Download as pdf
Download as pdf
You are on page 1of 536
Aeroplane Performance Aeroplane Performance About This Book Important Information Nothing in this book must be taken as superseding the legislation, rules, regulations, notices or procedures contained in applicable national or international regulations, laws and conventions or otherwise published by the appropriate State aviation authorities. Nothing in this book overrides the recommendations, guidance, restrictions or limitations imposed by the training or aviation organization under whose rules you are operating. Nothing in this book overrides the recommendations, guidance, restrictions or limitations imposed by the manufacturer or Design Authority of the aircraft you fly. This book is sold as is without warranty of merchantability and fitness for a particular purpose. Neither the author, the publisher nor their distributors or dealers assume liability for any alleged or actual damages arising from its use. Aeroplane Performance Publication Information EASA ATPL Aeroplane Performance Version 2.7 Published by: Padpilot Ltd Unit 19, The Glenmore Centre, Gloucester, UK, GL2 2AP ISBN 978-1-909600-06-5 © 2012-2020 Padpilot Limited All rights reserved. No part of this book may be reproduced, stored in a retrieval system, or transmitted in any form or by any person or thing without the express written permission of Padpilot Ltd. No part of this publication may be used for commercial purposes without the express permission of Padpilot Ltd. Aeroplane Performance Chapter 1 - Performance Fundamentals Chapter 2 - Lift, Weight De Chapter 3 - Climbing, Descending and Level Flight Chapter 4 - Thmust and Power = Chapter.5 - Take-Off - General Principles Chapter 6 - Angle of Climb Chapter 7 - Rate of Climb Chapter 8 - Climb Speeds and Calculations Chapter 9 - The Cruise Chapter 10 - Range Chapter 11 - Endurance Aeroplane Performance Chapter 12 - The Descent Chapter 13 - Landing i Chapter 14 - Class B Regulations Chapter 15 - Sel Engine Piston - SEP1 ClHapter 16 - Multi Engine Piston - vas Chapter 17%-.Class A Take-Off Theory Chapter 18 - Aerodrome Distances Chapter 19 - MRUT Take-Off Chapter 20 - Additional Take-Off Techniques Chapter 21 - The Initial Take-Off Climb Chapter 22 - MRJT En-Route Chapter 23 - MRJT Landing - Performance Fundamentals roe CoM Lael] iS Clatrehaiul ss SEE Section 1 Performance and as Pilot 1.1.1 Performance and the Pilot THREAT On the 6” March 2003 an Air Algerie Boeing 737 began its take-off run from Tamanrasset Airport in Algeria. The take-off was uneventful, until an engine failed at a height of 78 ft. This was way past the most critical point (the decision point). As a result of poor crew teamwork and inadequate speed control, the aeroplane stalled and crashed 1645 m past the end of the runway, killing all but one of the 103 Passengers and crew. An engine failure on take-off is the one most tightly considered contingencies within performance. Even with the engine failing at the most critical point on the runway, the aeroplane should have cleared all obstacles after take-off by at least 35 ft, provided the crew had performed at least averagely. At this point you are likely to be thinking that, during your flying career, you will never perform this badly. In fact you're probably thinking that you are most likely to perform at an above average level. The trouble is that almost everyone reading this will believe that he or she is likely to perform above the average. Clearly this can’t be true for all of you. It’s much more likely that you are, and will only ever be, an average pilot! But even someone with average piloting skills can keep themselves well into the safer half of average through knowledge, average skill and practice. CHAPTER 1 Performance Fundamentals The Air Algerie accident is alarming enough in its implications, but if we look further into performance accidents and incidents an even more extraordinary fact emerges: many performance related accidents involve no engine or major system failure at all! ERROR On the 14th October 2004 a cargo Boeing 747 began its take-off run at Halifax International Airport in Canada on a reduced thrust setting. At the planned rotate speed of 130 kt the aeroplane started to rotate but, despite increasing speed, and a further 2700 ft of runway, the aeroplane didn’t lift-off. In the last 600 ft of runway, the crew increased thrust to 92% of maximum but even this wasn’t enough to prevent the aircraft from rolling off the end of the runway at 155 kt. to critical underperformance After a further 200 m of cross-country excursion, it eventually managed to struggle into the air, but hit the ground again 100 m later, after which it broke up and burst into flames. All 7 crew members died The accident was caused by a miscalculation. The crew had used a take-off mass of 240 000 kg, but the actual mass was 353 000 kg. The consequent reduced thrust take-off and the calculated rotate speed were completely inadequate. A fact made more surprising because the crew knew they were close to being fully loaded. Unfortunately calculations errors such as this lead to incidents or near incidents, all too often. On the 20" March 2009, an Airbus A340 started its take-off run from Melbourne. It was 100 000 kg heavier than its planned weight. When the aeroplane refused to rotate the captain reacted quickly, selecting maximum take-off thrust. Although a serious accident was avoided, the following tail strike caused substantial damage to the aircraft and the airport ILS. CHAPTER 1 Performance Fundamentals Another all too common performance mistake occurs when a take-off is planned for a high flap setting to reduce the take-off speed but the crew (potentially you) mistakenly selects a lower flap setting for the take-off. On 5" September 2005, 100 passengers and crew of a Boeing 737 lost their lives in Indonesia precisely because of this mistake Finally, all the accidents and incidents we have looked at have been at, or immediately following, take-off. The table below, published annually by Boeing, summarises all fatal aeroplane accidents with a mass exceeding 60 000 Ib, but excluding military aeroplanes and those manufactured in CIS and USSR. Between 2006 and 2015, 47% of all fatal accidents occurred during the landing or on the final approach. ‘ax, oat! 40% triad partes Inst) cimb Intel 7 Fa tow | Takeo | cmb | (tapeup) Descent | sppreach | approach Feta Figure 1.1 Percentage of fatal accidents and onboard fatalities by phase of flight 2006-2015 (Boeing Statistical Survey 2018) Section 2 Introduction to Performance 1.2.1 Introduction to Performance Aeroplane performance centres on providing a statistical level of safety. In this section we'll look at the regulations and requirements that provide the required margins of safety. We finish the chapter by introducing aerodrome descriptions and requirements. Defined terms are described in the text and, for ease of reference, we have also included them in the glossary. Aircraft Certification Within the European Union (EU) the European Aviation Safety Agency (EASA) is responsible for regulating the airworthiness requirements for new aeroplanes. Part of these requirements is to state minimum performance criteria. To allow for the considerable difference in performance, two sets of performance regulations exist: CS 25 (Certification Specification 25) covers larger, better performing aeroplanes ™ CS 23 (Certification Specification 23) covers poorer performing, and usually smaller, piston propeller aeroplanes. To determine if a new aeroplane meets these minimum performance standards and other certification requirements, it must be extensively flight tested. The flight test data is also used to produce the performance section of the Performance Manual or Operating Data Manual (ODM), which supports the many performance decisions the flight crew make on every flight. CHAPTER 1 Performance Fundamentals 1.2.2 Aeroplane Performance Classes Under EASA, aeroplanes are grouped into 3 Performance Classes: A, B, and C. Ie Ne Capacity ae Rass CHAPTER 1 Performance Fundamentals Performance Class A Aeroplanes Class A includes all multi-engine jet aeroplanes and multi-engine turbo propeller aeroplanes with ten or more passenger seats and/or a maximum take-off mass (MTOM) exceeding 5700 kg. | Figure 1.2 The Airbus A350 is a Performance Class A aircraft Performance Class A aeroplanes have the most stringent performance requirements; these are documented in CS 25. CHAPTER 1 Performance Fundamentals Performance Class B Aeroplanes. Class B includes all propeller-driven aeroplanes having nine or less passenger seats and a MTOM of 5700 kg or less. Figure 1.3 Class 8 aircraft have 9 or fewer pax seats AND an MTOM less than 5700 kg The class includes single and twin-engine aeroplanes. The performance requirements, which are much less stringent than Class A, are detailed in CS 23. CHAPTER 1 Performance Fundamentals Performance Class C Aeroplanes Class C includes large piston-powered multi-engine aeroplanes (for example the old DC3 Dakota). They must have either ten or more passenger seats and/or have a maximum take-off mass exceeding 5700 kg. There are few aeroplanes in this class still flying. Class C aeroplane performance is not considered further in this book. Section 3 Performance Legislation 1.3.1 Performance Legislation The airworthiness requirements legislated in CS 23 and 25 state the appropriate minimum acceptable performance levels for each stage of flight. However to ensure similar safety standards between the performance classes, a second layer of legislation, operating regulations, is needed. The operating regulations are found in IR- OPS. Performance Class A aeroplanes have the most demanding airworthiness requirements and, therefore, the /east stringent operating regulations. Class A aeroplanes are allowed to operate in Poor weather conditions and to use contaminated runways. Conversely, single-engine Class B aeroplanes have the least stringent airworthiness requirements and therefore the most stringent operating requirements. This prohibits them from public transport operations at night or in IMC. CHAPTER 1 Performance Fundamentals Determination of Aeroplane Performance One of the key concepts underlying the regulations is planned performance. The purpose of planned performance is to determine the maximum aeroplane weight at different stages of flight to ensure a predetermined level of safety. For this to be possible, empirical data relating to the aeroplane’s performance is required. This establishes both the average and the spread of: @ Distances used for take-off and landing @ Rates and gradients of climb and descent. Performance figures, such as take-off distance required and climb gradient, are described as Measured, Gross, or Net performance. It is absolutely essential that you understand the meaning of and relationship between Measured, Gross and Net performance. Measured Performance Before a new aircraft type can enter service, the manufacturer must measure the performance of the pre-production aircraft. Measured data is collected for distances, gradients and rates of climb and descent in a variety of configurations of flap, slats and landing gear - with all engines working and, if applicable, with one engine out. The average of each set of data, achieved by the new aeroplane when flown by test pilots, is known as Measured Performance. CHAPTER 1 Performance Fundamentals This graph plots the actual climb paths achieved after take-off by a test pilot under given conditions. In this case, one engine failed. 400 8 8 8 x 8 8 climb paths flown by company test pilots Height (ft) above reference zero ° 1000 2000 3000 4000 5000 Distance (ft) from the the end of take-off reference zero Figure 1.4 Measured performance 1 engine failed The average of the recorded climb paths is calculated and is shown on the diagram by the blue climb path labelled Measured Performance. However, Measured Performance is the average for the aeroplane. It is also the performance achieved by experienced and highly skilled test pilots. To establish figures that represent an average for an aeroplane within a fleet flown by average line pilots, the measured performance must be degraded. CHAPTER 1 Performance Fundamentals Gross Performance Gross Performance is the average performance that a fleet of aircraft can be expected to achieve, when satisfactorily maintained and flown in accordance with the techniques described in the flight manual. Gross Performance is established for all engines working (the normal situation) and, for some aeroplanes, one-engine out - throughout a range of aircraft configurations and environmental conditions. The graph below shows that Gross Performance is worse than Measured Performance. 400 8 8 8 x 8 8 Recorded climb paths flown by ‘company test pilots 8 Height (ft) above reference zero & = O 1000 2000 3000 4000 5000 Distance (ft) from the the end of take-off reference zero Figure 1.5 Gross performance Gross Performance is very useful to the pilot because it states how the aeroplane should perform on an average day. In other words it tells you where the aeroplane is likely to be after take-off, or how much runway on average you will use. CHAPTER 1 Performance Fundamentals However, because Gross Performance is an average figure this means that half the time the aeroplane will perform better (climb steeper or use less runway) and half the time worse (use more runway and climb at a shallower angle). Although Gross Performance tells you how the aeroplane is likely to perform on an average day, some to almost all of the variation from flight to flight, must be allowed for to guarantee a suitable safety margin. This is achieved by using a safety factor. Gross Performance is degraded by the application of the safety factor, to give Net Performance. CHAPTER 1 Performance Fundamentals Net Performance In public transport operations, Net Performance is used to determine the safety related performance limiting masses (for example; determining the maximum mass that allows for adequate obstacle clearance and determining the maximum take-off and landing masses within the runway distances available). Net Performance is the Gross Performance diminished by a safety factor or margin, laid down by the appropriate authority. The graph below shows that the Net climb path is shallower than the Gross climb path. 400 8 38 8 y 8 8 Recorded climb paths flown by company test pilots 3 8 Height (ft) above reference zero a g 0 1000 2000 3000 4000 5000 Distance (ft) from the the end of take-off reference zero Figure 1.6 Net performance The Net performance climb path is calculated to allow the aircraft to clear obstacles by at least 35 ft in the initial climb. So, if your aircraft achieves Gross performance, your actual obstacle clearance will somewhat be greater than 35 ft. CHAPTER 1 Performance Fundamentals Now look at the lowest (worst) climb gradients recorded in the previous graph. Notice that if your climb path is not Gross (average) but is actually one of the worst performances in the recorded range you may not have 35 ft obstacle clearance. Let's summarise Measured, Gross and Net performance. ™ Measured Performance is the average performance for the new aeroplane flown by test pilots. ™ Gross Performance is the average performance for the fleet flown by line pilots using standard techniques. ™@ Net Performance is the performance used to determine the safety related performance-limiting masses. Net performance is the gross performance diminished by a safety factor. CHAPTER 1 Performance Fundamentals How the Size of the Safety Factor is Determined An aeroplane could perform better or worse than Gross on a particular flight, so a safety margin must be applied to establish Net Performance. In general, the crew, passengers and the general public would prefer large safety factors. The greater the size of the safety factor for a particular phase of flight, the lower the risk of an accident. However, larger safety factors result in lighter take-off masses and therefore less revenue-earning load. Performance regulations must therefore balance the conflicting demands of safety and profitability, when determining the size of the safety margin. Consequently the safety margin required for public transport, is based on the probability of an event/incident/accident being less than one in one million (10°) when the passenger buys his ticket. Statistically this is known as a ‘remote probability’. To achieve this level of safety, both the likelihood of the situation and the range of the performance figures recorded must be considered. CHAPTER 1 Performance Fundamentals Prato fae eur.) Up to 1/1000 1/1000 to 1/100 000 Engine failure Low speed overrun panennen Failure to achieve net take-off 1710000000 fier natn High speed overrun Ditching 4/10000000to Double engine failure on a twin 1/1000 000 000 engined aircraft Hitting an obstacle in the net take-off flight path Paneer Section 4 Likelihood of Event Occurring 1.4.1 Likelihood of an Event Occurring The graph below shows the likelihood of a situation occurring for a particular aeroplane. You can see that, once you are airborne, the likelihood of landing is certain (e.g. 1:1) because every flight must end in a landing! The probability of take-off however, is slightly less than certain, because some take-offs will be rejected due to problems occurring early in the take-off run. Take-Off sat 120429 AllEngines g 3 3 5 8 g 8 g 5 a Engine Fail = som : EnRoute 3 Bt a & 1:100.000 Engine Fail iat Cm & 1:1900 000 us 8 5 é F 2x Engine Fait oon Take-Off 41:100 000 000 | Figure 1.7 Likelihood of an event The probability of an engine failure at V, (the most critical point on take-off) is shown as 1:1 000 000. However, the probability increases to more frequently than 1 in 1 million, when the later part of the take- off run and initial climb are also considered. 24 CHAPTER 1 Performance Fundamentals An engine failure at some point en route is shown as being between 1:10 000 and 1:100 000. A double engine failure on take-off is shown as 1:100 000 000, which means that it is extremely improbable. The following graph also shows a horizontal red dashed line at 1:1 000 000 (1 in 1 million) which is the minimum level of safety that is required for public transport. Because the likelihood of a double engine failure on take-off is much less than 1 in 1 million (below the red line), there is no safety factor (and in this case, no need to demonstrate that the aeroplane could continue the take-off). However in situations having a likelihood of occurrence greater than 1 in 1 million, a safety factor is required. This is shown by the one engine failed take-off situation, where a small safety factor is applied for the initial climb, but not (just) for the actual take-off. A safety factor is required for the en-route engine failure situation and large safety factors are required for the certain landing for the (near certain) all-engine take-off situations. This is important so it’s worth recapping again. Safety factors are required when the probability of an event is greater than 1 in 1 million, and The greater the probability of an event (greater than 1 in 1 million), the larger the safety factor required. CHAPTER 1 Performance Fundamentals Range of Performance Data The size of the actual safety factor is also affected by the range of the performance data. We've seen that the landing and normal all- engine take-off situations, are at or close to being certain (1:1), so we can expect to see correspondingly large safety factors. This however is not the case. The safety factor for landing (67% for a jet) is much larger than for the all-engine take-off safety factor (15%). This is due to the much wider range of landing distances, compared to the established range of take-off distances. Gross Net Distance Distance > g 3| ° , Satey factor OOM Distance 15% t Gross Net Distance Distance Frequency ° Landing Distance Safety factor 67% Figure 1.8 Safety Factors 26 CHAPTER 1 Performance Fundamentals The wider range of landing distances is due to acceptable small variations in speed and height at the landing screen and variations in actual landing technique. By contrast, the range of normal take-off distances are much narrower. This is because the brakes release take-off position is accurately known, pilots will normally initiate rotation within a knot or so of the rotate speed and rates of rotation between pilots are very similar. In summary, the safety factor (and therefore the difference between gross and net performance) will be larger when: ™ The probability of an event is greater, and/or @ The range of the acceptable performance achieved is wider. meer Section 5 Comparing Gross to Net 1.5.1 Comparing Gross to Net We will finish this section of the chapter by considering the relationship between Gross and Net performance for take-off and landing distances and climb and descent gradients. Gross and Net distances The following image shows the distance used by an aeroplane landing. The green line shows where the aeroplane would stop on an average day. The red line shows where an identical aeroplane would stop on a one-in-one-million ‘bad’ day. ten! Oe Gross landing distance Net landing distance 7 Figure 1.9 Gross and Net landing distance CHAPTER 1 Performance Fundamentals Note that the Net landing distance is much longer than the gross landing distance. For public transport, the runway distance available must be at least as long as the Net landing distance required. Net distances are longer than gross distances (for take-off and landing). Climb Gradients Versus Descent Gradients Let's look again at the take-off initial climb situation. Normally an engine won't fail so normally the aircraft will climb steeply at the all- engine Gross gradient. 400 300 200 Recorded climb paths flown by company test pilots Height (ft) above reference zero - ° 1000 2000 3000 4000-5000 Distance (ft) from the the end of take-off reference zero Figure 1.10 Comparing Net and Gross climb gradients On the rare day that an engine fails, the aeroplane should achieve on average the one-engine failed Gross climb path. CHAPTER 1 Performance Fundamentals However, on a ‘one-in-a-million bad day’, when an engine failure has occurred, AND the aeroplane is performing slightly worse than the engine-out climb Gross gradient, the aeroplane would climb at the shallower Net climb path shown in red. This is why Net gradient is used to assess obstacle clearance. However if the aeroplane performs worse than this, the achieved Net climb gradients are smaller than gross climb gradients. gradient would be shallower than the Net climb gradient. In such circumstances obstacle clearance would not be guaranteed. Net descent gradients are greater than gross descent gradients However, in the one-engine out descent situation, the Net descent gradient (which provides minimum obstacle clearance) is steeper than the Gross descent gradient, as is shown in the following diagram. CHAPTER 1 Performance Fundamentals Figure 1.11 Comparing Net and Gross descent gradients 31 Paneer Section 6 Aerodromes 1.6.1 Aerodromes Whether or not an aircraft can operate at an aerodrome is determined by a number of factors including: the size of manoeuvring areas, the length of the runway and the strength of the manoeuvring surfaces. Many of these aspects are covered in greater depth in Operational Procedures. For this subject we need to focus on those aerodrome characteristics that affect the take-off and landing. Consequently, it’s very important that you understand the various lengths and distances described. This is our final task for this chapter. However, before we can understand the various performance distances available it will be helpful to know the runway end markings in common use. 32 CHAPTER 1 Performance Fundamentals 1.6.2 Runway End Markings Threshold The beginning of that portion of the runway available for landing is known as the threshold. This is indicated by a solid white line across the width of the runway. A series of white lines immediately after the threshold, and parallel to the runway centerline, indicates the width of the runway. These are known as the piano keys. A numerical, or alpha-numerical, runway designator is marked after the piano keys. roam Figure 1.12 Runway end markings - threshold and runway designator 33 CHAPTER 1 Performance Fundamentals Displaced Threshold A threshold may be permanently displaced or temporarily displaced, shown in the image below on the left and right respectively. A displaced threshold may have an area before the threshold which can be used for taxiing or for the initial acceleration at the start of take-off run. It is never suitable for landing, hence the reason why the threshold is displaced. During landing this area may also be suitable for the ground roll at the end of a landing. This area is marked by arrows pointing towards the displaced el IBAAAA ama Figure 1.13 Runway end markings: displaced threshold 34 CHAPTER 1 Performance Fundamentals If the area before a permanently displaced threshold is not fit for manoeuvring under any circumstances, it will be marked with white crosses. If the area can be used, but only as a stopway, it will be marked by a series of yellow chevrons. rea roam 1 Figure 1.14 Runway end markings: no normal movement areas 35 CHAPTER 1 Performance Fundamentals Stopway The stopway is an obstacle-free rectangular area on the ground at the end of the runway. It is at /east the same width as the associated runway and centred along the runway’s centreline. The stopway must be prepared so that the aircraft can roll over it without hazard to it or its occupants. It doesn’t need to be as well maintained, or as strong as, the associated runway. In some cases the stopway may comprise, for example, a section of disused runway. A stopway is clearly marked with yellow chevrons and has no runway edge marking. Figure 1.15 Runway showing stopway and displaced threshold at Palm Springs International Airport (Image courtesy: D Ramey Logan) 36 eer CHAPTER 1 Performance Fundamentals 1.6.3 Take-Off Distances Take-off Run Available (TORA) Take-off run available (TORA) is the distance between the point on the surface of the aerodrome at which an aeroplane can begin its take-off run to the nearest point, in the direction of take-off, at which the surface of the aerodrome is incapable of bearing the weight of the aeroplane under normal operating conditions. The TORA often (but not always) corresponds to the physical length of the prepared and maintained runway pavement. The TORA may also include a displaced threshold but only that portion of the runway indicated by white arrows may be used for take-off. Areas marked by white crosses or yellow chevrons cannot be included. The TORA may also include the displaced threshold area at the departure end of the runway. TORA Displaced Threshold Displaced Threshold Figure 1.16 Take-off run available (TORA) 37 CHAPTER 1 Performance Fundamentals Clearway The clearway is an area that may be provided at the end of the TORA, in the direction of take-off, which is free from any obstacles that could cause a hazard to aeroplanes in flight. The clearway extends 75 m either side of the extended runway centreline. Lengthwise, it extends to the first non-frangible obstacle. (Frangible means easily breakable). In situations where no obstacles exist for a considerable distance from the departure end of the runway, the clearway’s length is restricted to a maximum of 50% of the TORA. Clearway TORA : i ' ' , ' ' \ ' ‘ ' 1 ' TODA i First Signieant Figure 1.17 The Cleanway Obstacle CHAPTER 1 Performance Fundamentals Because a clearway is only used for safe overflight (and not for the ground roll) it may comprise ground or water (provided it is under the control of the appropriate authority). The clearway may also extend over certain marked areas of the paved surface, i.e. the stopway. Stopway Clearway Figure 1.18 The Clearway over paved area The clearway exists beneath an imaginary climbing surface that starts at ground level at the end of TORA and then slopes upward at a gradient of 1.25% (0.75°) away from the runway. Within the clearway, no part of the ground or any objects (exceeding 0.9 m) may protrude through this surface. CHAPTER 1 Performance Fundamentals Take-Off Distance Available (TODA) The take-off distance available is normally the runway plus the clearway. Because of the 50% of TORA limit mentioned earlier, the TODA is therefore the lesser of: mi TORA plus clearway, or Clearway TORA == ae A TODA First significant obstacle Figure 1.19 Take-off Distance Available (TODA) ™ 1.5 times TORA Clearway (50% of TORA) TORA TODA (1.5 x TORA) First significant obstacle Figure 1.20 Alternative calculation of TODA CHAPTER 1 Performance Fundamentals Example 1 If the TORA is 2000 m and the measured clearway is 1300 m, what is the TODA? Solution: TODA is the lesser of: mTORA + clearway (2000 m + 1300 m) = 3300 m, and 1.5 x TORA (1.5 x 2000 m) = 3000 m. Answer: TODA = 3000 m Example 2 A runway’s TORA is 2200 m and the measured clearway is 1000 m, what is the TODA? Solution: TODA is the lesser of: mTORA + clearway (2200 m + 1000 m) = 3200 m, and 1.5 x TORA (1.5 x 2200 m) = 3300 m. Answer: TODA = 3200 m CHAPTER 1 Performance Fundamentals Acceleration Stop Distance Available (ASDA) The acceleration stop distance available (ASDA) is the distance from the point on the surface of the aerodrome at which an aeroplane begins its take-off roll to the nearest point in the direction of take-off at which it can’t roll over the surface of the aerodrome and be brought to rest in an emergency without risk of accident. Stopway TORA ASDA Figure 1.21 ASDA equals TORA plus Stopway ASDA equals TORA plus stopway. It used to be known as emergency distance available (EMDA), a term still used occasionally in aviation and in exams. CHAPTER 1 Performance Fundamentals Balanced Field Balanced field refers to a situation in which TODA equals ASDA. This means that any clearway equals the length of the stopway. A balanced field is useful because it simplifies performance calculations about the maximum field length for a given take-off mass. TORA 1 ASDA 1 TODA Figure 1.22 The Balanced Field Runways which are not balanced, because the clearway exceeds the stopway (or vice versa), can be balanced by ignoring the additional distance. This speeds up the performance calculations, with a potential small reduction in the runway limited take-off mass. CHAPTER 1 Performance Fundamentals TORA, ASDA and TODA The diagram below shows the complete situation. This runway has displaced thresholds, and stopways and clearways in both directions. Note that the field is not balanced. +++ Wind Direction Clearway Stopway Stopway Cleanway Displaced Threshold TORA ASDA TODA Figure 1.23 Runway end markings CHAPTER 1 Performance Fundamentals 1.6.4 Pavement and Aircraft Classification Numbers Even if a runway is long enough to land on, it may not be strong enough to support the weight of an aircraft; or, perhaps the runway is strong enough but the other manoeuvring areas, such as the taxiway and apron, are not. Consequently ICAO has developed the aircraft classification number (ACN) and pavement classification number (PCN) system as a standard method for reporting the bearing strength of pavements. Aircraft classification numbers give a relative load rating of the aircraft on pavements for certain specified sub-grade strengths. ACN values for most aeroplanes have been calculated by ICAO and are published in aeronautical information publications. Figure 1.24 Avery different kind of load bearing surface! CHAPTER 1 Performance Fundamentals The pavement classification number gives the load bearing strength of the pavement as the highest aircraft classification number that it can accept for unrestricted use. So, if the runway’s PCN is lower than your aircraft’s ACN, then you can’t use it. A PCN is reported in a five-part format. Apart from the numerical value, notification is also required of the pavement type (rigid or flexible) and the sub-grade support category. Additionally, the aerodrome authority can specify a maximum allowable tyre pressure. The final part of the PCN indicates whether the assessment was made from a formal technical evaluation or from past experience of aircraft using the pavement. Example: PCN80/R/B/W/T The PCN is 80. The pavement is rigid, medium strength and there is no tyre pressure limitation. It was assessed by technical evaluation. If your ACN is 80 or lower, you can use it. ACN/PCN and Overload Operations For a limited number of movements, an aeroplane with an ACN slightly greater than the PCN, may be permitted to operate at an aerodrome, provided permission has been granted from that aerodrome’s operating management. yd Speed, Lift, Weight and Drag Ere Section 1 Speed, Lift, Weight and Drag 2.1.1 Speed, Lift, Weight and Drag Pressure The starting point when trying to understand and measure aircraft speed is air pressure. This is because some measures of aircraft speed, although called “airspeed” are actually nothing more than measures of air pressure. Pressure is the force exerted over an area: Force Area Pressure = The relevant international standard units are N/m’, pascals, or millibars. 48 CHAPTER 2 Speed, Lift, Weight and Drag Static Pressure The atmosphere comprises molecules, which have mass and are attracted by gravity to the Earth. The lower you are in the atmosphere, the greater the mass of atmosphere above you and therefore the weight of air pressing down on you. However, this weight force is not just pressing downwards, it exerts its force in all directions. Consequently, any other object in the atmosphere experiences an equal static pressure from all directions.. 49 CHAPTER 2 Speed, Lift, Weight and Drag Dynamic Pressure Dynamic pressure is generated whenever there is relative movement between an object and the air around it. Dynamic pressure only acts in one direction, depending on the direction of motion. For example, when you hold your hand out of a moving car window, you can feel dynamic pressure pushing your hand backwards. Static pressure is also acting on your hand, but because this acts equally from all directions you can’t sense it. The only pressure you feel is the unbalanced force created by dynamic pressure. Dynamic pressure increases when the car increases speed. This is because the kinetic energy of the air acting on your hand is increasing. Note: Actually, in this case it is the kinetic energy of your hand which has increased. However, because all forces have equal and opposite reactions it is perfectly reasonably to understand the forces in terms of the kinetic energy of the air. This is the traditional method for aerodynamic explanations because aerodynamicists work with wind tunnels in which the model aircraft stays still whilst the air is accelerated. Dynamic pressure increases with relative speed and increases with air density. The kinetic energy of the air depends very much on the air’s density. Expressed as a formula: dynamic pressure = ¥2 p V? The above equation shows that dynamic pressure is proportional to air density (p) and proportional to true airspeed (V) squared. Dynamic pressure (2 p V’) is common in all aerodynamic forces. CHAPTER 2 Speed, Lift, Weight and Drag Total Pressure Total pressure is the sum of dynamic pressure and static pressure Total pressure = Dynamic pressure + Static pressure Total pressure around a body at a given speed and air density remains constant provided energy doesn’t change from one form to another - such as heat or sound. This simple statement has an important consequence because we can now say that if the dynamic pressure increases when it flows around an object there must be a corresponding decrease in the static pressure. The diagram below shows that the area of lowest pressure above a wing occurs in the region where the local air flow is fastest. Highest velocity Figure 2.2 The lowest static pressure occurs where the speed of flow is highest eer) Section 2 Measuring and Defining Speed 2.2.1 Measurement of Speed Air Speed True air speed (TAS) is the speed of an aircraft relative to the air mass through which it is traveling. Although TAS is a useful measure of speed for navigation purposes, it does not directly relate to the aerodynamic performance of the aircraft. Consequently, most of the time we use other measures of air speed, which are closely related to dynamic air pressure. Dynamic air pressure is the additional air pressure in a volume of air which is attributable to the air’s kinetic energy. Air speeds relating to dynamic pressure are grouped into a family of indicated air speeds. Before we look at the various indicated air speeds in detail, we first need to look briefly at how an air speed indicator senses dynamic pressure to measure air speed. Isolating and Measuring Dynamic Pressure Total air pressure comprises static pressure and dynamic pressure. So, to measure dynamic pressure only, an air speed indicator (ASI) must somehow subtract static pressure from total pressure. 52 CHAPTER 2 Speed, Lift, Weight and Drag ommend —= Total erates pitot o——> pressure —

+ Direction of flight Figure 2.3 A rudimentary method for measuring air speed The very simplest way to isolate dynamic pressure from static pressure (and thus obtain an approximate indicated air speed), is to fix a flexible strip in the air flow. The flexible strip bends as airspeed increases; the greater the speed, the greater the degree of bend. A simple calibrated scale could then tell us our airspeed. The flexible strip is exposed to static pressure on both sides, but it is only exposed to dynamic pressure on the side facing the air flow. Consequently, the amount of bend is caused solely by the effect of dynamic pressure. CHAPTER 2 Speed, Lift, Weight and Drag 2.2.2 The Traditional Air Speed Indicator (ASI) A traditional ASI uses the same principle to sense and compare static and total pressure. A small sample of air is introduced into an otherwise closed flexible capsule via a pitot tube. This sample comprises dynamic and static pressure. Simultaneously, static pressure from two or more static ports is fed into the sealed case of the instrument. An increase in pitot pressure inside the capsule tends to cause it to expand. An increase in static pressure outside the capsule tends to cause it to contract. Ultimately the capsule only expands or contracts when there is a difference between pitot and static pressure. A suitable system of gears and levers translates the expansion and contraction of the capsule into rotary movement of a needle on the face of the ASI. Dynamic lus static pressure Slatc eg) pressure Movie 2.1 A traditional air speed indicator (ASI) CHAPTER 2 Speed, Lift, Weight and Drag Because the ASI uses both pitot and static sources it can be said to be part of the pitot-static system. Static Pressure Sources Static pressure is sensed by a pair of static pressure ports, one on each side of the fuselage. The ports are set in an aerodynamically quiet area and are oriented exactly perpendicular to the local airflow over that part of the fuselage. The ports are electrically heated to prevent icing. Pitot Pressure Sources The pitot tube points directly into the air flow. When the aircraft moves forward, air is forced into the pitot tube. Because it has no exit (other than a tube leading to the ASI capsule) all the air entering the pitot tube is brought to a complete standstill. The kinetic energy of the flow is thus converted to pressure energy which adds to the static pressure already present. This produces the stagnation pressure or pitot pressure which is a reasonably accurate measure of the airflow’s total pressure. Dynamic pressure + static pressure = Total pressure CHAPTER 2 Speed, Lift, Weight and Drag 2.2.3 Speeds Obtained from Dynamic Pressure Indicated Air Speed (IAS) Indicated air speed (IAS) is the speed indicated on an elementary air speed indicator (ASI). IAS is closely related to, and an approximate measure of, dynamic pressure. Ignoring minor errors associated with the IAS we can say that: m Dynamic pressure (q) « IAS* Taking the square root of both sides: m IAS =” Dynamic pressure (q) Because IAS is a measure of dynamic pressure we can say that: mIAS’ « q = %p TAS This equation tells us that if an aeroplane is to produce a constant amount of aerodynamic force, as altitude increases and air density reduces the aeroplane’s TAS must increase. CHAPTER 2 Speed, Lift, Weight and Drag Indicated Air Speed (IAS) Errors On a traditional ASI, IAS is not a particularly accurate measure of speed because of 3 categories of error which are inherent in all basic air speed indicators: ™ Instrument Error. The instrument itself suffers from small errors caused by the limitations of the manufacturing process. ™ Pressure Error. Errors in the measurement of both total and static pressure can occur due to the position and orientation of the pitot tube and static ports relative to the airflow. For example, a pitot tube which isn’t exactly aligned into the airflow will not sense full dynamic pressure. Similarly, static ports can be influenced by local changes to static pressured caused by events such as lowering flap. ® Compressibility Error. At high speed, air entering the pitot tube becomes compressed, which artificially raises its density and thus induces errors in the measurement of total pressure. In practice we can correct for almost all of these errors. Depending on which errors have been corrected, we arrive at two other types of air speed: ™@ Calibrated air speed @ True air speed Advanced Sensors Modern airliners don’t use such a basic system to find air speed. Pitot and static pressure is converted to digital signals at the sensing units. These signal are transmitted to an air data computer which calculates air speed, taking into account all the factors, such as compressibility, that may affect the sensed pressure. The result is shown on display units. The 3 errors described above are largely, but not entirely, eliminated in a modern system. CHAPTER 2 Speed, Lift, Weight and Drag Calibrated Air Speed (CAS) Calibrated air speed is IAS corrected for instrument error and predictable pressure errors. These corrections are usually small. In modern airliners they are normally removed by an air data computer. Consequently, unlike a basic flying trainer with its simple air speed indicator displaying IAS, most airliners display CAS. Nevertheless, pilots still usually talk about TAS even though their instrument is showing them CAS. In this book we will use the same convention and use the term “IAS” to mean IAS ona light aeroplane and CAS on more advanced aeroplanes. Equivalent Air Speed (EAS) Equivalent air speed is CAS corrected for compressibility and thus provides the most accurate indication of dynamic pressure. Consequently it is most accurate to revise our earlier equation to say that: EAS? x q = ‘29 TAS? At slower speeds IAS, CAS and EAS will either be the same or differ by only a few knots. At higher speed, compressibility error produces a significant difference between CAS and EAS. However the majority of this book assumes low speed flight, in which the air is considered to be incompressible. Consequently we will not often need to draw a distinction between IAS, CAS and EAS. CHAPTER 2 Speed, Lift, Weight and Drag True Air Speed (TAS) True air speed (TAS) is the actual speed of the aeroplane relative to the air mass. After taking into account the wind velocity it can be used to determine the aeroplane’s ground speed (GS), which, as suggested, is the speed over the ground. You may sometimes see the relationship between EAS and TAS written as an equation (below) where the air density at the aeroplane is p and the air density at mean sea level is ro. EAS = TAS Vp/t. At mean sea level in ISA conditions, EAS = TAS. But at higher altitudes where the air density is less, EAS/CAS/IAS will be less than the TAS. 2.2.4 Mach Number (M) Mach number is not a speed but a ratio of TAS to the local speed of sound (LSS). TAS Mach ber = —— lach number iss LSS varies with air temperature. As altitude increases, the air temperature normally decreases resulting in LSS decreasing. This means that, if an aircraft climbs at constant Mach number, which is normal at higher altitudes, the TAS (and CAS) must decrease. TAS Mach ber = ——— lach number iss In jet transport performance, Mach numbers are used enroute and during the upper parts of the climb and descent due to flight Mach number limitations. Ere Section 3 Forces Acting On An Aircraft 2.3.1 The Forces Acting on an Aircraft During flight, the forces acting on an aircraft are weight, lift, drag and (in powered flight) thrust. Thrust is examined more closely in Chapter 4. For now we will concentrate on weight, lift, and drag. 60 nr Beer CHAPTER 2 Speed, Lift, Weight and Drag Weight An aircraft has mass, which, when acted on by gravity, produces the force weight. The weight of an aircraft always acts vertically down from the aircraft's centre of gravity. Centre of gravity - sessecesecccssccesh ts sersnnscsiWVeccassesseee a W6IEM Figure 2.4 weight The acceleration due to gravity, ‘g’, is assumed to be constant at 9.81 m/s’, which is often given as a rounded figure of 10 m/s’ in exam questions. Weight is measured in Newtons (N) whereas mass is measured in kilogrammes (kg). On planet Earth, an aircraft mass of 60 000 kg produces a weight force of 60 000 kg x 9.81 m/s’ = 588 600 N. EOE CHAPTER 2 Speed, Lift, Weight and Drag Lift The aircraft is moving through the air; consequently the air is flowing along and past all of its surfaces, but in the opposite direction. This is called the relative airflow. Lift Drag Centre of pressure Figure 2.5 Lift and Drag The relative airflow is always along, and in the opposite direction to, the aircraft’s flight path. The action of the relative air flow on the wing (of area S) produces an aerodynamic force, which is shown in vector terms as a total reaction or net aerodynamic force on the wing. This acts through the centre of pressure. CHAPTER 2 Speed, Lift, Weight and Drag Lift is the component of the total reaction that is perpendicular to the relative airflow. The size of the lift force depends on the IAS squared, the wing area, and the wing's coefficient of lift. This is shown in the equations below. Since IAS* x ¥2 pV" (where V is TAS) = dynamic pressure: L« IAS?°SC, or L=*%pv sc, The lift equation is fundamental to an understanding of the principles of flight and performance. Note the effect of ‘squaring’ the airspeed; if IAS doubles, with no change in the coefficient of lift, the amount of lift created by the wing will increase by four times (2°). Put another way, if the IAS decreases by half, to maintain the same amount of lift from the wing the coefficient of lift must increase by four. This must be done by increasing the angle of attack and/or extending flaps. EOeTEe eT Section 4 Drag 2.4.1 Drag Drag is the component of the total reaction, which is parallel and opposite to the relative airflow. The two main types of drag are: ™ Parasite Drag @ Induced Drag Parasite drag is the drag caused solely by the relative motion of the aircraft to the air. The amount of parasite drag (D,) generated is proportional to the IAS squared. Note: at a given altitude we can also say that parasite drag is proportional to TAS squared. Parasite Drag « IAS’, or YapV" This means that, at low calibrated airspeeds, there is little parasite drag while at high speeds there is a large amount of parasite drag. 64 EOE CHAPTER 2 Speed, Lift, Weight and Drag Parasite Drag Parasite Drag = Y2pV’ S Co» Therefore, the factors affecting parasite drag are: Airspeed (pV) ™S (area) ™C,, (coefficient of parasite drag) The coefficient of parasite drag is a fixed constant for a given aircraft and wing of set thickness, camber, streamlining and smoothness. An aircraft with a large frontal area, highly cambered wings with little streamlining and a frosty or dirty surface would therefore have much greater parasite drag than a sleek, streamlined, polished aircraft. Parasite Drag TAS (knots) Figure 2.6 Parasite drag varies as the square of TAS at a given altitude CHAPTER 2 Speed, Lift, Weight and Drag Induced Drag Induced drag is the drag which results from creating lift. Consequently, it is dominated by the coefficient of lift. The diagram below shows that induced drag is inversely proportional to IAS? in straight and level flight. This means that, for a particular wing the induced drag is greatest at the stalling speed (when the coefficient of lift is greatest) and reduces as TAS increases. Induced drag « —4 IAS Induced Drag TAS (knots) Figure 2.7 induced drag CHAPTER 2 Speed, Lift, Weight and Drag To understand other factors affecting induced drag requires the induced drag equation: Induced Drag = pV’ S C., ™ Airspeed (2 p V’) mS (area) §C,, (coefficient of induced drag) At first glance you would think that induced drag should be proportional to air speed squared, just like parasite drag, and so should increase with speed. This would be true if the coefficient of induced drag remained constant. However, as air speed increases, the angle of attack is reduced to maintain level flight. This causes the coefficient of induced drag to reduce in straight and level flight. It is the coefficient of induced drag which dominates the amount of induced drag created. The equation for the coefficient of induced drag is shown below. C2 Cor = ———+___ pr 7X aspect ratio The coefficient of induced drag increases rapidly with the increase in coefficient of lift (high angles of attack). To maintain straight and level flight (constant lift), when the aircraft's speed halves, the coefficient of lift must be 4 times larger, due to: i Ge However, because the coefficient of induced drag depends on the coefficient of lift squared, the C., will now be 16 times larger (47). nr Su CHAPTER 2 Speed, Lift, Weight and Drag So what happens to induced drag if the speed halves in straight and level flight? Since induced drag = pV" S C,, the C,, will be 16 times larger. However, the airspeed has halved so the dynamic pressure (¥/pV’) will be % of its starting value. Induced drag will therefore be 16 x %, thus 4 times larger. This is why induced drag increases at slower speeds, as shown in the previous ‘Induced drag’ graph and in the initial equation: Induced drag « 1 TA nr Su CHAPTER 2 Speed, Lift, Weight and Drag 2.4.2 Total Drag The total drag acting on an aircraft in flight is the sum of parasite drag and induced drag, shown on the graph below. The graph assumes that lift remains constant as speed changes. Notice that total drag is smallest at medium speed. The speed where total drag is least is known as the minimum drag speed (Vwv). Vols a very important speed for performance. Any time that you fly faster or slower than Vs. drag will increase. Drag 0 TAS (knots) Figure 2.8 Total drag CHAPTER 2 Speed, Lift, Weight and Drag 2.4.3 The Effect of Weight on Drag The graph below shows total drag curves for two identical aircraft types, the only difference between them being their mass. One aircraft is described as being a ‘heavy’ aircraft and the other a ‘light’ aircraft. Note that the heavier aircraft: ™ Generates more drag at all speeds, and @ Has a higher speed corresponding to Vw. » J g| 6 y 0 0 a) TAS (kt) Vino Viewo ight heavy Figure 2.9 The effect of weight on total drag Put simply, the effect of mass is to move the drag curve up and to the right. The cause of this movement is shown on the following graph. CHAPTER 2 Speed, Lift, Weight and Drag Compared with a lighter aircraft, a heavier aircraft needs to generate more lift force at any given IAS. So, at any given IAS, the heavier aircraft must fly at a higher angle of attack. Increased alpha means an increase in C.. As C, increases, induced drag increases, so the induced drag curve moves right. By contrast, an increase in aircraft weight has almost no effect on parasite drag. Summing the two produces the characteristic up and right movement of the total drag curve associated with an increase in mass. IAS (kt) Figure 2.10 The effect of woight on V, In practice (for civilian air operations) aircraft mass never increases during flight; but it does decrease as fuel is burned. So, in flight, the total drag curve moves down and left. CHAPTER 2 Speed, Lift, Weight and Drag As the flight progresses, total drag decreases and V,. (together with the other V speeds) reduce. To put this into approximate numbers: @ If an aircraft becomes 10% lighter, V.. would be 5% slower @ If an aircraft becomes 20% lighter V.. would be 10% slower Thinking about the take-off and climb, an aircraft which is 20% heavier than it’s identical twin would have a V,,. approximately 10% faster than the lighter aircraft. At Vio, induced drag equals parasite drag. WEL & Vo" Taking the square root of both sides means that Vio is proportional to the square root of weight. VW =L & Vv This is useful to us. For small changes the percentage change approximately halves, i.e. a 10% reduction in weight causes an approximate 5% reduction in Vy. CHAPTER 2 Speed, Lift, Weight and Drag The Effect of Speed Brakes, Flaps and Landing Gear on Drag We use common terms to describe how an aircraft is configured: ™ Clean: fully streamlined for the climb and cruise, with gear, flaps and slats retracted Dirty: for slow-speed flight or landing, with gear, flaps and slats extended. The following graph shows how total drag varies with speed for 2 identical aircraft, where one is in a clean configuration and the other is in a dirty configuration. required TAS (knots) dirty clean Figure 2.11 The effect of configuration on drag CHAPTER 2 Speed, Lift, Weight and Drag There are a couple of points to notice. The dirty aircraft has: ™ More drag at all speeds, and ®V,.. is at a slower speed. The following graph shows the original parasite drag as a dashed green line and the original induced drag as a solid orange line. For a constant value of lift, extending flaps and landing gear causes an increase in parasite drag - shown by the continuous green line. Drag Induced drag IAS (kt) Figure 2.12 The etlect of configuration on V, CHAPTER 2 Speed, Lift, Weight and Drag The new total drag, shown by the continuous blue line, has moved up and to the left. This increase in drag is generally detrimental to performance. However, during the approach to land, the increase in drag and reduction in V,. are desirable. The increase in drag is useful because it allows a higher thrust setting in what is otherwise a very low thrust phase of flight. The reduction in Vu: is useful because it increases speed stability - our next topic for discussion! Section 5 Speed Stability 2.5.1 Speed Stability Speed Stable An aircraft is said to be speed stable if it shows a natural tendency to return to its original speed following a brief disturbance without the pilot needing to adjust the thrust setting. Look at the diagram on the next page. The thrust is set to maintain a speed of 270 kt. 76 CHAPTER 2 Speed, Lift, Weight and Drag If a brief disturbance to the flight path causes the speed to drop to 250 kt, the value for total drag will also drop. At 250 kt the thrust force is greater than the drag force so the aircraft will naturally accelerate and will continue to do so until it returns to 270 kt, at which point thrust and drag are, once again, exactly in balance. The opposite is also true. If a brief disturbance to the flight path causes the speed to rise to 290 kt, the total drag force becomes greater than the thrust force and so the aircraft will naturally decelerate back to 270 kt. In this condition the aircraft is said to be speed stable. Speed stability only occurs when the aircraft is flying at speeds greater than Vy... Total Drag o T 250270 2501s (at Vino Figure 2.13 The aircraft is speed stable faster than V, CHAPTER 2 Speed, Lift, Weight and Drag Speed Unstable An aircraft is said to be speed unstable if it shows a natural tendency to diverge from its original speed following a brief disturbance. Look at the diagram below. The aircraft is flying at 140 kt, at which the thrust force exactly counterbalances the total drag force. A small disturbance to the flight path causes the speed to drop to 125 kt; total drag increases. Drag now exceeds thrust so unless thrust is increased the aircraft will continue to decelerate at an increasing rate until it stalls. A Total Drag Drag at 125 kt Thrust 0 oO ore TAS (kt) 125kt 150kt 190 kt Figure 2.14 The aircraft is speed unstable slower than V, Conversely, if a disturbance causes the aircraft's speed to rise to 150 kt the total drag will fall. Thrust exceeds drag and so the aircraft accelerates, continuing to do so until stabilising at 190 kt. An aircraft flying in the speed unstable regime requires constant inputs from the pilot to maintain a constant speed. This is very undesirable, CHAPTER 2 Speed, Lift, Weight and Drag Back of the Drag Curve At speeds slower than V,., increasing amounts of thrust are needed to fly at slower speeds. Acceleration back to the original speed requires even more thrust. In some aircraft it is possible to get yourself into a slow speed situation in which you have only enough thrust to maintain speed but nothing left in reserve to accelerate back to the original speed. The region slower than V,,. is known as the back of the drag curve, an area which requires to you respond quickly to any speed change. The slower your speed the further you will be into the back of the drag curve and the greater the thrust change you will need to recover the situation. This region may be encountered on the final approach, which is why you must monitor your speed closely during this phase. ‘Back of the drag curve’ Total Drag Speed unstable Speed stable 3 3 a neutral Vu Figure 2.15 The speed stability regions CHAPTER 2 Speed, Lift, Weight and Drag Speed Neutral In the region close to Vj. the drag curve is very shallow. So the change in drag following a small disturbance in speed will be insignificant. Consequently, the aircraft will exhibit neither a tendency to return to speed nor to diverge further from the speed. In this part of the drag curve the aircraft is considered to be speed neutral. The faster the air speed is above Vu., the greater the speed stability. Conversely, the slower the air speed is below Vio, the greater the speed instability. CHAPTER 2 Speed, Lift, Weight and Drag 2.5.2 The Relationship Between V,,. and the Stall Having looked at speed stability and its relationship to Vie we can now link in another important relationship: where Vw. lies relative to the stall speed. ™@ For jet aeroplanes Vy. is at approximately 1.6 V;. ™@ For propeller-driven aeroplanes: Vw. is at approximately 1.3 V; For example, if the clean stall speed on a jet aeroplane is 140 kt IAS, Vwwo is approximately 224 kt IAS. Similarly, if the clean stall speed on a propeller-driven aeroplane is 120 kt IAS, Vo is approximately 156 kt. These relationships have fundamental importance in performance. They explain, for example, why jets are much more vulnerable to speed instability. Jet aeroplanes have a wider back-of-the-drag curve which is problematic because they also have a much slower response time to demands for increased thrust. In later chapters you will relate operational speeds such as V,, endurance and range speeds to the stall speed. Section 6 The Drag Polar 2.6.1 The Drag Polar The drag polar is a graph which plots the coefficient of lift (C.) against the coefficient of drag (C.). The coefficient of drag (C,) is the sum of the coefficients of parasite drag (C,.) and induced drag (C.:). Figure 2.16 The drag polar 82 CHAPTER 2 Speed, Lift, Weight and Drag The coefficient of parasite drag (C.,) is constant for a given wing and is shown by the offset in C, at zero C,. The shape of the drag polar is due to the coefficient of induced drag (C,:) which is proportional to C,’(i.e. C.,increases very rapidly as the coefficient of lift increases). Two important points can be identified on the drag polar: ™C/C, maximum (tangent to the curve in the drag polar graph) ™ C, maximum (top edge of the curve) This polar graph is often used as an annex in examinations, with four points, 1, 2, 3 and 4 annotated as shown on the following page. CHAPTER 2 Speed, Lift, Weight and Drag Figure 2.17 The EASA exam polar diagram ™ Point 1: The angle of attack for zero lift. (Only parasite drag.) ™ Point 2: The point where C,//C, is maximum and Vv. @ Point 3: The point equating to minimum power, Vu @ Point 4: The point where C,is at its maximum. Cine OCCurS at the critical angle of attack, so Point 4 occurs just before the stall. We know that C,/C, max. occurs at V.... Point 3 must be a speed between V,,.and the stall, e.g. minimum power speed (Vw). Any point between Points 1 and 2 will be faster than V.. One key speed is 1.32 V.., the optimum maximum range speed for jet aircraft. CHAPTER 2 Speed, Lift, Weight and Drag Drag Polar and Configuration The following diagram shows the effect of extending flaps on the drag polar. When flap setting is increased; Ratio of C, decreases. eC C, MAX increase Cop Cc, Increase Figure 2.18 The effect on the polar diagram of extending flaps ™ C, maximum increases mC, increases ™ C/C, decreases CHAPTER 2 Speed, Lift, Weight and Drag Looking at the movement of the drag polar practically: ™ Extending flaps increases both C,,and C,. Increasing C, allows generation of the same amount of lift at a slower airspeed. This enables us to take-off and land at slower airspeeds and use shorter runways, @ The reduction in C/C, ratio means that the aircraft is now producing more drag per unit of lift, reducing climb performance. Climbing, Descending and Level Flight roe CoM Lael] iS Clatrehaiul ss Section 1 Climbing, Descending, Level Fligh 3.1.1 Climbing, Descending and Level Flight Typical flight conditions include: @ Straight and level @ Climbing ™ Descending § Gliding In all these phases we assume steady flight. This is flight at constant speed, with no change in the flight-path direction. In steady flight, the forces acting on an aircraft are in equilibrium, where opposing forces are in balance. The aircraft is neither turning nor accelerating. This is also known as unaccelerated flight. During unaccelerated flight the speed and direction of the aircraft remain constant. If an exam question doesn’t mention speed, assume that it is constant. 88 Section 2 Straight, Level and Steady Flight 3.2.1 Straight, Level and Steady Flight In straight-and-level flight, the flight path and the relative airflow are horizontal. Thus, lift acts vertically upwards and drag horizontally. Figure 3.1. Forces in balance during straight level and unaccelerated flight In unaccelerated, straight-and-level flight: @ Thrust pushing the aircraft forward = drag opposing forward flight @ Lift (upward) force = weight (downward) force In steady, unaccelerated, straight-and-level flight: L = Wand T =D 89 CHAPTER 3 Climbing, Descending and Level Flight Note: In this book, and in EASA examinations, thrust is assumed to be parallel to the flight path. However, this is not always the case, e.g. during low-speed flight, due to the high angle of attack the thrust vector is angled downwards relative to the direction of flight. Steady Climbing Flight Once an aircraft has stabilised at a constant speed in the climb, each of the opposing forces must once again be equal and opposite. Although lift is still at 90° to drag and thrust, weight is no longer perpendicular to the other two forces. It is therefore easiest to resolve weight (shown as a green vector) into two components: @ Wcosé - perpendicular to the flight path ™@ Wsiné - parallel to the flight path @ is the angle formed between the two forces acting vertically downwards (Weight) and perpendicular to the flight path (Wcos8). Figure 3.2 Weight is resolved into two vector components CHAPTER 3 Climbing, Descending and Level Flight ww an 3 ong ware Figure 3.3 Thrust and drag in the climb In the diagram above the forces parallel to the flight path are shown in red. The only force acting forward along and up the flight path is thrust. If the aircraft is to remain at constant speed, a force (or components of forces) must oppose and balance the thrust. These forces are drag, plus that proportion of weight referred to as Wsin8. Therefore: T=D+ Wsin@ CHAPTER 3 Climbing, Descending and Level Flight wl o NN é\ : aw” 90 Figure 3.4 Lift and weight in the climb Similarly, if the aircraft is not to diverge from a straight climbing flight path, the forces perpendicular to the flight path must also be in balance. This means that lift must be equal and opposite to that component of weight perpendicular to the flight path, Wcos8. Therefore: . Wcos® To summarise, in a steady climb: mL = Wcos8 mT =D + Wsind This means that in a steady climb: Weight is greater than lift (W > L), and @ Thrust is greater than drag (T > D) CHAPTER 3 Climbing, Descending and Level Flight Note: The first of these two statements, that weight exceeds lift, often confuses students. How does the aircraft climb if weight exceeds lift? It may help to think of the forces acting vertically. Now, weight is balanced by lift plus a part of thrust equal to Tsin®. W =L + Tsin@ Hence, weight must be greater than lift. Remember once in a steady climb, the aeroplane must be in equilibrium with the forces in any direction, equal and opposite. Section 3 Steady Descending Flight 3.3.1 Steady Descending Flight Exactly the same principle applies to steady descending flight. Again, if the flight is to be steady, all forces must be in equilibrium. Figure 3.5 Forces in the descent Look at the forces perpendicular to the flight path in a descent at a constant angle. Lift equals the component of weight, Wcos6: L= Wcos8 94 CHAPTER 3 Climbing, Descending and Level Flight From practical flying experience, you may already know that when you lower the nose to descend you must reduce thrust, otherwise the aircraft will accelerate. The diagram below shows why. ‘A component of weight (Wsin8) is now acting in the same direction as thrust. For these two forces to be balanced by drag, which is the only force acting backwards up the flight path, thrust must be somewhat smaller than that required for straight-and-level flight. 929 8 ye o\ 4360 Figure 3.6 Thrust must be reduced in the descent to maintain constant speed To summarise, in a steady descent: D=T + Wsind L = Wcos8 This means that: m Weight is greater than lift (W > L) ™ Drag is greater than thrust (D > T) CHAPTER 3 Climbing, Descending and Level Flight The Glide The terms descent and descending flight normally refer to a powered descent. However, not all descents are powered and an aircraft may need to conduct a glide descent. The term glide is used to describe a descent in which the engines are producing no appreciable thrust, e.g. when the power levers are set to idle or, in extreme circumstances, after all engines have failed. Because the term glide always means a descent with no thrust, only three forces act on the aircraft in a glide: lift, weight, and drag. 0) Figure 3.7 Only three forces are acting in the glide descent Looking at the above diagram, the balancing forces are now simplified with: D = Wsin@ L = Wcos8 CHAPTER 3 Climbing, Descending and Level Flight However, of greatest significance is the fact that the two components of weight are equal to lift and drag. There is a special relationship between the angle of descent, lift, and drag. D p= Tan 0 L This means that the angle of the glide path is steepest when the ratio of drag to lift is greatest. The glide path is shallowest, and range maximum, when the ratio of drag to lift is minimum (D/L min). This is the same as saying the ratio of lift to drag is maximum (L/D max). This means that only the lift to drag ratio determines glide range with respect to the air and not aircraft weight. For a typical training aircraft, the L/D ratio is maximum at about 4 degrees angle of attack. This corresponds to the speed V.., the speed where drag is minimum. CHAPTER 3 Climbing, Descending and Level Flight The Effect of Weight on the Glide The diagram below shows that if the angle of attack remains constant at 4° (Vj), a heavier aircraft will increase its lift and drag in the same proportion, and glide the same distance. Therefore, the heavier aircraft has the same glide path and angle of descent as the lighter aircraft. However, because the heavier aircraft is flying faster to remain at Vso, its rate of descent will be faster and its time in the glide will be reduced. Weary g 35a Figure 3.8 Tho otfect of weight in the glide descent Section 4 Turning 3.4.1 Turning The diagram below shows an aircraft in a level turn to the left. The angle of bank is indicated by the greek symbol ‘0’ (phi). The forces acting on the aircraft in the vertical plane are lift and weight. However, during a turn lift has 2 components: @ The vertical component of lift @ The horizontal component of lift To remain at a constant height during a turn, the vertical component of lift force must be exactly equal and opposite to the weight force. Horizontal Component of Lift (m) J Figure 3.9 Forces in a level tun 99 CHAPTER 3 Climbing, Descending and Level Flight Referring to the diagram: Weight Cosé = i Lift Lift Fe = Because Load Factor Weight Load Factor is inversely proportional to cos ©. Load Factor = — Cos & Apply this to a turn at a bank angle of 30°: Load Factor = ——~— ea "* Cos 30° We find that the load factor = 1.15. The lift required to maintain altitude in a level turn is now 1.15 x the original value. The required increase in lift is obtained by increasing the back pressure on the control column. This increases the angle of attack, which increases the coefficient of lift and causes induced drag to increase. This has implications for both the level and climbing turn. Entering a level turn, the increase in drag now exceeds thrust, which causes the aircraft to slow down. To keep airspeed constant in the turn, thrust must be increased to balance increased drag. Ina climbing turn, induced and total drag will again both increase due to the forces created by the turn. However, the aircraft is already likely to be at full power and the increase in drag will cause the excess thrust, and thus angle of climb, to decrease. 100 Section 5 Height and Altitude 3.5.1 Height and Altitude We have one last topic to address in the chapter, although it's not directly related to what we've been talking about. In performance, two main vertical measurements are used: @ Height: the vertical distance above the ground ™ Pressure altitude: the vertical distance above the standard 1013.25 hPa pressure surface Density altitude is the pressure altitude corrected for non ISA standard temperatures. It determines the performance of the engines. Density altitude is not directly factored into aeroplane performance charts, but its effects can be taken into account by entering pressure altitude and temperature (or ISA deviation). 101 Thrust and mee hiicl a Lig -t a neg -lo(0] Ke] your study time Section 1 Introduction 4.1.1 Thrust and Power Introduction Thrust is the force produced by an engine when it accelerates a mass of air rearwards. Since every action has an equal and opposite reaction, the reaction to the mass of air accelerating backwards is a force which drives the aircraft forwards. This is the force we know as thrust. The same thrust force can be produced by applying a relatively small acceleration to a relatively large mass of air or by giving a large acceleration to a relatively small mass of air. Modern (low bypass) military turbojets are about the only type of propulsion system that still relies on giving a large acceleration to the air. Most power plants including piston, turbo-propeller and high bypass turbofan engines accelerate a relatively large mass relatively slowly. 103 CHAPTER 4 Thrust and Power Figure 4.1 Modern high-bypass engines impart a relatively small acceleration to a relatively large mass of air 104 Section 2 Factors Affecting Thrust 4.2.1 Variation of Jet Thrust with Speed The diagram below shows us that jet thrust does not change much with airspeed. This is because thrust depends on mass (air) flow, multiplied by the velocity change (acceleration) imparted by the engine. As speed increases, mass flow through the engine increases due to ram effect. But the acceleration imparted to the air decreases due to intake momentum drag. The overall outcome is that thrust remains almost constant with speed. Thrust Overall thrust variation with speed Thrust lost due Thrust gained to momentum drag due to ram effect TAS Figure 4.2 Jet engine thrust remains relatively constant over the operating speed range 105 CHAPTER 4 Thrust and Power Intake momentum drag describes the loss of thrust resulting from an increase in forward speed. When a jet aircraft's speed increases, so too does the velocity of the air entering approaching the intake. But the exit speed remains constant. Consequently, the acceleration given by the engine to the air reduces, resulting in reduced thrust. Intake momentum drag at different speeds is shown by the red area in the diagram. However, as Mach number increases there is an increase in the mass of air that is accelerated. This is because compressibility effects cause the pressure in the intake to rise. This increases the air density and thus the mass flow. Ram effect is the increased thrust due to the increase in mass flow. The contribution to thrust caused by ram effect is shown in green in the diagram. At any given airspeed, the loss of thrust from intake momentum drag is approximately equal to the increase in thrust from ram effect. CHAPTER 4 Thrust and Power High and Low Bypass Ratio Engines We have been economical with the truth. Our claim that thrust remains constant with speed is only true for low bypass engines. The diagram above shows that for a modern, high bypass, engine thrust reduces significantly with forward speed. This is because the ram effect isn’t so apparent in the bypass duct and is therefore insufficient to compensate for all of the intake momentum drag. ry Low-bypass engine Thrust ~—___ High-bypass engine ee v Figure 4.3 Comparison of low and high bypass engine However, for the purposes of this course, we assume we are dealing with a low bypass ratio engine because this permits a simpler introduction to the concepts of jet performance. So, from now on we shall assume that thrust is constant with speed, except during the take-off run when it is assumed to reduce slightly as speed increases. CHAPTER 4 Thrust and Power 4.2.2 Variation of Jet Thrust with Altitude As altitude increases, thrust reduces, because the air density and therefore the mass flow reduces. The diagram below shows the reduction in thrust available as altitude increases. Low altitude High altitude Thrust Available Figure 4.4 Reduction in thrust available with altitude CHAPTER 4 Thrust and Power 4.2.3 Variation of Thrust with Temperature An increase in ambient air temperature slightly reduces air density and therefore mass flow. In practice, it is usually either the turbine gas temperature (TGT) limit or RPM limit that restricts the available thrust. The RPM limit prevents the maximum pressure ratio of the compressor from being exceeded and it also keeps the compressor within acceptable limits of centrifugal forces. On hot days (engine specific but normally above ISA+15°C), the TGT limit is reached first. If the outside air temperature gets hotter, the TGT is reached at a slightly slower RPM, resulting in the mass flow and thrust available reducing. Conversely, the thrust available increases as the outside air temperature gets colder. This is because the RPM at which the TGT becomes limiting is higher, allowing a greater maximum mass flow. In short, as temperature increases above ISA+15°C, thrust reduces. The orange line in the following diagram shows that the effect of temperature on thrust become appreciable when operating in ambient temperatures greater than 30°C at mean sea level ISA. 109 CHAPTER 4 Thrust and Power = & ~S 2 id Limiting RPM 5 2 = = Thrust limited by RPM in this area Flat rating cut-off ISA deviation ISA-15°C ISA ISA+15°C. ISA +30°C. ISA +45°C ISA Deviation Figure 4.5 Thrust limited either by RPM or by temperature At outside air temperatures below ISA+15°C, the RPM limit is reached first. Within the temperature range where thrust is regulated by RPM, thrust does not change with temperature. This is shown by the grey horizontal line in the above graph. On older 3-crew aircraft the RPM used to be limited manually by a flight engineer. Now this is normally achieved electronically by flat rating the engine below the cut-off ISA deviation temperature (normally ISA+15°C). Section 3 Flat Rating and Thrust Settings 4.3.1 Flat-rated Engines At all temperatures, an increase in pressure altitude reduces thrust available. This means that an aircraft has more thrust available at an airfield at mean sea level than an airfield on a high plateau. However, at a given pressure altitude, the thrust available may or may not vary with temperature change. Thrust available does not change if the temperature changes between temperatures below ISA+15°C, because the engines are flat-rated at this ISA deviation. Conversely, thrust does reduce at a given pressure altitude if the temperature increases above ISA+15°C. The following diagram is a reproduction of Figure 4.5 from CAP 698. This graph establishes the climb-limited take-off mass for a given airfield pressure altitude and temperature. The climb limited take-off mass is the heaviest mass at which the aircraft can just achieve the minimum climb gradient required after take-off. The greater the thrust available, the greater the climb mass for the same gradient. At hot temperatures (above ISA+15°C), the climb mass reduces as temperature increases. However, at temperatures below ISA+15°C (shaded yellow), the pressure altitude lines are now almost horizontal, showing that air temperature does not affect the climb mass. This is because the engine is flat-rated below ISA+15°C. The kink shown in red in the previous diagram is at ISA+15°C. This represents the different temperatures at different altitudes, below which the engines are flat-rated. 1” CHAPTER 4 Thrust and Power CLIMB LIMIT BRAKE RELEASE MASS 1000 kg 65 60 65 50 45 40 35 30 Temperature area in which engines are flatrated t | 1 | 1 | 1 | S 1 ! t lw | 2 2 h 18 5 10 20 30 40 50 60 70 FLAP & AIRPORT OAT °C POSITION BELOW Figure 4.6 Climb limited take-off mass for a given airfield pressure altitude and temperature CAP 698 CHAPTER 4 Thrust and Power 4.3.2 Engine Thrust Settings The two main engine thrust settings are TOGA - Take-Off Go Around and MCT or CON - Maximum Continuous Thrust. TOGA is used for take off or a go-around and for the engines in the reference MRJT aircraft, TOGA is normally limited to a maximum of 5 minutes for all engine operation and a maximum of 10 minutes with one engine out. The maximum thrust setting that can be used without a time limit in flight is Maximum Continuous Thrust. However depending on aircraft type there can be other engine thrust settings such as CLB (climb) which is a higher thrust setting than maximum continuous. Section 4 Propeller Thr 4.4.1 Propeller Thrust A propeller produces thrust in exactly the same way as a jet engine; by accelerating a mass of air rearwards. However, the method it uses to generate the acceleration is somewhat different. A rotating propeller blade produces an aerodynamic total reaction. But because the propeller is mounted at approximately 90° to the relative airflow, the total reaction faces approximately forwards rather than approximately upwards. This force is resolved into thrust and torque (the propeller equivalent of induced drag). Propeller thrust depends on both the blades’ angle of attack and the amount of thrust generated in relation to torque (i.e. the direction of the total reaction force in relation to the propeller axis of rotation). For a fixed pitch propeller the angle of attack reduces as the TAS of the aircraft increases. This means that thrust will be maximum when the aircraft is stationary and reduce with forward air speed. The thrust of a variable pitch propeller also reduces as forward speed increases, but at a slower rate. This is achieved by increasing the propeller blade angle, to maintain the same most efficient angle of attack over a wide range of normal flying speeds. However, compared to a turbo-jet or turbo-fan engine neither type of propeller is nearly as effective at imparting an acceleration to the air. Thus the relative acceleration given to the air reduces rapidly with an increase in TAS. And with no ram effect to compensate, propeller thrust reduces much more quickly with TAS than jet engine thrust. 14 CHAPTER 4 Thrust and Power Thrust CAS Figure 4.7 Variation of propeller thrust with air speed Variation of Propeller Thrust with Altitude Propeller thrust reduces as altitude increases because of the reduction in air density. Similarly thrust will decrease as temperature increases. A high density altitude, high pressure altitude and/or high ambient temperature all reduce engine thrust. Section 5 Power 4.5.1 Power Power is often confused with thrust. Although related, they are not the same. Thrust is a force. Power is force multiplied by speed. Power is defined as the rate of doing work: _ Work Power = Time Work is equal to force multiplied by distance: Work = Force x Distance Therefore, substituting this into the original equation for power: Force x Distance Pe = ower The Because: Distance Velocity = == locity Time It can be said that: Power = Force x Velocity In aviation terminology there are two power terms of interest to us: ™ Power available @ Power required 116 CHAPTER 4 Thrust and Power Power Available - Jet Power available is thrust multiplied by TAS: Power Available = Thrust x Velocity The diagram below compares power available with thrust. The orange line plots thrust for a jet against TAS. Thrust is approximately constant with forward speed. The purple line plots power available for a jet against TAS. Power available (PA) increases as TAS increases. This is because thrust, which is constant, is multiplied by the increasing TAS. Thrust Jet Tiet PA jet TAS Figure 4.8 Power available - jet Before brake release there is no power available, because TAS is zero, However, as anyone who has stood behind a commercial aircraft at take-off thrust can testify, there most certainly is considerable thrust!

You might also like