative engines running at both the five minute power and maximum continuous ratings. Although the . performance should not depend on which engine it is decided to disengage, it may be advisable to carry out tests with each engine throttled back in turn to check on the cooling of the engines in use. C C Typical experimental results obtained on Hoverfly I aircraft (Ref. 1) are shown in Figs. 12:8, 12:9, and 12:10. When the steady performance is known, the loss of height following an engine failure during level flight can be found by disen- gaging one engine as quickly as possible and using an auto-observer to record control movements, rotor speed, altitude, airspeed and accelerations. At this stage it should be possible to judge the effect on fall-off of rator speed of a delay before taking control action after failure of one engine and to decide whether an engine failure warning indicator is necessary. (b) Multi-Engine Helicopters The test program is designed to discover the performance following engine failure in various conditions of flight. There remains the question of behavior after an engine failure during take-off or landing. If the earlier tests show that the helicopter is capable of hovering with one engine out and the remainder at emergency" 1 K TIME AT WHICH ENGINE WAS CUT STICK POSITION ba 2.5 FWD o AFT 2.5 STICK POSITION, INCHES FROM CENTRAL 300 COLLECTIVE PITCH, DEGREES COLLECTIVE PITCH 0 0 60 50 200 AIRSPEED AIR SPEED, KNOTS 40 HEIGHT LOST IN REATTAINING INITIAL ROTOR SPEED (236 R.PM) 250 30 ROTOR SPEED HEIGHT LOST, FEET SPEED, ROTOR R.P.M. 225 100 200 NORMAL ACCELERATION 200 9 NORMAL ACCELERATION, O HEIGHT LOST, FEET HEIGHT LOST IN ATTAINING APPROX. STEADY RATE OF DESCENT 100 HEIGHT_LOST o 2 8 10 6 TIME, SECS. o 20 30 40 50 60 INITIAL AIR SPEED IN LEVEL FLIGHT, KNOTS Fig. 12:8 Transition to Autorotation after Engine Cut in Level Flight Fig. 12:9 Height Lost in Transition C 12:8 240 INITIAL ROTOR SPEED IN LEVEL FLIGHT 230 0_ ROTOR SPEED AT TIME OF ATTAINING STEADY RATE OF DESCENT investigate the effect on the control of the aircraft of failure of one engine and the effect of opening the others to full power. If, as will probably be the case with a twin-engine machine, the helicopter cannot hover with one engine failed some such flight path as that shown in Fig. 12:7 will have to be adopted in the event of an engine failure during take-off. Since the shape of this flight path will determine the size of the site and the obstruction line for safe operation it is important that it should be measured accurately. 220 ROTOR SPEED, R.P.M. 0 MINIMUM ROTOR SPEED DURING TRANSITION 210 NORMAL MINIMUM PERMISSIBLE ROTOR SPEED 200 190 20 30 40 50 60 INITIAL AIRSPEED IN LEVEL FLIGHT, KNOTS Flight tests have been carried out in Great Britain in which one engine of a twin-engine helicopter was throttled back suddenly during vertical and backward take- offs and the flight path during the subsequent transition to forward flight recorded by kine-theodolite. An auto-observer and Hussenot continuous trace recorders were fitted in the aircraft. Full results of these tests are not not yet available, however it appears that auto-observer records of altim- eter and A.S.I. readings may yield a suf- ficiently accurate flight path shape, as shown by the comparison of one actual flight path measured by theodolite with that deduced from auto-observer instruments (Fig. 12:11). Fig. 12:10 Variation of Rotor Speed in Transition or "take-off" rating, it will be possible to make a safe landing under these conditions from any height and it is only necessary to 400 0 HEIGHT, FEET 200 O FROM AUTO-OBSERVER FROM KINE THEODOLITES 400 800 1200 HORIZONTAL DISTANCE, FEET Fig. 12:11 Flight Path after Single Engine Failure 12:9 REFERENCES 1. O'Hara and Mather, “The Performance after Power Failure of a Helicopter with Blade Pitch Control, Parts I and II," British R. & M. 2797. 2. Oliver, “The Low Speed Performance of a Helicopter," British A.R.C. C.P. 122. 3. Cheeseman and Bennett, “The Effect of the Ground on a Lifting Rotor,'' Unpublished British M. O. S. Report. 4. Wilkinson, “The Performance of a Single Rotor Helicopter Following Power Cut whilst Hovering. Flight Tests with a Hoverfly I near the Ground," Unpublished British M. U. S. Report. 5. . Oliver, “The Performance of a Multi-Engine Helicopter Following Failure of One Engine during Take-Off or Landing," Unpublished British M. O, S, Report. 6. Hafner, "The Domain of the Helicopter," Seventh Bleriot Lecture, Journal of the Royal Aeronautical Society, October 1954. 7. Langdon, “An Experimental Investigation into the Performance of a Helicopter Fol- lowing Sudden Reduction in Power," Unpublished British M. O. S. Report. را C Air Force-USAFE, Wsbn, Ger- 56-2628 12:10 APPENDIX I, PARTI ICAN STANDARD ATMOSPHERE Temperatures Pressures* PP Altitude (feet) ICAN Standard ос °C Abs. psi Absolute Speed** of sound ft.per sec. Millibars p/Po 0 1,000 2,000 3,000 4,000 5,000 1.000 0.985 0.971 0.957 0.942 0.928 15.0 13.0 11.0 9.1 7.1 5.1 288.0 286.0 284.0 282.1 280.1 278.1 1,013.2 977.1 942.1 908.1 875.1 843.0 14.70 14.17 13.66 13.17 12.69 12.23 1.000 0.964 0.930 0.896 0.864 0.832 1,117 1,113 1,109 1,105 1,102 1,098 6,000 7,000 8,000 9,000 10,000 0.914 0.900 0.887 0.873 0.859 3.1 1.1 -0.8 -2.8 -4.8 276.1 274.1 272.2 270.2 268.2 812.0 781.8 752.6 724.3 696.8 11.78 11.34 10.92 10.51 10.11 0.801 0.772 0.743 0.715 0.688 1,094 1,090 1,086 1,082 1,078 11,000 12,000 13,000 14,000 15,000 0.846 0.833 0.819 0.806 0.793 -6.8 -8.8 -10.7 -12.7 -14.7 266.2 264.2 262.3 260.3 258.3 670.2 644.4 619.4 595.2 571.8 9.72 9.35 8.98 8.63 8.29 0.661 0.636 0.611 0.588 0.564 1,074 1,070 1,066 1,062 1,058 16,000 17,000 18,000 19,000 20,000 0.780 0.768 0.755 0.742 0.730 -16.7 -18.7 -20.6 -22.6 -24.6 256.3 254.3 252.4 250.4 248.4 549.1 527.2 506.0 485,5 465.6 7.97 7.65 7.34 7.04 6.75 0.542 0.520 0.499 0.479 0.460 1,054 1,050 1,046 1,041 1,037 21,000 22,000 23,000 24,000 25,000 0.718 0.705 0.693 0.681 0.669 -26.6 -28.6 -30.5 -32.5 -34.5 246.4 244,4 242.5 240.5 238.5 446.4 427.9 410.0 392,7 376.0 6.48 6.21 5.95 5.70 5.45 0.441 0.422 0.405 0.388 0.371 1,033 1,029 1,025 1,021 1,017 ICAN standard conditions ** ICAN standard conditions. tional to T. For conditions other than ICAN, the speed of sound is propor- A-1 APPENDIX I, PART I ICAN STANDARD ATMOSPHERE Temperatures Pressures* Speed** Altitude (feet) volpo ICAN Standard OC OC Abs. psi Absolute Millibars of sound ft.per sec. p/PO 26,000 27,000 28,000 29,000 30,000 0.658 0.616 0.635 0.623 0.612 - 36.5 -38.5 -40.4 -42.4 -44.4 236.5 234.5 232.6 230.6 228.6 359.9 344.3 329.3 314.9 300.9 5.22 4.99 4.78 4.57 4.36 0.355 0.340 0.325 0.311 0.297 1,012 1,003 1,004 999 995 31,000 32,000 33,000 34,000 35,000 0.601 0.589 0.578 0,568 0.557 -46.4 -18.4 -50.3 -52.3 -54.3 226.6 224.6 222.7 220.7 218.7 287.5 274.5 262.0 250.0 238.4 4.17 3.98 3.80 3.63 3.46 0.284 0.271 0.259 0.247 0.235 991 987 982 973 973 -56.3 -56.5 216.7 216.5 969 968 36,000 37,000 38,000 39,000 40,000 0.546 0.533 0.521 0.508 0.496 227.3 216.7 206.5 196.8 187.6 3.30 3.14 2.99 2.85 2.72 0.224 0.214 0.204 0.194 0.185 41,000 42,000 43,000 44,000 45,000 0.485 0.473 0.462 0,451 0.440 178.8 170.4 162.4 154.8 147.5 2.60 2.47 2.35 2.24 2.14 0.176 0.168 0.160 0.153 0.146 46,000 47,000 48,0000 49.000 50,000 0.430 0.419 0.409 0.400 0.390 140.6 134.0 127.7 121.7 116.0 2.04 1.94 1.85 1.76 1.68 0.139 0.132 0.126 0.120 0.115 ICAN standard conditions ** ICAN standard conditions. tional to ✓T. For conditions other than ICAN, the speed of sound is propor- A-2 APPENDIX I, PART II PROPERTIES OF THE NACA STANDARD ATMOSPHERE h, T, o Vc, v X 104 (1000 ft.) | °F, Abs. P, P. in. Hg. slugs per cu.ft. mph ft. per sec. 0 ܘܫ ܘܬ ܚ ܛ 2 3 4 5 518.4 514.8 511.3 507.7 504.1 500.6 29.92 28.86 27.82 26.81 25.84 24.89 0.002378 .002309 .002242 .002176 .002112 .002049 1.0000 0.9710 .9428 .9151 .8881 .8616 1.0000 0.9854 .9710 9566 .9424 .9282 760.9 758.3 755.7 753.0 750.4 747.7 1.567 1.604 1.644 1.684 1.725 1.768 Onda 497.0 493.4 489.9 486.3 482.7 23.98 23.09 22.22 21.38 20.58 ,001988 .001928 .001869 ,001812 .001756 .8358 .8106 .7859 7619 .7384 9142 .9003 .8865 .8729 .8593 745.1 742.3 739.7 737.0 734.3 1.812 1.857 1.905 1.954 2.004 10 11 12 13 14 15 479,2 475.6 472.0 468.5 464.9 19.79 19.03 18.29 17.57 16.88 .001702 .001648 .001596 .001545 .001496 .7154 .6931 .6712 .6499 .6291 .8458 .8325 .8193 .8062 .7931 731.6 728.8 726.1 723.4 720.6 2.055 2,109 2.165 2.223 2.281 16 17 18 19 20 416.3 457.8 454.2 450.6 447.1 16.21 15.56 14.94 14.33 13.75 .001448 .001401 .001355 .001311 .001267 .6088 .5891 .5698 .5509 .5327 .7803 .7675 .7548 .7422 .7299 717.8 715.0 712.2 709.4 706.6 2.342 2.405 2.471 2.537 2.608 21 22 23 24 25 443.5 439.9 436.4 432.8 492.2 13.18 12.63 12.10 11.59 11.10 .001225 ,001183 .001143 .001103 .001065 .5148 .4974 .4805 .4640 .4480 .7175 .7053 .6932 .6812 .6693 703.8 701.0 698.1 695.3 692.4 2.680 2.756 2.834 2.916 3.000 26 27 28 29 30 425.7 422.1 418.5 415.0 411.4 10.62 10.16 9.720 9.293 8.880 .001028 .000992 .000957 .000922 .000889 .4323 .4171 .4023 .3879 .3740 .6575 .6458 .6343 .6228 .6116 689.5 686.6 683.7 680.8 677.9 3.086 3.175 3.268 3.369 3.468 * ICAN standard conditions ** ICAN standard conditions. For conditions other than ICAN, the speed of sound is propor- tional to T. A-3 APPENDIX I, PART II PROPERTIES OF THE NACA STANDARD ATMOSPHERE I, h, (1000 ft.) v X 104 °F, Abs. P, P, in. Hg. slugs per cu.ft. Vc, mph ft. per sec 31 32 33 34 35 407.8 404.3 400.7 397.2 393.6 8.483 8.101 7.732 7.377 7.036 .000857 .000826 .000795 .000765 .000736 .3603 .3472 .3343 .3218 .3098 .6003 .5892 .5781 .5673 .5566 674.9 672.0 669.0 666.0 663.0 3.570 3.678 3.792 3.911 4.034 35.332 36 37 38 39 40 392.4 392.4 392.4 392.4 392.4 392.4 6.926 6.711 6.397 6.098 5.813 5.544 .000727 .000705 .000672 .000640 .000610 .000582 .3058 .2963 .2824 2692 .2567 .2448 .5530 .5443 .5314 .5188 .5067 .4948 662.0 662.0 662.0 662.0 662.0 662.0 4.073 4.204 4.410 4.625 4.852 5.089 41 42 43 44 45 392.4 392.4 392.4 392,4 392.4 5.284 5.038 4.802 4.578 4.365 .000555 .000529 .000504 .000480 .000458 .2333 .2225 .2120 2021 .1927 .4830 .4717 4604 .4496 .4390 662.0 662.0 662.0 662.0 662.0 5.338 5.599 5.873 6.161 6.462 46 47 48 49 50 392.4 392,4 392.4 392.4 392.4 4.162 3.967 3.782 3.605 3.438 .000437 .000417 .000397 .000379 .000361 .1838 .1752 .1670 .1592 1518 .4287 .4186 4086 .3990 .3896 662.0 662.0 662.0 662.0 662.0 6.778 7.110 7.459 7.824 8.206 51 52 53 54 55 392.4 392,4 392.4 392.4 392.4 3.276 3.124 2.979 2.840 2.707 .000344 .000328 .000313 .000298 .000284 .1447 .1379 .1315 .1254 .1195 .3804 .3714 .3626 .3541 .3457 662.0 662.0 662.0 662.0 662.0 8.607 9.028 9.470 9.933 10.42 56 57 58 59 60 392.4 392.4 392.4 392.4 392,4 2.581 2.461 2.346 2.236 2.132 .000271 .000258 .000246 .000235 .000224 .1140 ,1087 .1036 .0987 .0941 .3376 .3296 .3219 .3142 .3068 662.0 662.0 662.0 662.0 662,0 10.93 11.46 12.02 12.61 13.23 * ICAN standard conditions ** ICAN standard conditions. For conditions other than ICAN, the speed of sound is propor- tional to VT. A-4 .50 45 40 35 .45 30 25 20 15 10 .40! 5 SEA LEVEL .35 .30 .25 PARAMETERS OF 1,000 FEET Hp .20 .15 100 120 140 160 180 200 220 240 260 280 300 320 340 VCAL IN KNOTS APPENDIX II (1) Graphical Relations Between Calibrated Airspeed, Mach Number and Pressure Altitude for Subsonic Speeds (NACA Standard Atmosphere for Altimeter and Airspeed Indicator Calibration) A-5 .88 PARAMETERS OF Hp 1,000 FEET .85 .80 80 55 .75 40 35 .70 30 25 .65 20 15 .60 10 .33 SEA LEVEL .50 100 120 140 160 180 200 280 300 320 340 360 VCAL IN 220 240 260 IN KNOTS APPENDIX II (2) Graphical Relations Between Calibrated Airspeed, Mach Number and Pressure Altitude for Subsonic Speeds (NACA Standard Atmosphere for Altimeter and Airspeed Indicator Calibration) A-6 1.00 .95 Os .90 .85 20 .80 .75 710 .70 .65 SEA LEVEL .60 PARAMETERS OF Hp 2,000 FEET .55 .50 320 340 360 380 400 420 440 460 480 500 VCAL IN KNOTS APPENDIX II (3) Graphical Relations Between Calibrated Airspeed, Mach Number and Pressure Altitude for Subsonic Speeds (NACA Standard Atmosphere for Altimeter and Airspeed Indicator Calibration) A-7 440 420 50,000 FT. 400 45,000 FT. 40,000 FT. 380 35,000 FT. 360 30,000 FT. X 340 ,25,000 FT. 20,000 FT. 320 15,000 FT. 10,000 FT. V, IN KNOTS 300F 5,000 FT. 1,000 FT. 280 SEA LEVEL 260 240 220 200 STANDARD DAY CONDITIONS 180 160 160 180 200 220 240 260 280 300 320 340 VCAL IN KNOTS APPENDIX III Graphical Relation Between True Airspeed, Altitude and Calibrated Airspeed Under Standard Conditions (NACA Standard Atmosphere) A-8 ! Too I M 7 1+ € m2(2), 8-1.40 То T..-T. E : RECOVERY FACTOR : Ty-Ta - 1.20 E = 1.0 E = E = 0.9 1.16 E = 0.8 E = 0.7 1.12 Te = 0.6 1 € =0.5 ܘܘܟ Ta 1.08 1.04 1.00 .10 .20 .30 .40 .50 .60 .70 .80 .90 1.00 ma M APPENDIX IV Mach Number Temperature Relations for Various Recovery Factors A-9 Zone 150.000 15 50004.h 250 300 350 400 450 500 550 kom 11. RELATIONS BETWEEN TRUE SPEED V, CALIBRATED AIR- SPEED Vcal, MACH NUMBER M, AND PRESSURE ALTI- (SUBSONIC) VALID FOR M€ 1.0 14 Veal = 100 kt TUDE hp 13-1 M-0.60 M=0.50 M=0.40 M10.70 M0.80 40,000) T, CONSTANT. 56.5°C M1.00 M0.90 140,00d 1 12 NOTATION: Di impact pressure. D, (MCO) Distatic pressure cspeed of sound 1) BASIC RELATIONS PD ..(M) (for isentropic 150 kt flow 10 Neal 30,000 9 Vcal: 200 kr 8 • 250 ky Vcol 7 Vcal Vcal. 300 km +50°C (2) M. Vic T.14, R.227 (m2,211°C 30,000 2) CALIBRATED AIR SPEED (1) At eoch altitude Rwas computed 40°C or a function of and M from (1) (2) At eoch value of 0, -, the value of calibrated airspeed is determin ed from: +30°C Pro Subscript, designatos * level conditions popold 3) Under non standard conditions Vis given by the abscind values +20°C corresponding to the intersec- tion of the ordinate with the constant Mach number line which DOIS through the junction of the +10°c curves Ycal and no 20,000 Ncal. 350k1 6 cal*400** NCC 450kt Cal 500k 10,000 3 ON Ncal. 550kt goood qi- toºc 2 Vcal f600 ki Foc M0.30 M: 0.40 M +0.50 M=0.60 M+0.70 M-0.80 M0.90 M=1.00 +20°C 250 300 350 200 400 400 600 450 850 500 950 550 1,000 650 knots 1,200 km/h 450 500 550 600 650 700 750 800 900 1,050 1,100 1,150 APPENDIX V (1) Relations Between True Airspeed V, Calibrated Airspeed Vcal , Mach Number M and Pressure Altitude hp (Supersonic) A-10 1 ܪ 200000 250 300 350 90008 Ap 400 450 550 500 15 kan 11. Veal = 100 kt RELATIONS BETWEEN TRUE SPEED V, CALIBRATED AIR- SPEED Vcal , MACH NUMBER M, AND PRESSURE ALTI- (SUBSONIC) VALID FOR M€ 1.0 14 TUDE hp 13- M = 0.40 M = 0.60 M=0.50 M = 10.70 M = 0.80 40,000 12- M1.00 M +0.90 for Tr. CONSTANT:56.5" 40,00 NOTATION: Di impoct pressure D, (M , static pressure cspeed of sound 1) BASIC RELATIONS 11 10 cal" 150 k1 30,000 9 Vcal: 200 ml 8 Vcal.- 250 ky D. . (M) (for isentropic flow) +50°C (2) M. Vic 7:14R.227 ( m2,21/C 30.00 2) CALIBRATED AIR SPEED (1) At eoch altitude D-, - computed +40°C os a function of Boy and M from (1) (2) Al och value of D-Ps the value of calibrated airspeed is determin ed from: Vcoi -30°C Subscript,o, designates 10 level conditions 20.00 3) Under non standard conditions Vis given by the abscisso values +20°c corresponding to the internec- tion of the ordinate with the constant Moch number line which pane through the junction of the Floºc curvas Vcal and 7 Vcol. 300 kt 20,000 Ncal. 350kt/ 6 09 Ncal 400k/ 5 Ncore 450 kt 4 •500kr 10,000 3 10.000 Neol Near"550kt to°c 2 Vcai 600 ki . 1 FIOC O M=0.30 M = 0.40 M-0.50 M=0.60 M0.70 M20.80 M-0.90 001.no (20°C 300 200 400 250 450 350 650 400 750 450 850 500 950 700 550 1,000 600 1,100 650 knots 1,200 km/h 500 550 600 BOO 900 1,050 1,150 APPENDIX V (1) Relations Between True Airspeed V, Calibrated Airspeed Vcal, Mach Number M and Pressure Altitude hp (Supersonic) A-10 . 1 UNIVERSITY OF MICHIGAN 3 9015 06435 3850 The HF Group Indiana Plant TO