Category AVIATORS

CONFIGURATION EFFECTS

TRANSONIC AND SUPERSONIC FLIGHT

Any object in subsonic flight which has some finite thickness or is producing lift will have local velocities on the surface which are greater than the free stream velocity. Hence, compressibility effects can be expected to occur at flight speeds less than the speed of sound. The transonic regime of flight pro­vides the opportunity for mixed subsonic and supersonic flow and accounts for the first | significant effects of compressibility.

Consider a conventional airfoil shape as shown in figure 3-9. If this airfoil is at a flight Mach number of 0.50 and a slight posi­tive angle of attack, the maximum local velocity on the surface will be greater than the flight speed but most likely less than sonic speed. Assume that an increase in flight Mach number to 0.72 would produce first evidence of local sonic flow. This condition of flight would be the highest flight speed possible without supersonic flow and would be termed the ‘ ‘critical Mach number. ’ ’ Thus, critical Mach number is the boundary between subsonic and transonic flight and is an im­portant point of reference for all compressi – | bility effects encountered in transonic flight. By definition, critical Mach number is the “free stream Mach number which produces first evidence of local sonic flow.” Therefore, shock waves, buffet, airflow separation, etc., take place above critical Mach number.

As critical Mach number is exceeded an area of supersonic airflow is created and a normal

Revised January 1965

M = .72

(CRITICAL MACH NUMBER)

M = .800 a = +2° CL= .442 SHOCK FORMATION IS APPARENT AT 25 TO 30 % CHORD POSITION

M = .875 a =4-2° CL= .450 SHOCK INDUCED SEPARATION ALONG AFT PORTION OF WING PLANFORM

shock wave forms as the boundary between the supersonic flow and the subsonic flow on the aft portion of the airfoil surface. The acceleration of the airflow from subsonic to supersonic is smooth and unaccompanied by shock waves if the surface is smooth and the transition gradual. However, the transition of airflow from supersonic to subsonic is always accompanied by a shock wave and, when there is no change in direction of the airflow, the wave form is a normal shock wave.

Recall that one of the principal effects of the normal shock wave is to produce a large increase in the static pressure of the airstream behind the wave. If the shock wave is strong, the boundary layer may not have sufficient kinetic energy to withstand the large, adverse pressure gradient and separation will occur. At speeds only slightly beyond critical Mach number the shock wave formed is not strong enough to cause spearation or any noticeable change in the aerodynamic force coefficients. However, an increase in speed above critical Mach number sufficient to form a strong shock wave can cause sepa­ration of the boundary layer and produce sudden changes in the aerodynamic force coefficients. Such a flow condition is shown in figure 3.9 by the flow pattern for M=0.77. Notice that a further increase in Mach number to 0.82 can enlarge the supersonic area on the upper surface and form an additional area of supersonic flow and normal shock wave on the lower surface.

As the flight speed approaches the speed of sound the areas of supersonic flow enlarge and the shock waves move nearer the trailing edge. The boundary layer may remain sepa­rated or may reattach depending much upon the airfoil shape and angle of attack. When the flight speed exceeds the speed of sound the ‘ ‘ bow’ ’ wave forms at the leading edge and this typical flow pattern is illustrated in figure 3 9 by the drawing for M=1.05. If the speed is increased to some higher supersonic value all oblique portions of the waves incline more greatly and the detached normal shock portion of the bow wave moves closer to the leading edge.

Of course, all components of the aircraft are affected by compressibility in a manner somewhat similar to that of basic airfoil. The tail, fuselage, nacelles, canopy, etc. and the effect of the interference between the various surfaces of the aircraft must be considered.

FORCE DIVERGENCE. The airflow sepa­ration induced by shock wave formation can create significant variations in the aerody­namic force coefficients. When the free stream speed is greater than critical Mach number some typical effects on an airfoil section are as follows :

(1) An increase in the section drag coeffi­cient for a given section lift coefficient.

(2) A decrease in section lift coefficient for a given section angle of attack.

0) A change in section pitching moment coefficient.

A reference point is usually taken by a plot of drag coefficient versus Mach number for a constant lift coefficient. Such a graph is shown in figure З. Ю. The Mach number which produces a sharp change in the drag coefficient is termed the “force divergence” Mach number and, for most airfoils, usually exceeds the critical Mach number at least 5 to 10 percent. This condition is also referred to as the “drag divergence” or “drag rise.” PHENOMENA OF TRANSONIC FLIGHT. Associated with the "drag rise” are buffet, trim and stability changes, and a decrease in control surface effectiveness. Conventional aileron, rudder, and elevator surfaces sub­jected to this high frequency buffet may “buzz,” and changes in hinge moments may produce undesirable control forces. Of course, if the buffet is quite severe and prolonged, structural damage may occur if this operation is in violation of operating limitations. When airflow separation occurs on the wing due to

shock wave formation, there will be a loss of lift and subsequent loss of downwash aft of the affected area. If the wings shock unevenly due to physical shape differences or sideslip, a rolling moment will be created in the direction of the initial loss of lift and con­tribute to control difficulty (‘‘wing drop”). If the shock induced separation occurs sym­metrically near the wing root, a decrease in downwash behind this area is a corollary of the loss of lift. A decrease in downwash on the horizontal tail will create a diving moment and the aircraft will ‘‘tuck under.” If these conditions occur on a swept wing planform, the wing center of pressure shift contributes to the trim change—root shock first moves the wing center of pressure aft and adds to the diving moment; shock formation at the wing tips first moves the center of pressure forward and the resulting climbing moment and tail

downwash change can contribute to ‘‘pitch up.”

Since most of the difficulties of transonic flight are associated with shock wave induced flow separation, any means of delaying or alleviating the shock induced separation will improve the aerodynamic characteristics. An aircraft configuration may utilize thin surfaces of low aspect ratio with sweepback to delay and reduce the magnitude of transonic force divergence. In addition, various methods of boundary layer control, high lift devices, vortex generators, etc., may be applied to improve transonic characteristics. For exam­ple, the application of vortex generators to a surface can produce higher local surface veloci­ties and increase the kinetic energy of the boundary layer. Thus, a more severe pressure gradient (stronger shock wave) will be neces­sary to produce airflow separation.

Once the configuration of a transonic air­craft is fixed, the pilot must respect the effect of angle of attack and altitude. The local flow I velocities on any upper surface increase with an increase in angle of attack. Hence, local sonic flow and subsequent shock wave formation can occur at lower free stream Mach numbers. A pilot must appreciate this reduction of force divergence Mach number with lift coefficient since maneuvers at high speed may produce compressibility effects which may not be en­countered in unaccelerated flight. The effect of altitude is important since the magnitude of any force or moment change due to com­pressibility will depend upon the dynamic pressure of the airstream. Compressibility effects encountered at high altitude and low dynamic pressure may be of little consequence in the operation of a transonic aircraft. How­ever, the same compressibility effects en­countered at low altitudes and high dynamic pressures will create greater trim changes, heavier buffet, etc., and perhaps transonic flight restrictions which are of principal inter­est only to low altitude.

PHENOMENA OF SUPERSONIC FLIGHT. While many of the particular effects of super­sonic flight will be presented in the detail of later discussion, many general effects may be anticipated. The airplane configuration must have aerodynamic shapes which will have low drag in compressible flow. Generally, this will require airfoil sections of low thickness ratio and sharp leading edges and body shapes of high fineness ratio to minimize the supersonic wave drag. Because of the aft movement of the aerodynamic center with supersonic flow, the increase in static longitudinal stability will demand effective, powerful control surfaces to achieve adequate controllability for super­sonic maneuvering.

As a corollary of supersonic flight the shock wave formation on the airplane may create special problems outside the immediate vicinity of the airplane surfaces. While the shock waves a great distance away from the airplane

can be quite weak, the pressure waves can be of sufficient magnitude to create an audible disturbance. Thus, “sonic booms" will be a simple consequence of supersonic flight.

The aircraft power pi ante for supersonic flight must be of relatively high thrust output. Also, in many cases it may be necessary to provide the air breathing powerplant with special inlet configurations which will slow the airflow to subsonic prior to reaching the compressor face or combustion chamber. Aero­dynamic heating of supersonic flight can pro­vide critical inlet temperatures for the gas turbine engine as well as critical structural temperatures.

The density variations in airflow may be shown by certain optical techniques. Schlieren photographs and shadowgraphs can define the various wave patterns and their effect on the airflow. The Schlieren photographs presented in figure 3-11 define the flow conditions on an aircraft in supersonic flight.

SECTIONS IN SUPERSONIC FLOW

In order to appreciate the effect of these various wave forms on the aerodynamic char­acteristics in supersonic flow, inspect figure 3.8. Parts (a) and (b) show the wave pattern and resulting pressure distribution for a thin flat plate at a positive angle of attack. The air – stream moving over the upper surface passes through an expansion wave at the leading edge and then an oblique shock wave at the trailing edge. Thus, a uniform suction pressure exists over the upper surface. The airstream moving underneath the flat plate passes through an oblique shock wave at the leading edge then an expansion wave at the trailing edge. This pro­duces a uniform positive pressure on the under­side of the section. This distribution of pres­sure on the surface will produce a net lift and incur a subsequent drag due to lift from the in­clination of the resultant lift from a perpen­dicular to the free stream.

Parts (c) and (d) of figure 3.8 show the wave pattern and resulting pressure distribu­tion for a double wedge airfoil at zero lift. The airstream moving over the surface passes through an oblique shock, an expansion wave, and another oblique shock. The resulting pressure distribution on the surfaces produces no net lift, but the increased pressure on the forward half of the chord along with the de­creased pressure on the aft half of the chord produces a “wave” drag. This wave drag is caused by the components of pressure forces which are parallel to the free stream direction. The wave drag is in addition to the drag due to friction, separation, lift, etc., and can be a very considerable part of the total drag at high supersonic speeds.

Parts (e) and (f) of figure 3.8 illustrate the wave pattern and resulting pressure distribu­tion for the double wedge airfoil at a small positive angle of attack. The net pressure

distribution produces an inclined lift with drag due to lift which is in addition to the wave drag at zero lift. Part (g) of figure 3.8 shows the wave pattern for a circular arc air­foil. After the airflow traverses the oblique shock wave at the leading edge, the airflow undergoes a gradual but continual expansion until the trailing edge shock wave is en­countered. Part (h) of figure 3-8 illustrates the wave pattern on a conventional blunt nose airfoil in supersonic flow. When the nose is blunt the wave must detach and become a normal shock wave immediately ahead of the leading edge. Of course, this wave form produces an area of subsonic airflow at the leading edge with very high pressure and density behind the detached wave.

The drawings of figure 3-8 illustrate the typical patterns of supersonic flow and point out these facts concerning aerodynamic surfaces in two dimensional supersonic flow:

(1) All changes in velocity, pressure, density and flow direction will take place quite suddenly through the various. wave forms. The shape of the object and the required flow direction change dictate the type and strength of the wave formed.

(2j) As always, lift results from the distri­bution of pressure on a surface and is the net force perpendicular to the free stream direc­tion. Any component of the lift in a direc­tion parallel to the wind stream will be drag due to lift.

(3) In supersonic flight, the zero lift drag of an airfoil of some finite thickness will include a "wave drag.” The thickness of the airfoil will have an extremely powerful effect on this wave drag since the wave drag varies as the square of the thickness ratio— if the thickness is reduced 50 percent, the wave drag is reduced 75 percent. The lead­ing edges of supersonic shapes must be sharp or the wave formed at the leading edge will be a strong detached shock wave.

(4) Once the flow on the airfoil is super­sonic, the aerodynamic center of the surface

will be located approximately at the 50 per­cent chord position. As this contrasts with the subsonic location for the aerodynamic center of the 25 percent chord position, sig­nificant changes in aerodynamic trim and stability may be encountered in transonic flight.

HIGH SPEED AERODYNAMICS

Developments in aircraft and powerplants have produced high performance airplanes with capabilities for very high speed flight. The study of aerodynamics at these very high flight speeds has many significant differences from the study of classical low speed aero­dynamics. Therefore, it is quite necessary that the Naval Aviator be familiar with the nature of high speed airflow and the charac­teristics of high performance airplane configurations.

GENERAL CONCEPTS AND SUPERSONIC
FLOW PATTERNS

NATURE OF COMPRESSIBILITY

At low flight speeds the study of aero­dynamics is greatly simplified by the fact that air may experience relatively small changes in pressure with only negligible changes in density. This airflow is termed incompressible since the air may undergo changes

in pressure without apparent changes in den­sity. Such a condition of airflow is analogous to the flow of water, hydraulic fluid, or any other incompressible fluid. However, at high flight speeds the pressure changes that take place are quite large and significant changes in air density occur. The study of airflow at high speeds must account for these changes in air density and must consider that the air is compressible and that there will be * ‘compressibility effects.’ ’

A factor of great importance in the study of high speed airflow is the speed of sound. The speed of sound is the rate at which small pressure disturbances will be propagated through the air and this propagation speed is solely a function of air temperature. The accompanying table illustrates the variation of the speed of sound in the standard atmosphere.

TABLE 3-1. Variation of Temperature and Speed of Sound with Altitude in the Standard Atmosphere

Altitude

Temperature

Speed of sound

Ft,

° F.

"C.

Knots

Sea level………………………..

59.0

15.0

661.7

5,000…………………………….

41.2

5.1

650.3

10,000…………………………..

23.3

-4.8

638.6

15,000…………………………..

5.5

-14.7

616.7

20,000…………………………..

-12.3

-24.6

614.6

25,000…………………………..

-30.2

-34.5

602.2

30,000…………………………..

-48.0

-44.4

589.6

35,000…………………………..

-65.8

-54.3

576.6

40,000…………………………..

-69.7

-56.5

573-8

50,000…………………………..

-69.7

-56.5

573.8

60,000…………………………..

-69.7

-56.5

573.8

As an object moves through the air mass, velocity and pressure changes occur which create pressure disturbances in the airflow sur­rounding the object. Of course, these pressure disturbances are propagated through the air at the speed of sound. If the object is travel­ling at low speed the pressure disturbances are propagated ahead of the object and the airflow immediately ahead of the object is influenced by the pressure field on the object. Actually, these pressure disturbances are transmitted in all directions and extend indefinitely in all
directions. Evidence of this “pressure warn­ing’’ is seen in the typical subsonic flow pattern of figure 3.1 where there is upwash and flow direction change well ahead of the leading edge. If the object is travelling at some speed above the speed of sound the air­flow ahead of the object will not be influenced by the pressure field on the object since pres­sure disturbances cannot be propagated ahead of the object. Thus, as the flight speed nears the speed of sound a compression wave will form at the leading edge and all changes in velocity and pressure will take place quite sharply and suddenly. The airflow ahead of the object is not influenced until the air par­ticles are suddenly forced out of the way by the concentrated pressure wave set up by the object. Evidence of this phenomenon is seen in the typical supersonic flow pattern of figure 3.1.

The analogy of surface waves on the water may help clarify these phenomena. Since a surface wave is simply the propagation of a pressure disturbance, a ship moving at a speed much less than the wave speed will not form a “bow wave.” As the. ship’s speed nears the wave propagation speed the bow wave will form and become stronger as speed is increased beyond the wave speed.

At this point it should become apparent that all compressibility effects depend upon the relationship of airspeed to the speed of sound. The term used to describe this rela­tionship is the Mach number, M, and this term is the ratio of the true airspeed to the speed of sound.

M=Mach number F=true airspeed, knots a= speed of sound, knots

= Яо^в

Oo=speed of sound at standard sea level conditions, 661 knots 0 = temperature tatio

= r/re

Revised January 1965

It is important to note that compressibility effects are not limited to flight speeds at and above the speed of sound. Since any aircraft will have some aerodynamic shape and will be developing lift there will be local flow velocities on the surfaces which are greater than the flight speed. Thus, an aircraft can experience compressibility effects at flight speeds well below the speed of sound. Since there is the possibility of having both subsonic and supersonic flows existing on the aircraft it is convenient to define certain regimes of flight. These regimes are defined approxi­mately as follows:

Subsonic—Mach numbers below 0.75

Transonic—Mach numbers from 0.75 to

1.20

Supersonic—Mach numbers from 1.20 to

5.00

Hypersonic—Mach numbers above 5.00 While the flight Mach numbers used to define these regimes of flight are quite approximate, it is important to appreciate the types of flow existing in each area. In the subsonic regime it is most likely that pure subsonic airflow exists on all parts of the aircraft. In the transonic regime it is very probable that flow on the aircraft components may be partly sub­sonic and partly supersonic. The supersonic and hypersonic flight regimes will provide definite supersonic flow velocities on all parts of the aircraft. Of course, in supersonic flight there will be some portions of the boundary layer which are subsonic but the predominating flow is still supersonic.

The principal differences between subsonic and supersonic flow are due to the compres­sibility of the supersonic flow. Thus, any change of velocity or pressure of a supersonic flow will produce a related change of density which must be considered and accounted for. Figure 3-2 provides a comparison of incom­pressible and compressible flow through a closed tube. Of course, the condition of con­tinuity must exist in the flow through the closed tube; the mass flow at any station along the tube is constant. This qualification must exist in both compressible and incompressible cases.

The example of subsonic incompressible flow is simplified by the fact that the density of flow is constant throughout the tube. Thus, as the flow approaches a constriction and the streamlines converge, velocity increases and static pressure decreases. In other words, a convergence of the tube requires an increasing velocity to accommodate the continuity of flow. Also, as the subsonic incompressible flow enters a diverging section of the tube, velocity decreases and static pressure increases but density remains unchanged. The behavior of subsonic incompressible flow is that a con­vergence causes expansion (decreasing pressure) while a divergence causes compression (in­creasing pressure).

The example of supersonic compressible flow is complicated by the fact that the variations of flow density are related to the changes in velocity and static pressure. The behavior of supersonic compressible flow is that a con­vergence causes compression while a divergence causes expansion. Thus, as the supersonic compressible flow approaches a constriction and the streamlines converge, velocity de­creases and static pressure increases. Con­tinuity of mass flow is maintained by the increase in flow density which accompanies the decrease in velocity. As the supersonic com­pressible flow enters a diverging section of the tube, velocity increases, static pressure de­creases, and density decreases to accommodate the condition of continuity.

The previous comparison points out three significant differences between supersonic com­pressible and subsonic incompressible flow.

(a) Compressible flow includes the addi­tional variable of flow density.

(b) Convergence of flow causes expansion of incompressible flow but compression of compressible flow.

(r) Divergence of flow causes compression of incompressible flow but expansion of compressible flow.

Revised January 1965

DECREASING VELOCITY
INCREASING PRESSURE
INCREASING DENSITY

INCREASING VELOCITY
DECREASING PRESSURE
DECREASING DENSITY

Figure 3.2. Comparison of Compressible and Incompressible Flow Through a Closed Tube


When supersonic flow is clearly established, all changes in velocity, pressure, density, flow direction, etc., take place quite suddenly and in relatively confined areas. The areas of flow change are generally distinct and the phenom­ena are referred to as “wave” formations. All compression waves occur suddenly and are wasteful of energy. Hence, the compression waves are distinguished by the sudden “shock” type of behavior. All expansion waves are not so sudden in their occurrence and are not waste­ful of energy like the compression shock waves. Various types of waves can occur in supersonic flow and the nature of the wave formed depends upon the airstream and the shape of the object causing the flow change. Essentially, there are three fundamental types of waves formed in supersonic flow: (1) the oblique shock wave (compression), (2) the normal shock wave (compression), (3) the expansion wave (no shock).

OBLIQUE SHOCK WAVE. Consider the case where a supersonic airstream is turned into the preceding airflow. Such would be the case of a supersonic flow “into a comer” as shown in figure 3.3- A supersonic airstream passing through the oblique shock wave will experience these changes:

(1) The airstream is slowed down; the velocity and Mach number behind the wave are reduced but the flow is still supersonic

(2) The flow direction is changed to flow along the surface

(3) The static pressure of the airstream behind the wave is increased

(4) The density of the airstream behind the wave is increased

(5) Some of the available energy of the airstream (indicated by the sum of dynamic and static pressure) is dissipated and turned into unavailable heat energy. Hence, the shock wave is wasteful of energy.

A typical case of oblique shock ■’wave forma­tion is that of a wedge pointed into a super­sonic airstream. The oblique shock wave will form on each surface of the wedge and the inclination of the shock wave will be a func­tion of the free stream Mach number and the wedge angle. As the free stream Mach number increases, the shock wave angle decreases; as the wedge angle increases the shock wave angle increases, and, if the wedge angle is in­creased to some critical amount, the shock wave will detach from the leading edge of the wedge. It is important to note that detach­ment of the shock wave will produce subsonic flow immediately after the central portion of the shock wave. Figure 3-4 illustrates these typical flow patterns and the effect of Mach number and wedge angle.

The previous flow across a wedge in a supersonic airstream would allow flow in two dimensions. If a cone were placed in a super­sonic airstream the airflow would occur in three dimensions and there would be some noticeable differences in flow characteristics. Three-dimensional flow for the same Mach number and flow direction change would pro­duce a weaker shock wave with less change in pressure and density. Also, this conical wave formation allows changes in airflow that con­tinue to occur past the wave front and the wave strength varies with distance away from the surface. Figure 3.5 depicts the typical three-dimensional flow past a cone.

Oblique shock waves can be reflected like any pressure wave and this effect is shown in figure 3-5- This reflection appears logical and necessary since the original wave changes the flow direction toward the wall and the reflected wave creates the subsequent flow change to cause the flow to remain parallel to the wall surface. This reflection phenomenon places definite restrictions on the size of a model in a wind tunnel since a wave reflected back to the model would cause a pressure distribution not typical of free flight.

NORMAL SHOCK WAVE. If a blunt­nosed object is placed in a supersonic airstream the shock wave which is formed will be de­tached from the leading edge. This detached

Figure 3.4. ShockWaves Formed by VariousWedge Shapes

REFLECTED OBLIQUE WAVES

MOOEL IN WIND
TUNNEL WITH WAVES
REFLECTED FROM
WALLS

Figure 3.5. Three Dimensional and Reflected Shock Waves

Revised January 1965

wave also occurs when a wedge or cone angle exceeds some critical value. Whenever the shock wave forms perpendicular to the up­stream flow, the shock wave is termed a ‘ ‘normal” shock wave and the flow immediately behind the wave is subsonic. Any relatively blunt object in a supersonic airstream will form a normal shock wave immediately ahead of the leading edge slowing the airstream to subsonic so the airstream may feel the presence of the blunt nose and flow around it. Once past the blunt nose the airstream may remain subsonic or accelerate back to supersonic depending on the shape of the nose and the Mach number of the free stream.

In addition to the formation of normal shock waves described above, this same type of wave may be formed in an entirely different manner when there is no object in the super­sonic airstream. It is particular that whenever a supersonic airstream is slowed to subsonic without a change in direction a normal shock wave will form as a boundary between the supersonic and subsonic regions. This is an important fact since aircraft usually encounter some ‘‘compressibility effects” before the flight speed is sonic. Figure 3.6 illustrates the man­ner in which an airfoil at high subsonic speeds has local flow velocities which are supersonic. As the local supersonic flow moves aft, a normal shock wave forms slowing the flow to subsonic. The transition of flow from subsonic to supersonic is smooth and is not accompanied by shock waves if the transition is made gradually with a smooth surface. The transition of flow from supersonic to subsonic without direction change always forms a normal shock wave.

A supersonic airstream passing through a normal shock wave will experience these changes:

(1) The airstream is slowed to subsonic;

the local Mach number behind the wave is

approximately equal to the reciprocal of the

Mach number ahead of the wave—e. g., if

Mach number ahead of the wave is 1.25, the Mach number of the flow behind the wave is approximately 0.80.

(2) The airflow direction immediately behind the wave is unchanged.

(3) The static pressure of the airstream behind the wave is increased greatly.

(4) The density of the airstream behind the wave is increased greatly.

(5) The energy of the airstream (indi­cated by total pressure—dynamic plus static) is greatly reduced. The normal shock wave is very wasteful of energy.

EXPANSION WAVE. If a supersonic air­stream were turned away from the preceding flow an expansion wave would form. The flow “around a corner” shown in figure 3 7 will not cause sharp, sudden changes in the airflow except at the corner itself and thus is not actually a “shock” wave. A supersonic airstream passing through an expansion wave will experience these changes:

(1) The airstream is accelerated; the ve­locity and Mach number behind the wave are greater.

(2) The flow direction is changed to flow along the surface—provided separa­tion does not occur.

(3) The static pressure of the airstream behind the wave is decreased.

(4) The density of "the airstream behind the wave is decreased.

(5) Since the flow changes in a rather gradual manner there is no “shock” and no loss of energy in the airstream. The expansion wave does not dissipate air­stream energy.

The expansion wave in three dimensions is a slightly different case and the principal difference is the tendency for the static pres­sure to continue to increase past the wave.

The following table is provided to summa­rize the characteristics of the three principal wave forms encountered with supersonic flow.

TABLE 3-2. Supeitonic Wave Choracttrittict

Type of wave formation………..

Oblique shock wave………….

Normal shoe

:k wave………….

Expansion wave. –

Flow direction change……………

‘“Flow into a corner," turned into preceding flow.

No change…………………………

"Flow around a corner," turned away from pre­ceding flow.

Effect on velocity and Mach number.

Decreased but still super­sonic.

Decreased to subsonic………

Increased to higher super­sonic.

Effect on static pressure and density.

Increase……………………………..

Great increase……………………

Decrease.

Effect on energy or total pres­sure.

Decrease…………………………….

Great decrease…………………..

No change (no shock).

TAKEOFF AND LANDING PERFORMANCE

The majority of pilot caused airplane acci­dents occur during the takeoff and landing phase of flight. Because of this fact, the Naval Aviator must be familiar with all the many variables which influence the takeoff and landing performance of an airplane and must strive for exacting, professional techniques of operation during these phases of flight.

Takeoff and landing performance is a con­dition of accelerated motion. For instance, during takeoff the airplane starts at zero veloc­ity and accelerates to the takeoff velocity to become airborne. During landing, the air­plane touches down at the landing speed and decelerates (or accelerates negatively) to the zero velocity of the stop. In fact, the landing performance could be considered as a takeoff in reverse for purposes of study. In either case, takeoff or landing, the airplane is ac­celerated between zero velocity and the takeoff or landing velocity. The important factors of takeoff or landing performance are:

(1) The takeoff" or landing velocity which

will generally be a function of the stall

speed or minimum flying speed, e. g., 15 per­cent above the stall speed.

(2) The acceleration during the takeoff or landing roll. The acceleration experienced by any object varies directly with the un­balance of force and inversely as the mass of the object.

(3) The takeoff or landing roll distance is a function of both the acceleration and velocity.

In the actual case, the takeoff and landing dis­tance is related to velocity and acceleration in a very complex fashion. The main source of the complexity is that the forces acting on the airplane during the takeoff or landing roll are ’difficult to define with simple relationships. Since the acceleration is a function of these forces, the acceleration is difficult to define in a simple fashion and it is a principal variable affecting distance. However, some simplifica­tion can be made to study the basic relationship of acceleration, velocity, and distance While the acceleration is not necessarily constant or uniform throughout the takeoff or landing roll, the assumption of uniformly acceler­ated motion will facilitate study of the princi­pal variables. affecting takeoff and landing distance.

From basic physics, the relationship of velocity, acceleration, and distance for uni­formly accelerated motion is defined by the following equation: where

S= acceleration distance, ft.

V= final velocity, ft. per sec., after accel­erating uniformly from zero velocity a~ acceleration, ft. per sec.2 This equation could relate the takeoff distance in terms of the takeoff velocity and acceleration when the airplane is accelerated uniformly from zero velocity to the final takeoff velocity. Also, this expression could relate the landing distance in terms of the landing velocity and deceleration when the airplane is accelerated (negatively) from the landing velocity to a complete stop. It is important to note that

TAKEOFF OR LANDING DISTANCE, FT.

Figure 2.31. Relationship of Velocity, Acceleration, and Distance for Uniformly Accelerated Motion

the distance varies directly as the square of the velocity and inversely as the acceleration.

As an example of this relationship, assume that during takeoff an airplane is accelerated uniformly from zero velocity to a takeoff velocity of 150 knots (253 5 ft. per sec.) with an acceleration of 6.434 ft. per sec.1 (or, 0.2g, since g = 32.17 ft. per sec.1). The takeoff distance would be:

„ (253-5)2 (2X6.434)

= 5,000 ft.

If the acceleration during takeoff were reduced 10 percent, the takeoff distance would increase

11.1 percent; if the takeoff velocity were increased 10 percent, the takeoff distance would increase 21 percent. These relation­ships point to the fact that proper accounting must be made of altitude, temperature, gross weight, wind, etc. because any item affecting acceleration or takeoff velocity will have a definite effect on takeoff distance.

If an airplane were to land at a velocity of 150 knots and be decelerated uniformly to a stop with the same acceleration of 0.2g, the landing stop distance would be 5,000 ft. However, the case is not necessarily that an aircraft may have identical takeoff and landing performance but the principle illustrated is that distance is a function of velocity and accelera­tion. As before, a 10 percent lower accelera­tion increases stop distance 11.1 percent, and a 10 percent higher landing speed increases landing distance 21 percent.

The general relationship of velocity, accel­eration, and distance for uniformly accelerated motion is illustrated by figure 2.31. In this illustration., acceleration distance is shown as a function of velocity for various values of acceleration.

TAKEOFF PERFORMANCE. The mini­mum takeoff distance is of primary interest in the operation of any aircraft because it defines the runway requirements. The minimum take­off distance is obtained by takeoff at some minimum safe velocity which allows sufficient margin above stall and provides satisfactory control and initial rate of climb. Generally, the takeoff speed is some fixed percentage of the stall speed or minimum control speed for the airplane in the takeoff configuration. As such, the takeoff will be accomplished at some particular value of lift coefficient and angle of attack. Depending on the airplane character­istics, the takeoff speed will be anywhere from 1.05 to 1.25 times the stall speed or minimum control speed. If the takeoff speed is specified as 1.10 times the stall speed, the takeoff lift coefficient is 82.6 percent of CLmax and the angle of attack and lift coefficient for takeoff are fixed values independent of weight, altitude, wind, etc. Hence, an angle of attack indicator can be a valuable aid during takeoff.

To obtain minimum takeoff distance at the specified takeoff velocity, the forces which act on the aircraft must provide the maximum acceleration during the takeoff roll. The various forces acting on the aircraft may or may not be at the control of the pilot and various techniques may be necessary in certain airplanes to maintain takeoff acceleration at the highest value.

Figure 2.32 illustrates the various forces which act on the aircraft during takeoff roll. The powerplant thrust is the principal force to provide the acceleration and, for minimum takeoff distance, the output thrust should be at a maximum. Lift and drag are produced as soon as the airplane has speed and the values of lift and drag depend on the angle of attack and dyhamic. pressure. Rolling friction results when there is a norinal force on the wheels and the friction force is the product of the normal force and the coefficient of rolling friction. The normal force pressing the wheels against the runway surface is the net of weight and lift while the rolling friction coefficient is a function of the tire type and runway surface texture.

The acceleration of the airplane at any instant during takeoff roll is a function of the net accelerating force and the airplane mass. From Newton’s second law of motion:

a=Fn/M

or

a=g(Fn/W~)

where

a = acceleration, ft, per se<

F«=net accelerating force,

W= weight, lbs. g=gravitational accelerat = 32,17 ft. per sec.2 iVf = mass, slugs = Wjg

The net accelerating force on the airplane, F„, is the net of thrust, T, drag, D, and rolling friction, F. Thus, the acceleration at any instant during takeoff roll is:

Figure 2.32 illustrates the typical variation of the various forces acting on the aircraft throughout the takeoff roll. If it is assumed that the aircraft is at essentially constant angle of attack during takeoff roll, CL and CD are constant and the forces of lift and drag vary as the square of the speed. For the case of uniformly accelerated motion, distance along the takeoff roll is proportional also to the square of the velocity hence velocity squared and distance can be used almost synon – omously. Thus, lift and drag will vary line­arly with dynamic pressure (q) or V2 from the point of beginning takeoff roll. As the rolling friction coefficient is essentially un­affected by velocity, the rolling friction will vary as the normal force on the wheels. At zero velocity, the normal force on the wheels is equal to the airplane weight but, at takeoff velocity, the lift is equal to the weight and the normal force is zero. Hence, rolling fric­tion decreases linearly with q or V2 from the beginning of takeoff roll and reaches zero at the point of takeoff.

The total retarding force on the aircraft is the sum of drag and rolling friction (D+F) and, for the majority of configurations, this sum is nearly constant or changes only slightly during the takeoff roll. The net accelerating force is then the difference between the power – plant thrust and the total retarding force,

Fn = T—D—F

The variation of the net accelerating force throughout the takeoff roll is shown in figure 2.32. The typical propeller airplane demon­strates a net accelerating force which decreases with velocity and the resulting acceleration is initially high but decreases throughout the takeoff roll. The typical jet airplane demon­strates a net accelerating force which is essen­tially constant throughout the takeoff roll. As a result, the takeoff performance of the typical turbojet airplane will compare closely with the case for uniformly accelerated motion.

The pilot technique required to achieve peak acceleration throughout takeoff roll can vary considerably between airplane configurations. In some instances, maximum acceleration will be obtained by allowing the airplane to remain in the three-point attitude throughout the roll until the airplane simply reaches lift-equal-to – weight and flies off the ground. Other air­planes may require the three-point attitude until the takeoff speed is reached then rotation to the takeoff angle of attack to become air­borne. Still other configurations may require partial or complete rotation to the takeoff angle of attack prior to reaching the takeoff speed. In this case, the procedure may be necessary to provide a smaller retarding force (D-f F) to achieve peak acceleration. When­ever any form of pitch rotation is necessary the pilot must provide the proper angle of attack since an excessive angle of attack will cause excessive drag and hinder (or possibly pre­clude) a successful takeoff. Also, insufficient rotation may provide added rolling resistance or require that the airplane accelerate to some excessive speed prior to becoming airborne.

Revised January 1965

iW* vx ywx

In this sense, an angle of attack indicator is especially useful for night or instrument takeoff conditions as well as the ordinary day VFR takeoff conditions. Acceleration errors of the attitude gyro usually preclude accurate pitch rotation under these conditions.

FACTORS AFFECTING TAKEOFF PER­FORMANCE. In addition to the important factors of proper technique, many other vari­ables affect the takeoff performance of an air­plane. Any item which alters the takeoff velocity or acceleration during takeoff roll will affect the takeoff distance. In order to evalu­ate the effect of the many variables, the prin­cipal relationships of uniformly accelerated motion will be assumed and consideration will be given to those effects due to any nonuni­formity of acceleration during the process of takeoff. Generally, in the case of uniformly accelerated motion, distance varies directly with the square of the takeoff velocity and in­versely as the takeoff acceleration.

where

S = distance.

V= velocity

a= acceleration

<‘ condition (1) applies to some known takeoff distance, Ті, which was common to some original takeoff velocity, Vu and acceleration, ax.

condition (2) applies to some new takeoff distance, S2, which is the result of some different value of takeoff velocity, V2, or acceleration, a2.

With – this basic relationship, the effect of the many variables on takeoff distance can be approximated.

The effect of gross weight on takeoff distance is large and proper consideration of this item must be made in predicting takeoff distance. Increased gross weight can be considered to produce a threefold effect on takeoff perform­ance: (1) increased takeoff velocity, (2) greater
mass to accelerate, and (3) increased retarding force (P + F). If the gross weight increases, a greater speed is necessary to produce the greater lift to get the airplane airborne at the takeoff lift coefficient. The relationship of takeoff speed and gross weight would be as follows:

тгШ (EAS°’CAV

where •

Vi — takeoff velocity corresponding to some original weight, Wx V2 = takeoff velocity corresponding to some different weight, W2

Thus, a given airplane in the takeoff configura­tion at a given gross weight will have a specific takeoff speed (EAS or CAS") which is invariant with altitude, temperature, wind, etc. because a certain value of q is necessary to provide lift equal to weight at the takeoff Cb. As an ex­ample of the effect of a change in gross weight a 21 percent increase in takeoff weight will require a 10 percent increase in takeoff speed to support the greater weight.

A change in gross weight will change the net accelerating force, Fn> and change the mass, M, which is being accelerated. If the airplane has a relatively high thrust-to-weight ratio, the change in the net accelerating force is slight and the principal effect on accelera­tion is due to the change in mass.

To evaluate the effect of gross weight on takeoff distance, the following relationship are used:

the effect of weight on takeoff velocity is

(VA2-Wi

01 ~W>

if the change in net accelerating force is neglected, the effect of weight on accelera­tion is

£t 2 Й2 Ц71

7TWX or71=W2


the effect of these items on takeoff dis­tance is

£>/»!,У

Si wj

(at least this effect because weight will alter the net accelerating force)

This result approximates the effect of gross weight on takeoff distance for airplanes with relatively high thrust-to-weight ratios. In effect, the takeoff distance will vary at least as the square of the gross weight. For ex­ample, a 10 percent increase in takeoff gross weight would cause:

a 5 percent increase in takeoff velocity at least a 9 percent decrease in acceleration at least a 21 percent increase in takeoff distance

For the airplane with a high thrust-to-weight ratio, the increase in takeoff distance would be approximately 21 to 22 percent but, for the airplane with a relatively low thrust-to – weight ratio, the increase in takeoff distance would be approximately 25 to 30 percent. Such a powerful effect requires proper con­sideration of gross weight in predicting takeoff distance.

The effect of wind on takeoff distance is large and proper consideration also must be provided when predicting takeoff distance. The effect of a headwind is to allow the airplane to reach the takeoff velocity at a lower ground velocity while the effect of a tailwind is to require the airplane to achieve a greater ground velocity to attain the takeoff velocity. The effect of the wind on acceleration is relatively small and, for the most part, can be neglected. To evaluate the effect of wind on takeoff distance, the following relationships are used:

the effect of a headwind is to reduce the takeoff ground velocity by the amount of the headwind velocity, V„

Vt=Vi-rv

the effect of wind on acceleration is negligible,

<>i,

a%=ai or—= 1 at

the effect of these items on takeoff distance is

ГУ,-у IP

Л L Ух J

or

where

zero wind takeoff distance St— takeoff distance into the head­wind

Vm—headwind velocity Vi~ takeoff ground velocity with zero wind, or, simply, the takeoff airspeed

As a result of this relationship, a headwind which is 1G percent of the takeoff airspeed will reduce the takeoff distance 19 percent. How­ever, a tailwind (or negative headwind) which is 10 percent of the takeoff airspeed will in­crease the takeoff distance 21 percent. In the case where the headwind velocity is 50 percent of the takeoff speed, the takeoff distance would be approximately 25 percent of the zero wind takeoff distance (75 percent reduction).

The effect of wind on landing distance is identical to the effect on takeoff distance. Figure 2.33 illustrates the general effect of wind by the percent change in takeoff or land­ing distance as a function of the ratio of wind velocity to takeoff or landing speed.

NAVWEPS 00—80T—80 AIRPLANE PERFORMANCE

і

The effect of runway slope on takeoff distance is due to the component of weight along the inclined path of the airplane. A runway slope of 1 percent would provide a force com­ponent along the path of the airplane which is 1 percent of the gross weight. Of course, an upslope would contribute a retarding force component while a downslope would contri­bute an accelerating force component. For the case of the upslope, the retarding force component adds to drag and rolling friction to reduce the net accelerating force. Ordinarily, a 1 percent runway slope can cause a 2 to 4 percent change in takeoff distance depending on the airplane characteristics. The airplane with the high thrust-to-weight ratio is least affected while the airplane with the low thrust – to-weight ratio is most affected because the slope force component causes a relatively greater change in the net accelerating force.

The effect of runway slope must be consid­ered when predicting the takeoff distance but the effect is usually minor for the ordinary run­way slopes and airplanes with moderate thrust-to-weight ratios. In fact, runway slope considerations are of great significance only when the runway slope is large and the airplane has an intrinsic low acceleration, i. e., low thrust-to-weight ratio. In the ordinary case, the selection of the takeoff runway will favor the direction with an upslope and headwind rather than the direction with a downslope and tailwind.

The effect of proper takeoff velocity is important when runway lengths and takeoff distances are critical. The takeoff speeds specified in the flight handbook are generally the minimum safe speeds at which the airplane can become airborne. Any attempt to take off below the recommended speed may mean that the air­craft may stall, be difficult to control, or have very low initial rate of climb. In some cases, an excessive angle of attack may not allow the airplane to climb out of ground effect. On the other hand, an excessive airspeed at takeoff may improve the initial rate of climb and “feel” of the airplane but will produce an un­desirable increase in takeoff distance. Assum­ing chat the acceleration is essentially un­affected, the takeoff distance varies as the square of the takeoff velocity,

i?=/T? Y Ti vj

Thus, 10 percent excess airspeed would increase the takeoff distance 21 percent. In most criti­cal takeoff conditions, such an increase in takeoff distance would be prohibitive and the pilot must adhere to the recommended takeoff speeds.

The effect of pressure altitude and ambient temperature is to define primarily the density altitude and its effect on takeoff performance. While subsequent corrections are appropriate for the effect of temperature on certain items of powerplant performance, density altitude defines certain effects on takeoff performance. An increase in density altitude can produce a two-fold effect on takeoff performance: (1) in­creased takeoff velocity and (2) decreased thrust and reduced net accelerating force. If a given weight and configuration of airplane is taken to altitude above standard sea level, the airplane will still require the same dynamic pressure to become airborne at the takeoff lift coefficient. Thus, the airplane at altitude will take off at the same equivalent airspeed (EAS) as at sea level, but because of the reduced density, the true airspeed (TAS) will be greater. From basic aerodynamics, the rela­tionship between true airspeed and equivalent airspeed is as follows:

TAS 1 EAS~ffJ

where

ТЛТ=тіе airspeed EAS=equivalent airspeed n = altitude density ratio

The effect of density altitude on powerplant thrust depends much on the type of power – plant. An increase in altitude above standard sea level will bring an immediate decrease in power output for the unsupercharged or ground boosted reciprocating engine or the turbojet and turboprop engines. However, an increase in altitude above standard sea level will not cause a decrease in power output for the super­charged reciprocating engine until the altitude exceeds the critical altitude. For those power – plants which experience a decay in thrust with an increase in altitude, the effect on the net accelerating force and acceleration can be ap­proximated by assuming a direct variation with density. Actually, this assumed vari­ation would closely approximate the effect oft airplanes with high thrust-to-weight ratios. This relationship would be as follows:

£2 _ Fti’i _ _p _

ai Ffh po

where

au F«j = acceleration and net accelerating force corresponding to sea level a2, Fn2 = acceleration and net accelerating force corresponding to altitude cr= altitude density ratio

In order to evaluate the effect of these items on takeoff distance, the following relationships are used:

if an increase in altitude does not alter ac­celeration, the principal effect would be due to the greater TAS

Tl <r

where

Si=standard sea level takeoff distance T2= takeoff distance at altitude <r=altitude density ratio

if an increase in altitude reduces accelera­tion in addition to the increase Іп TAS, the combined effects would be approximated for the case of the airplane with high in­trinsic acceleration by the following:

where

ii = standard sea level takeoff distance Т2=takeoff distance at altitude <r=altitude density ratio

As a result of these relationships, it should, be appreciated that density altitude will affect takeoff performance in a fashion depending much on the powerplant type. The effect of density altitude on takeoff distance can be appreciated by the following comparison:

TABLE S-1. Approximate Effect of Altitude on Takeoff Dlstaate

Density ikitude

<r

Percent inc off (list stand»

Super­

charged

recipro­

cating

airplane

below

critical

altitude

tease it ante frt d sea le

Tur­

bojet

Sth

СЙІСС-

>m

vel

Tur­

bojet

low

СГ/И0

Sea level…………

1.000

1.000

1.000

0

0

0

1,000 ft………….

.9711

1.0298

1.0605

2.9B

6.05

9.8

2,000 ft………….

.9428

1.0605

1.125

6.05

12.5

19.9

3,000 ft………….

.9151

1.0928

1.195

9-28

19.5

3Q.1

4,000 ft………….

.8881

1.126

1.264

12.6

26.4

40.6

5,000 ft………….

.8617

1.1605

1.347

16.05

34.7

52.3

6,000 ft………….

.8359

1.1965

1.432

19.65

43.2

65.8

From the previous table, some approximate rules of thumb may be derived to illustrated the differences between the various airplane types. A 1,000-ft. increase in density altitude

will cause these approximate increases in takeoff distance:

ЪУ2 percent for the supercharged recipro­cating airplane when below critical altitude

7 percent for the turbojet with high thrust – to-weight ratio

10 percent for the turbojet with low thrust-to-weight ratio

These approximate relationships show the turbojet airplane to be much more sensitive to density altitude than the reciprocating powered airplane. This is an important fact which must be appreciated by pilots in transition from propeller type to jet type airplanes. Proper accounting of pressure altitude (field elevation is a poor substitute) and temperature is mandatory for accurate prediction of takeoff roll distance.

The most critical conditions of takeoff performance are the result of some combination of high gross weight, altitude, temperature and unfavorable wind. In all cases, it be­hooves the pilot to make an accurate predic­tion of takeoff distance from the performance data of the Flight Handbook, regardless of the runway available, and to strive for a polished, professional takeoff technique.

In the prediction of takeoff distance from the handbook data, the following primary considerations must be given:

Reciprocating powered airplane

(1) Pressure altitude and temperature— to define the effect of density altitude on distance.

(2) Gross weight—a large effect on dis­tance.

0) Specific humidity—to correct take­off distance for the power loss associated with water vapor.

(4) Wind—a large effect due to the wind or wind component along the runway.

Turbine powered airplane

(1) Pressure altitude and temperature— to define the effect of density altitude.

(2) Gross weight.

(3) Temperature—an additional correc­tion for nonstandard temperatures to ac­count for the thrust loss associated with high compressor inlet air temperature. For this correction the ambient tempera­ture at the runway conditions is appro­priate rather than the ambient temperature at some distant location.

(4) Wind.

In addition, corrections are necessary to ac­count for runway slope, engine power defi­ciencies, etc.

LANDING PERFORMANCE. In many Cases, the landing distance of an airplane will define the runway requirements for flying operations. This is particularly the case of high speed jet airplanes at low altitudes where landing distance is the problem rather than takeoff performance. The minimum landing distance is obtained by landing at some mini­mum safe velocity which allows sufficient mar­gin above stall and provides satisfactory, con­trol and capability for waveoff Generally, the landing speed is some fixed percentage of the stall speed or minimum control speed for the airplane in the landing configuration. As such, the landing will be accomplished at some particular value of lift coefficient and angle of attack. The exact value of Cl and * for landing will depend on the airplane characteristics but, once defined, the values are independent of weight, altitude, wind, etc. Thus, an angle of attack indicator can be a valuable aid during approach and landing.

To obtain minimum landing distance at the specified landing velocity, the forces which act on the airplane must provide maximum deceleration (or negative acceleration} during the landing roll. The various forces acting, on the airplane during the landing roll may require various techniques to maintain landing deceleration at the peak value.

Figure 2.34 illustrates the forces acting on the aircraft during landing roll. The power – plant thrust should be a minimum positive value, or, if reverse thrust is available, a maxi­mum negative value for minimum landing dis­tance. Lift and drag are produced as long as the airplane has speed and the values of lift and drag depend on dynamic pressure and angle of attack. Braking friction results when there is a normal force on the braking wheel surfaces and the friction force is the product of the normal force and the coefficient of braking friction. The normal force on the braking surfaces is some part of the net of weight and lift, i. e., some other part of this net may be distributed to wheels which have no brakes. The maximum coefficient of braking friction is primarily a function of the runway surface con­dition (dry, wet, icy, etc.) and rather inde­pendent of the type of tire for ordinary condi­tions (dry, hard surface runway). However, the operating coefficient of braking friction is controlled by the pilot by the use of brakes.

The acceleration of the airplane during the landing roll is negative (deceleration) and will be considered to be in that sense. At any in­stant during the landing roll the acceleration is a function of the net retarding force and the airplane mass. From Newton’s second law of motion:

a= FrjM or

a = g (FrJW)

where

a = acceleration, ft. per sec.2 (negative) Fr= net retarding force, lbs. g= gravitational acceleration, ft. per sec.2 W= weight, lbs.

Af = mass, slugs

-Wig

The net retarding force on the airplane, Fr, is the net of drag, D, braking friction, F, and thrust, T. Thus, the acceleration (negative) at any instant during the landing roll is:

a=£(D+F-T)

w

Figure 2.34 illustrates the typical variation of the various forces acting on the aircraft throughout the landing roll. If it is assumed that the aircraft is at essentially constant angle of attack from the point of touchdown, CL and CD are constant and the forces of lift and drag vary as the square of the velocity. Thus, lift and drag will decrease linearly with ^ or Vі from the point of touchdown. If the braking coefficient is maintained at the maximum value, this maximum value of coefficient of friction is essentially constant with speed and the braking friction force will vary as the normal force on the braking surfaces. As the airplane nears a complete stop, the velocity and lift approach zero and the normal force on the wheels approaches the weight of the air­plane. At this point, the braking friction force is at a maximum. Immediately after touchdown, the lift: is quite large and the normal force on the wheels is small. As a re­sult, the braking friction force is small. A common error at this point is to apply exces­sive brake pressure without sufficient normal force on the wheels. This may develop a skid with a locked wheel and cause the tire to blow out so suddenly that judicious use of the brakes is necessary.

The coefficient of braking friction can reach peak values of 0.8 but ordinarily values near

0. 5 are typical for the dry hard surface runway. Of course, a slick, icy runway can reduce the maximum braking friction coefficient to values as low as 0.2 or 0.1.’ If the entire weight of the airplane were the normal force on the brak­ing surfaces, a coefficient of braking friction of

0. 5 would produce a deceleration of jig, 16.1 ft. per sec.2 Most airplanes in ground effect rarely produce lift-drag ratios lower than 3 or 4. If the lift of the airplane were equal to the weight, an L/D — 4 would produce a decelera­tion of %g, 8 ft. per sec.2 By this comparison it should be apparent that friction braking offers the possibility of greater deceleration than airplane aerodynamic braking. To this end, the majority of airplanes operating from

dry hard surface runways will require particular techniques to obtain minimum landing dis­tance. Generally, the technique involves low­ering the nose wheel to the runway and retract­ing the flaps to increase the normal force on the braking surfaces. While the airplane drag is reduced, the greater normal force can pro­vide greater braking friction force to com­pensate for the reduced drag and the net retard­ing force is increased.

The technique necessary for minimum land­ing distance can be altered to some extent in certain situations. For example, low aspect ratio airplanes with high longitudinal control power can create very high drag at the high speeds immediate to landing touchdown. If the landing gear configuration or flap or incidence setting precludes a large reduction of CLf the normal force on the braking surfaces and braking friction force capability are rela­tively small. Thus, in the initial high speed part of the landing roll, maximum deceleration would be obtained by creating the greatest possible aerodynamic drag. By the time the aircraft has slowed to 70 or 80 percent of the touchdown speed, aerodynamic drag decays but braking action will then be effective. Some form of this technique may be necessary to achieve minimum distance for some con­figurations when the coefficient of braking friction is low (wet, icy runway) and the braking friction force capability is reduced relative to airplane aerodynamic drag.

A distinction should be made between the techniques for minimum landing distance and an ordinary landing roll with considerable excess runway available. Minimum landing distance will be obtained from the landing speed by creating a continuous peak decelera­tion of the airplane. This condition usually requires extensive use of the brakes for maxi­mum deceleration. On the other hand, an ordinary landing roll with considerable excess runway may allow extensive use of aero­dynamic drag to minimize wear and tear on the tires and brakes. If aerodynamic drag is
sufficient to cause deceleration of the airplane it can be used in deference to the brakes in the early stages of the landing roll, i. e., brakes and tires suffer from continuous, hard use but airplane aerodynamic drag is free and does not | wear out with use. The use of aerodynamic drag is applicable only for deceleration to 60 or 70 percent of the touchdown speed. At speeds less than 60 to 70 percent of the touch­down speed, aerodynamic drag is so slight as to be of little use and braking must be utilized to produce continued deceleration of the airplane.

Powerplant thrust is not illustrated on figure 2.34 for there are so many possible variations. Since the objective during the landing roll is to decelerate, the powerplant thrust should be the smallest possible positive value or largest possible negative value. In the case of the turbojet aircraft, the idle thrust of the engine is nearly constant with speed throughout the landing roll. The idle thrust is of significant magnitude on cold days | because of the low compressor inlet air temper­ature and low density altitude. Unfortu­nately, such atmospheric conditions usually have the corollary of poor braking action be­cause of ice or water on the runway. The thrust from a windmilling propeller with the engine at idle can produce large negative thrust early in the landing roll but the negative force decreases with speed. The large negative thrust at high speed is valuable in adding to drag and braking friction to increase the net retarding force.

Various devices can be utilized to provide greater deceleration of the airplane or to mini­mize the wear and tear on tires and brakes. The drag parachute can provide a large retard­ing force at high ^ and greatly increase the de­celeration during the initial phase of landing roll. It should be noted that the contribution of the drag chute is important only during the high speed portion of the landing roll. For maximum effectiveness, the drag chute must be deployed immediately after the airplane is in contact with the runway. Reverse thrust of

propellers is obtained by rotating the blade angle well below the low pitch stop and applying engine power. The action is to ex­tract a large amount of momentum from the airstream and thereby create negative thrust. The magnitude of the reverse thrust from pro­pellers is very large, especially in the case of the turboprop where a very large shaft power can be fed into the propeller. In the case of reverse propeller thrust, maximum effective­ness is achieved by use immediately after the airplane is in contact with the runway. The reverse thrust capability is greatest at the high speed and, obviously, any delay in pro­ducing deceleration allows runway to pass by at a rapid rate. Reverse thrust of turbojet engines will usually employ some form of vanes, buckets, or clamshells in the exhaust to turn or direct the exhaust gases forward. Whenever the exit velocity is less than the in­let velocity for negative), a negative momen­tum change occurs and negative thrust is produced. The reverse jet thrust is valuable and effective but it should not be compared with the reverse thrust capability of a com­parable propeller powerplant which has the high intrinsic thrust at low velocities. As with the propeller reverse thrust, jet reverse thrust must be applied immediately after ground contact for maximum effectiveness in reducing landing distance.

FACTORS AFFECTING LANDING PER­FORMANCE, In addition to the important factors of proper technique, many other vari­ables affect the landing performance of an air­plane. Any item which alters the landing velocity or deceleration during landing roll will affect the landing distance. As with takeoff performance, the relationships of uni­formly accelerated motion will be assumed applicable for studying the principal effects on landing distance. The case of uniformly ac­celerated motion defines landing distance as varying directly as the square of the landing velocity and inversely as the acceleration dur­ing landing roll.

where

= landing distance resulting from certain values of landing velocity, V, and acceleration, a

Tj = landing distance resulting from some different values of landing velocity, V2, or acceleration, a2

With this relationship, the effect of the many variables on landing distance can be approxi­mated.

The effect of gross weight on landing distance is one of the principal items determining the landing distance of an airplane One effect of an increased gross weight is that the airplane will require a greater speed to support the airplane at the landing angle of attack and lift coefficient. The relationship of land­ing speed and gross weight would be as follows:

vr^W, or CAS)

where

Fi = landing velocity corresponding to some original weight,

V2 = landing velocity corresponding to some different weight, W2 Thus, a given airplane in the landing con­figuration at a given gross weight will have a specific landing speed (EAS or CAS) which is invariant with altitude, temperature, wind, etc., because a certain value of q is necessary to provide lift equal to weight at the landing CL. As an example of the effect of a change in gross weight, a 21 percent increase in landing weight will require a 10 percent increase in landing speed to support the greater weight.

When minimum landing distances are con­sidered, braking friction forces predominate during the landing roll and, for the majority of airplane configurations, braking friction is the main source of deceleration. In this case, an increase in gross weight provides a greater

normal force and increased braking friction force to cope with the increased mass. Also, the higher landing speed at the same CL and CD produce an average drag which increased in the same proportion as the increased weight. Thus, increased gross weight causes like in­creases in the sum of drag plus braking friction and the acceleration is essentially unaffected.

To evaluate the effect of gross weight on landing distance, the following relationships are used;

the effect of weight on landing velocity is

Уг fWi Vi W1

if the net retarding force increases in the same proportion as the weight, the accel­eration is unaffected.

the effect of these items on landing dis­tance is,

or

S2 _Wi

In effect, the minimum landing distance will vary directly as the gross weight. For ex­ample, a 10 percent increase in gross weight at landing would cause:

a 5 percent increase in landing velocity a 10 percent increase in landing distance A contingency of the previous analysis is the relationship between weight and braking fric­tion force. The maximum coefficient of brak­ing friction is relatively independent of the usual range of normal forces and rolling speeds, e. g., a 10 percent increase in normal force would create a like 10 percent increase in braking friction force. Consider the case of two air­planes of the same type and c. g. position but of different gross weights. If these two air­planes are rolling along the runway at some speed at which aerodynamic forces are negli­gible, the use of the maximum coefficient of
braking friction will bring both airplanes to a stop in the same distance. The heavier air­plane will have the greater mass to decelerate but the greater normal force will provide a greater retarding friction force. As a result, both airplanes would have identical accelera­tion and identical stop distances from a given velocity. However, the heavier airplane would have a greater kinetic energy to be dis­sipated by the brakes and the principal differ­ence between the two airplanes as they reach a stop would be that the heavier airplane would have the hotter brakes. Therefore, one of the factors of braking performance is the ability of the brakes to dissipate energy with­out developing excessive temperatures and losing effectiveness.

To appreciate the effectiveness of modern brakes, a 30,000-lb. aircraft landing at 175 knots has a kinetic energy of 41 million ft.-lbs. at the instant of touchdown. In a minimum distance landing, the brakes must dissipate most of this kinetic energy and each brake must absorb an input power of approximately 1,200 h. p. for 25 sccohds. Such requirements for brakes are extreme but the example serves to illustrate the problems of brakes for high performance airplanes.

While a 10 percent increase in landing weight causes:

a 5 percent higher landing speed

a 10 percent greater landing distance, it also produces a 21 percent increase in the kinetic energy of the airplane to be dissipated during the landing roll. Hence, high landing weights may approach the energy dissipating capability of the brakes.

The effect of wind on landing distance is large and deserves proper consideration when pre­dicting landing distance. Since the airplane will land at a particular airspeed independent of the wind, the principal effect of wind on landing distance is due to the change in the ground velocity at which the airplane touches down. The effect of wind on acceleration during the landing distance is identical to the

effect on takeoff distance and is approximated by the following relationship:

where

Ti=Zero wind landing distance ^2= landing distance into a headwind Fw = headwind velocity Vi= landing ground velocity with zero wind or, simply, the landing airspeed

As a result of this relationship, a headwind which is 10 percent of the landing airspeed will reduce the landing distance 19 percent but a tailwind (or negative headwind) which is 10 percent of the landing speed will increase the landing distance 21 percent. Figure 2.33 illus­trates this general effect.

The effect of runway slope on landing distance is due to the component of weight along the inclined path of the airplane. The relation­ship is identical to the case of takeoff per­formance but the magnitude of the effect is not as great. While account must be made for the effect, the ordinary values of runway slope do not contribute a large effect on landing distance. For this reason, the selection of the landing runway will ordinarily favor the direc­tion with a downslope and headwind rather than an upslope and tailwind.

The effect of pressure altitude and ambient tem­perature is to define density altitude and its effect on landing performance. An increase in dens­ity altitude will increase the landing velocity but will not alter the net retarding force. If a given weight and configuration of airplane is taken to altitude above standard sea level, the airplane will still require the same q to provide lift equal to weight at the landing Ct. Thus, the airplane at altitude will land at the same equivalent airspeed (EAS) as at sea level but, because of the reduced density, the true airspeed (TAS) will be greater. The relation­ship between true airspeed and equivalent air­speed is as follows:

TAS= 1 EAS – fa

where

TAS = true airspeed EAS= equivalent airspeed <r=altitude density ratio

Since the airplane lands at altitude with the same weight and dynamic pressure, the drag and braking friction throughout the landing roll have the same values as at sea level. As long as the condition is within the capability of the brakes, the net retarding force is un­changed and the acceleration is the same as with the landing at sea level.

To evaluate the effect of density altitude on landing distance, the following relationships are used:

since an increase in altitude does not alter acceleration, the effect would be due to the greater TAS

Ji=l Ті а

where

Ті = standard sea level landing dis­tance

T2 = landing distance at altitude <r=altitude density ratio

From this relationship, the minimum land­ing distance at 5,000 ft. (<r=0.8617) would be 16 percent greater than the minimum landing distance at sea level. The approximate increase in landing distance with altitude is approxi­mately 3K percent for each 1,000 ft. of altitude. Proper accounting of density altitude is neces­sary to accurately predict landing distance.

The effect of proper landing velocity is impor­tant when runway lengths and landing dis­tances are critical. The landing speeds specified in the flight handbook are generally the mini­mum safe speeds at which the airplane can be landed. Any attempt to land at below the specified speed may mean that the airplane may stall, be difficult to control, or develop high rates of descent. On the other hand, an exces­sive speed at landing may improve the control­lability (especially in crosswinds) but will cause an undesirable increase in landing dis­tance. The principal effect of excess landing speed is described by:

Si

Thus, a 10 percent excess landing speed would cause a 21 percent increase in landing distance. The excess speed places a greater working load on the brakes because of the additional kinetic energy to be dissipated. Also, the additional speed causes increased drag and lift in the nor­mal ground attitude and the increased lift will reduce the normal force on the braking sur­faces. The acceleration during this range of speed immediately after touchdown may suffer and it will be more likely that a tire can be blown out from braking at this point. As a result, 10 percent excess landing speed will cause at least 21 percent greater landing dis­tance.

The most critical conditions of landing per­formance are the result of some combination of high gross weight, density altitude, and un­favorable wind. These conditions produce the greatest landing distance and provide critical levels of energy dissipation required of the brakes. In all cases, it is necessary to make an accurate prediction of minimum landing dis­tance to compare with the available runway. A polished, professional lajiding technique is necessary because the landing phase of flight accounts for more pilot caused aircraft acci­dents than any other single phase of flight.

In the prediction of minimum landing dis­tance from the handbook data, the following considerations must be given:

(1) Pressure altitude and temperature—to define the effect of density altitude.

(2) Gross weight—which define the CAS or EAS for landing.

СЗ) Wind—a large effect due to wind or wind component along the runway.

(4) Runway slope—a relatively small cor­rection for ordinary values of runway slope. IMPORTANCE OF HANDBOOK PER­FORMANCE DATA. The performance sec­tion or supplement of the flight handbook con­tains all the operating data for the airplane. For example, all data specific to takeoff, climb, range, endurance, descent and landing are in­cluded in this section. The ordinary use of these data in flying operations is mandatory and great knowledge and familiarity of the air­plane can be gained through study of this material. A complete familiarity of an air­plane’s characteristics can be obtained only through extensive analysis and study of the handbook data.

ENDURANCE PERFORMANCE

The ability of the airplane to convert fuel energy into flying time is an important factor in flying operati on s. The ‘ ‘ specific end urance ’ ’ of the airplane is defined as follows:

specific endurance—тг^-7-r—j lb. of fuel

_________ 1________

fuel flow, lbs. per hr.

The specific endurance is simply the reciprocal of the fuel flow, hence maximum endurance conditions would be obtained at the lowest fuel flow required to hold the airplane in steady level flight. Obviously, minimum fuel flow will provide the maximum flying time from a given quantity of fuel. Generally, in subsonic performance, the speed at which maximum en­durance is achieved is approximately 75 per­cent of the speed for maximum range.

While many different factors can affect the specific endurance, the most important factors at the control of the pilot are the configuration and operating altitude. Of course, for maxi­mum endurance conditions the airplane must be in the clean configuration and operated at the proper aerodynamic conditions.

EFFECT OF ALTITUDE ON ENDUR­ANCE, PROPELLER DRIVEN AIRPLANES. Since the fuel flow of the propeller driven air­plane is proportional to power required, the propeller powered airplane will achieve maxi­mum specific endurance when operated at mini­mum power required. The point of minimum power required is obtained at a specific value of lift coefficient for a particular airplane con­figuration and is essentially independent of weight or altitude. However, an increase in altitude will increase the value of the minimum power required as illustrated by figure 2.27. If the specific fuel consumption were not in­fluenced by altitude or engine power, the spe­cific endurance would be directly proportional to v’ff, e. g., the specific endurance at 22,000 ft. (jt=0.498) would be approximately 70 percent of the value at sea level. This example is very nearly the case of the airplane with the recipro­cating engine since specific fuel consumption and propeller efficiency are not directly affected by altitude. The obvious conclusion is that maximum endurance of the reciprocating en­gine airplane is obtained at the lowest practical altitude.

The variation with altitude of the maximum endurance of the turboprop airplane requires consideration of powerplant factors in addition

to airplane factors. The turboprop power – plant prefers operation at low inlet air tem­peratures and relatively high power setting to produce low specific fuel consumption. While an increase in altitude will increase the mini­mum power required for the airplane, the powerplant achieves more efficient operation. As a result of these differences, maximum en­durance of the multiengine turboprop airplane at low altitudes may require shutting down some of the powerplants in order to operate the remaining powerplants at a higher, more efficient power setting.

EFFECT OF ALTITUDE ON ENDUR­ANCE, TURBOJET AIRPLANES. Since the fuel flow of the turbojet powered airplane is proportional to thrust required, the turbojet airplane will achieve maximum specific endur­ance when operated at minimum thrust re­quired or (L/D)™*. In subsonic flight, (L/D)m« occurs at a specific value of lift coefficient for a given airplane and is essentially independent of weight or altitude. If a given weight arid configuration of airplane is oper­ated at various altitudes, the value of the minimum thrust required is unaffected by the curves of thrust required versus velocity shown in figure 2.27. Hence, it is apparent that the aerodynamic configuration has no preference for altitude (within compressibility limits) and specific endurance is a function only of engine performance.

The specific fuel consumption of the turbojet engine is strongly affected by operating RPM and altitude. Generally, the turbojet engine prefers the operating range near normal rated engine speed and the low temperatures of the stratosphere to produce low specific fuel con­sumption. Thus, increased altitude provides the favorable lower inlet air temperature and requires a greater engine speed to provide the thrust required at (X/D)™*. The typical turbojet airplane experiences an increase in specific endurance with altitude with the peak values occurring at or near the tropopause. For example, a typical single-engine turbojet airplane will have a maximum specific endur­ance at 35,000 ft. which is at least 40 percent greater than the maximum value at sea level. If the turbojet airplane is at low altitude and it is necessary to hold for a considerable time, maximum time in the air will be obtained by beginning a climb to some optimum altitude dependent upon the fuel quantity available. Even though fuel is expended during the climb, the higher altitude will provide greater total endurance. Of course, the use of afterburner for the climb would produce a prohibitive re­duction in endurance.

OFF-OPTIMUM RANGE AND ENDUR­ANCE

There are many conditions of flying oper­ations in which optimum range or endurance conditions are not possible or practical. In many instances, the off-optimum conditions result from certain operational requirements or simplification of operating procedure. In addition, off-optimum performance may be the result of a powerplant malfunction or failure. The most important conditions are discussed for various airplanes by powerplant type.

RECIPROCATING POWERED AIR­PLANE. In the majority of cases, the recipro­cating powered airplane is operated at an engine dictated cruise. Service use will most probably define some continuous power setting which will give good service life and trouble-free operation of the powerplant. When range or endurance is of no special interest, the simple expedient is to operate the powerplant at the recommended power setting and accept what­ever speed, range, or endurance that results. While such a procedure greatly simplifies the matter of cruise control, the practice does not provide the necessary knowledge required for operating a high performance, long range airplane.

The failure of an engine on the multiengine reciprocating powered airplane has interesting ramifications. The first problem appearing is to produce sufficient powtr from the remaining engines to keep the airplane airborne. The

problem will be most. critical if the airplane is at high altitude, high gross weight, and with flaps and gear extended. Lower altitude, jettisoning of weight items, and cleaning up the airplane will reduce the power required for flight. Of course, the propeller on the in­operative engine must be feathered or the power required may exceed that available from the remaining operating powerplants.

The effect on range is much dependent on the airplane configuration. When the pro­peller on the’inoperative engine is feathered, the added drag is at a minimum, but there is added the trim drag required to balance the unsymmetrical power. When both these sources of added drag are accounted for, the (LjD’)ma is reduced but not by significant amounts. Generally, if the specific fuel con­sumption and propeller efficiency do not deteri­orate, the maximum specific range is not greatly reduced. On the twin-engine airplane the power required1 must be furnished by the one remaining engine and this, usually requires more than the maximum cruise-rating of the powerplants As a result the powerplant can­not be operated in the auto-lean or manual lean power range and the specific fuel con­sumption increases. greatly1. Thus, noticeable loss of range must be anticipated when one engine fails on the twin-engine airplane. The failure of one engine on the four (or more) engine airplartir may allow the required power to be developed і by the three remaining power – plants operating in an economical power range. If the airplane is clean, at low altitude, and low gross weight, the failure of one engine is not likely to cause a loss of range. However, the loss – of two engines is likely to cause a considerable loss of range.

When engine failure produces a critical power or range situation, improved perform­ance is possible with the"airplane in the clean configuration at low altitude. Also, jetti­soning of expendable weight items will reduce the power required and improve the specific range.

TURBOPROP POWERED AIRPLANE. The turbine engine has the preference for relatively high power settings and high alti­tudes to provide low specific fuel consumption. Thus, the off-optimum conditions of range or endurance can be concerned with altitudes less than the optimum. Altitudes less than the optimum can reduce the range but the loss can be minimized on the multiengine airplane by shutting down some powerplants and operating the remaining powerplants at a higher, more efficient output. In this case the change of range is confined to the variation of specific fuel consumption with altitude.

Essentially the same situation exists in the case of engine failure when cruising at optimum altitude. If the propeller on the inoperative engine is feathered, the loss of range will be confined to the change in specific fuel con­sumption from the reduced cruise altitude. If a critical power situation exists due to engine failure, a reduction in altitude provides im­mediate benefit because of the reduction of power required and the increase in power available from the power plants. In addition, the jettisoning of expendable weight items will improve performance and, of course, the clean configuration provides minimum parasite drag.

Maximum specific endurance of the turbo­prop airplane does not vary as greatly with altitude as the turbojet airplane. While each configuration has its own particular operating requirements, low altitude endurance of the turboprop airplane requires special considera­tion. The single-engine turboprop will gen­erally experience an increase in specific endur­ance with an increase in altitude from sea level. However, if the airplane is at low altitude and must hold or endure for a period of time, the decision to begin a climb or hold the existing altitude will depend on the quantity of fuel available. The decision depends primarily on the climb fuel requirements and the variation of specific endurance with altitude. A somewhat similar problem exists with the multiengine

turboprop airplane but additional factors are available to influence the specific endurance at low altitude. In other words, low altitude endurance can be improved by shutting down some powerplants and operating the remaining powerplants at higher, more efficient power setting. Many operati onal fact ors could dec ide whether such procedure would be a suitable technique.

TURBOJET TOWERED AIRPLANE. In­creasing altitude has a powerful effect on both the range and endurance of the turbojet air­plane. As a result of this powerful effect, the typical turbojet airplane will achieve maxi­mum specific endurance at or near the tropo – pause. Also, the maximum specific range will be obtained at even higher altitudes since the peak specific range generally occurs at the highest altitude at which the normal rating of the engine can sustain the optimum aero­dynamic conditions. At low altitude cruise conditions, the engine speed necessary to sus­tain optimum aerodynamic conditions is very low and the specific fuel consumption is rela­tively poor. Thus, at low altitude, the air­plane prefers the low speeds to obtain (VC/CtDffloa: but the powerplant prefers the higher speeds common to higher engine effi­ciency. The compromise results in maximum specific range at flight speeds well above the optimum aerodynamic conditions. In a sense, low altitude cruise conditions are engine dictated.

Altitude is the one most important factor affecting the specific range of the turbojet airplane. Any operation below the optimum altitude will have a noticeable effect on the range capability and proper consideration must be given to the loss of range. In addi­tion, turbojet airplanes designed specifically for long range will have a large percent of the gross weight as fuel. The large changes in gross weight during cruise will require partic­ular methods of cruise control to extract the maximum flight range. A variation from the optimum flight path of cruise (constant Mach number, cruise-climb, or whatever the appro­priate technique) will result in a loss of range capability.

The failure of an engine during the optimum cruise of a multiengine turbojet airplane will cause a noticeable loss of range. Since the optimum cruise of the turbojet is generally a thrust-limited cruise, the loss of part of the total thrust means that the airplane must descend to a lower altitude. For example, if a twin-engine jet begins an optimum cruise at

35,0 ft. (<r=0.31) and one powerplant fails, the airplane must descend to a lower altitude so that the operative engine can provide the cruise thrust. The resulting altitude would be approximately 16,000 ft. («г=0.61). Thus, the airplane will experience a loss of the range remaining at the point of engine failure and loss could be accounted for by the reduced velocity (TAS) and the increase in specific fuel consumption OO ftom the higher ambient air temperature. In the case of the example air­plane, engine failure would cause a 30 to 40 percent loss of range from the point of engine failure. Of course, the jettisoning of expend­able weight items would allow higher altitude and would increase the specific range.

Maximum endurance in the turbojet air­plane varies with altitude but the variation is due to the changes in fuel flow necessary to provide the thrust required at (L/D)mai. The low inlet air temperature of the tropopause and the greater engine speed reduce the specific fuel consumption to a minimum. If the single­engine turbojet airplane is at low altitude and must hold or endure for a period of time, a climb should begin to take advantage of the higher specific endurance at higher altitude. The altitude to which to climb will be deter­mined by the quantity of fuel remaining. In the case of the multiengine turbojet at low altitude, some slightly different procedure may be utilized. If all powerplants are oper­ating, it is desirable to climb to a higher altitude which is a function of the remaining fuel quantity. An alternative at low altitude

would, be to provide the endurance thrust with some engineCO shut down and the remaining engine(s) operating at a more efficient power output. This technique would cause a mini­mum loss of endurance if at low altitude. The feasibility of such a procedure is dependent on many operational factors.

In all cases, the airplane should be in the cleanest possible external configuration because the specific endurance is directly proportional to the (LjD).

MANEUVERING PERFORMANCE y,.

When the airplane is in turning flight, the airplane is not in static equilibrium for there must be developed the unbalance of force to produce the acceleration of the turn. During a steady coordinated turn, the lift is inclined to produce a horizontal component of force to equal the centrifugal force of the turn. In addition, the steady turn is achieved by pro­ducing a vertical component of lift which is equal to the weight of the airplane. Figure 2.28 illustrates the forces which act on the airplane in a steady, coordinated tufn.

For the case of the steady, coordinated turn, the vertical component of lift musr <=qual the weight of the aircraft so that there will be no acceleration in the vertical direction. This requirement leads to the following relation­ship:

From this relationship it is apparent that the steady, coordinated turn requires specific values of load factor, n, at various angles of bank, ф. For example, a bank angle of 60° requires a load factor of 2.0 (cos 60° = 0.5 or sec 60° = 2.0) to provide the steady, coordinated turn. If the airplane were at a 60° bank and lift were not provided to produce the exact load factor of 2.0, the aircraft would be accelerating in the vertical direction as well as the horizontal di­rection and the turn would not be steady. Also, any sideforce on the aircraft due to sideslip, etc., would place the resultant aero­dynamic force out of the plane of symmetry perpendicular to the lateral axis and the turn would not be coordinated.

As a consequence of the increase lift re­quired to produce the steady turn in a bank, the induced drag is increased above that in­curred by steady, wing level, lift-equal-weight flight. In a sense, the increased lift required in a steady turn will increase the total drag or power required in the same manner as increased gross weight in level flight. The curves of figure 2.28 illustrate the general effect of turn­ing flight on the total thrust and power re­quired. Of course, the change in thrust re­quired at any given speed is due to the change in induced drag and the magnitude of change depends on the value of induced drag in level flight and the angle of bank in turning flight. Since the induced drag generally varies as the square of CL, the following data provide an illustration of the effect of various degrees of bank:

Bank angle, ф

Load factor, n

Percent increase in induced drag from level flight

0°……………………………………………

1.000

0 (of course)

15°…………………………………………

1.036

7.2

30°………………………………………….

1.154

33.3

45°……………………………………… :.

1.414

100.0

60°………………………………………….

2.000

300.0

Since the. induced drag predominates at low speeds, steep turns at low speeds can produce significant increases in thrust or power required to maintain altitude. Thus, steep turns must be avoided after takeoff, during approach, and especially during a critical power situation from failure or malfunction of a powerplant. The greatly increased induced drag is just as

important—if not more important—as the increased stall speed in turning flight. It is important also that any turn be well coordi­nated to prevent the increased drag attendant to a sideslip.

TURNING PERFORMANCE. The hori­zontal component of Lift will equal the centrif­ugal force of steady, turning flight. This fact allows development of the following relation­ships of turning performance:

turn radius

_ Vі Г 11.26 tan ф

where

return radius, ft.

V= velocity, knots (TAS)

0 = bank angle, degrees

turn rate

R0T= 1-°9уаП ф

where

ROT= rate of turn, degrees per sec. tf>=bank angle, degrees V— velocity, knots, TAS

These relationships define the turn radius, r, and rate of turn, ROT, as functions of the two principal variables; bank angle, Ф, and velocity, V (TAS’). Thus, when the airplane is flown in the steady, coordinated turn at specific values of bank angle and velocity, the turn rate and turn radius are fixed and independent of the airplane type. As an example, an air­plane in a steady, coordinated turn at a bank angle of 45° and a velocity of 250 knots (TAS) would have the following turn performance:

_ (250)2

Г (U.26)(l.000)

= 5,550 ft.

and (1,091)0.000)

250

= 4.37 deg. per sec.

If the airplane were to hold the same angle of bank at 500 knots (TAS), the turn radius would quadruple (r= 22,200 ft.) and the turn rate would be one-half the original value (R0T= 2.19 deg. per sec.).

Values of turn radius and turn rate versus velocity are shown in figure 2.29 for various angles of bank and the corresponding load factors. The conditions are for the steady, coordinated turn at constant altitude but the results are applicable for climbing or descend­ing flight when the angle of climb or descent is relatively small. While the effect of alti­tude on turning performance is not immediately apparent from these curves, the principal effect must be appreciated as an increased true air­speed (TAS) for a given equivalent airspeed (EAS).

TACTICAL PERFORMANCE. Many tac­tical maneuvers require the use of the maxi­mum turning capability of the airplane. The maximum turning capability of an airplane will be defined by three factors:

(1) Maximum lift capability. The combi­nation of maximum lift coefficient,

and wing loading, WjS, will define the ability of the airplane to develop aero­dynamically the load factors of maneuvering flight.

(2) Operating strength limits will define the upper limits of maneuvering load factors which will not damage the primary struc­ture of the airplane. These limits must not be exceeded in normal operations because of the possibility of structural damage or failure.

(3) Thrust or power limits will define the ability of the airplane to turn at constant altitude. The limiting condition would al­low increased load factor and induced drag until the drag equals the maximum thrust available from the powerplant. Such a case would produce the maximum turning capa­bility for maintaining constant altitude.

The first illustration of figure 2.30 shows

how the aerodynamic and structural limits

define the maximum turning performance. The aerodynamic limit describes the minimum turn radius available to the airplane when operated at Cbmax – When the airplane is at the stall speed in level flight, all the lift is neces­sary to sustain the aircraft in flight and none is available to produce a steady turn. Hence, the turn radius at the stall speed is infinite. As speed is increased above the stall speed, the airplane at Сі^ая is able to develop lift greater than weight and produce a finite turn radius. For example, at a speed twice the stall speed, the airplane at CLmax is able to develop a load factor of four and utilize a bank angle of 75-5° Ceos 75-5° = 0.25). Continued increase in speed increases the load factor and bank angle which is available aerodynamically but, be­cause of the increase in velocity and the basic effect on turn radius, the turn radius approaches an absolute minimum value. When Ct^ax is unaffected by velocity, the aerodynamic mini­mum turn radius approaches this absolute value which is a function of CLpiaxt WjS, and <r. Actually, the one common denominator of aerodynamic turning performance is the wing level stall speed.

The aerodynamic limit of turn radius requires that the increased velocity be utilized to pro­duce increasing load factors and greater angles of bank. Obviously, very high speeds will require very high load factors and the absolute aerodynamic minimum turn radius will require an infinite load factor. Increasing speed above the stall speed will eventually produce the limit load factor and continued increase in speed above this point will require that load factor and bank angle be limited to prevent structural damage. When the load factor and bank angle are held constant at the structural limit, the turn radius varies as the square of the velocity and increases rapidly above the aerodynamic limit. The intersection of the aerodynamic limit and structural limit lines is the “maneuver speed.” The maneuver speed is the minimum speed necessary to develop aerodynamically the limit load factor

and it produces the minimum turn radius within aerodynamic and structural limitations. At speeds less than the maneuver speed, the limit load factor is not available aerodynami­cally and turning performance is aerody­namically limited. At speeds greater than the maneuver speed, Cl^ and maximum aerodynamic load factor are not available and turning performance is structurally limited. When the stall speed and limit load factor are known for a particular configuration, the maneuver speed is related by the following expression:

Vp = V,^n limit

where

Vp = maneuver speed, knots

V„ = stall speed, knots

n limit = limit load factor

For example, an airplane with a limit load factor of 4.0 would have a maneuver speed which is twice the stall speed.

The aerodynamic limit line of the first illustration of figure 2.30 is typical of an air­plane with a CLmax which is invariant with speed. While this is applicable for the ma­jority of subsonic airplanes, considerable differ­ence would be typical of the transonic or supersonic airplane at altitude. Compressi­bility effects and changes in longitudinal control power may produce a maximum avail­able CL which varies with velocity and an aerodynamic turn radius which is not an absolute minimum at the maximum of velocity.

The second illustration of figure 2.30 describes the constant altitude turning performance of an airplane. When an airplane is at high altitude, the turning performance at the high speed end of the flight speed range is more usually thrust limited rather than structurally limited. In flight at constant altitude, the thrust must equal the drag to maintain equilib­rium and, thus, the constant altitude turn radius is infinite at the maximum level flight speed. Any bank or turn at maximum level flight speed would incut additional drag and

TURN

RADIUS

Г

FT

VELOCITY, KNOTS (TAS)

cause the airplane to descend. However, as speed is reduced below the maximum level flight speed, parasite drag reduces and allows increased load factors and bank angles and reduced radius of turn, i. e., decreased parasite drag allows increased induced drag to accom­modate turns within the maximum thrust available. Thus, the considerations of con­stant altitude will increase the minimum turn radius above the aerodynamic limit and define a particular airspeed for minimum turn radius.

Each of the three limiting factors (aero­dynamic, structural, and power) may combine to define the turning performance of an air­plane. Generally, aerodynamic and structural limits predominate at low altitude while aero­dynamic and power limits predominate at high altitude. The knowledge of this turning per­formance is particularly necessary for effective operation of fighter and interceptor types of airplanes.

RANGE PERFORMANCE

The ability of an airplane to convert fuel energy into flying distance is one of the most important items of airplane performance. The problem of efficient range operation of an air­plane appears of two general forms in flying operations: (T) to extract the maximum flying distance from a given fuel load or (2) to fly a specified distance with minimum expenditure of fuel. An obvious common denominator for each of these operating problems is the “spe­cific range," nautical miles of flying distance per lb. of fuel. Cruise flight for maximum range conditions should be conducted so that the airplane obtains maximum specific range throughout the flight.

GENERAL RANGE PERFORMANCE. The principal items of range performance can be visualized by use of the illustrations of figure

2.23. From the characteristics of the aero­dynamic configuration and the powerplant, the
conditions of steady level flight will define various rates of fuel flow throughout the range of flight speed. The first graph of figure 2.23 illustrates a typical variation of fuel flow versus velocity. The specific range can be defined by the following relationship:

nautical miles lbs. of fuel

or,

.r nautical miles/hr.

specific range – lbs.„ffadTht.

thus,

.ґ velocity, knots

specific range-^д^^^-

If maximum specific range is desired, the flight condition must provide a maximum of velocity fuel flow. This particular point would be located by drawing a straight line from the origin tangent to the curve of fuel flow versus velocity.

The general item of range must be clearly distinguished from the item of endurance. The item of range involves consideration of flying distance while endurance involves consideration of flying time. Thus, it is appropriate to define a separate term, “specific endurance.”

flight hours lb. of fuel

or,

flight hours/hr. lbs. of fuel/hr.

________ 1________

fuel flow, lbs. per hr.

By this definition, the specific endurance is simply the reciprocal of the fuel flow. Thus, if maximum endurance is desired, the flight condition must provide a minimum of fuel flow. This point is readily appreciated as the lowest point of the curve of fuel flow versus velocity. Generally, in subsonic performance, the speed at which maximum endurance is

GROSS WEIGHT LBS.

Figure 2.23. GeneraI Range Performance


obtained is approximately 75 percent of the speed for maximum range.

A more exact analysis of range may be ob­tained by a plot of specific range versus velocity similar to the second graph of figure 2.23. Of course, the source of these values of specific range is derived by the proportion of velocity and fuel flow from the previous curve of fuel flow versus velocity. The maximum specific range of the airplane is at the very peak of the curve. Maximum endurance point is located by a straight line from the origin tangent to the curve of specific range versus velocity. This tangency point defines a maximum of (nmi/lb-) per (nmi/hr.) or simply a maximum of (hrs./lb.).

While the very peak value of specific range would provide maximum range operation, long range cruise operation is generally recom­mended at some slightly higher airspeed. Most long range cruise operation is conducted at the flight condition which provides 99 per­cent of the absolute maximum specific range. The advantage of such operation is that 1 percent of range is traded for 3 to 5 percent higher cruise velocity. Since the higher cruise speed has a great number of advantages, the small sacrifice of range is a fair bargain. The curves of specific range versus velocity are affected by three principal variables: airplane gross weight, altitude, and the external aero­dynamic configuration of the airplane. These curves are the source of range and endurance operating data and are included in the per­formance section of the flight handbook.

"Cruise control” of an airplane implies that the airplane is operated to maintain the recom­mended long range cruise condition through­out the flight. Since fuel is consumed during cruise, the gross weight of the airplane will vary and optimum airspeed, altitude, and power setting can vary, Generally, "cruise control" means the control of optimum air­speed, altitude, and power setting to maintain the 99 percent maximum specific range condi­tion. At the beginning of cruise, the high initial weight of the airplane will require spe­cific values of airspeed, altitude, and power setting to produce the recommended cruise condition. As fuel is consumed and the air­plane gross weight decreases, the optimum air­speed and power setting may decrease or the optimum altitude may increase. Also, the optimum specific range will increase. The pilot must provide the proper cruise control technique to ensure that the optimum condi­tions are maintained.

The final graph of figure 2.23 shows a typical variation of specific range with gross weight for some particular cruise operation. At the beginning of cruise the gross weight is high and the specific range is low. As fuel is con­sumed, and the gross weight reduces, the specific range increases. This type of curve relates the range obtained by the expenditure of fuel by the crosshatched area between the gross weights at beginning and end of cruise. For example, if the airplane begins cruise at 18,500 lbs. and ends cruise at 13,000 lbs., 5,500 lbs. of fuel is expended. If the average spe­cific range were 0.2 nmi/lb., the total range would be:

range=(0.2) (5,500) lb.

= 1,100 nmi.

Thus, the total range is dependent on both the fuel available and the specific range. When range and economy of operation predominate, the pilot must ensure that the airplane will be operated at the recommended long range cruise condition. By this procedure, the airplane will be capable of its maximum design operat­ing radius or flight distances less than the maximum can be achieved with a maximum of fuel reserve at the destination.

RANGE, PROPELLER DRIVEN AIR­PLANES. The propeller driven airplane com­bines the propeller with the reciprocating engine or the gas turbine for propulsive power. In the case of either the reciprocating engine or the gas turbine combination, powerplant fuel

flow is determined mainly by the shaft power put into the propeller rather than thrust. Thus, the powerplant fuel flow could be related di­rectly to power required to maintain the air­plane in steady, level flight. This fact allows study of the range of the propeller powered airplane by analysis of the curves of power required versus velocity.

Figure 2.24 illustrates a typical curve of power required versus velocity which, for the propeller powered airplane, would be analo­gous to the variation of fuel flow versus veloc­ity. Maximum endurance condition would be obtained at the point of minimum power re­quired since this would require the lowest fuel flow to keep the airplane in steady, level flight. Maximum range condition would occur where the proportion between velocity and power re­quired is greatest and this point is located by a straight line from the origin tangent to the curve.

The maximum range condition is obtained at maximum lift-drag ratio and it is important to note that (L! D’)max for a given airplane configuration occurs at a particular angle of attack and lift coefficient and is unaffected by weight or altitude (within compressibility limits). Since approximately 50 percent of the total drag at (L/jD)7MJ. is induced drag, the propeller powered airplane which is designed specifically fo’r long range will have a strong preference for the high aspect ratio planform.

The effect of the variation of airplane gross weight is illustrated by the second graph of figure 2.24. The flight condition of (L/D)mai is achieved at one particular value of lift coeffi­cient for a given airplane configuration. Hence, a variation of gross weight will alter the values of airspeed, power required, and spe­cific range obtained at (L/D)MBI. If a given configuration of airplane is operated at con­stant altitude and the lift coefficient for (L/D)raor, the following relationships will apply:

where

condition (2) applies to some known condi­tion of velocity, power required, and specific range for (L/D)m« at some basic weight, W

condition (2) applies to some new values of velocity, power required, and specific range for (L/D)mar at some different weight, W2

and,

V— velocity, knots W= gross weight, Ibs – Pr — power required, h. p.

SR— specific range, nmi/lb.

Thus a 10 percent increase in gross weight would create:

a 5 percent increase in velocity a 15 percent increase in power required a 9 percent decrease in specific range when flight is maintained at the optimum con­ditions of (L/D)mar. The variations of veloc­ity and power required must be monitored by the pilot as part of the cruise control to main­tain (L/D)„aj. When the airplane fuel weight is a small part of the gross weight and the range is small, the cruise control procedure can be simplified to essentially a constant speed and power setting throughout cruise. However, the long range airplane has a fuel weight which is a considerable part of the gross weight and cruise control procedure must employ sched­uled airspeed and power changes to maintain optimum range conditions.

The effect of altitude on the range of the propeller powered airplane may be appreciated by inspection of the final graph of figure 2.24. If a given configuration of airplane is operated at constant gross weight and the lift coefficient

for (LjD’)max, a change in altitude will produce the following relationships:

where

condition (2) applies to some known condi­tion of velocity and power required for (L/D)^ at some original, basic altitude condition (2) applies to some new values of velocity and power required for (L/D)^ at some different altitude and

V = velocity, knots ( TAS, of course)

Pr=power required, h. p. tr=altitude density ratio (sigma)

Thus, if flight is conducted at 22,000 ft. (<r=0.498), the airplane will have: a 42 percent higher velocity a 42 percent higher power required

than when operating at sea level. Of course, the greater velocity is a higher TAS since the airplane at a given weight and lift coefficient will require the same EAS independent of altitude. Also, the drag of the airplane at altitude is the same as the drag at sea level but the higher TAS causes a proportionately greater power required. Note that the same straight line from the origin tangent to the sea level power curve also is tangent to the altitude power curve.

The effect of altitude on specific range can be appreciated from the previous relationships. If a change in altitude causes identical changes in velocity and power required, the proportion of velocity to power required would be un­changed. This fact implies that the specific range of the propeller powered airplane would be unaffected by altitude. In the actual case, this is true to the extent that powerplant specif­ic fuel consumption (r) and propeller efficiency (fjp) are the principal factors which could cause a variation of specific range with altitude.

If compressibility effects are negligible, any variation of specific range with altitude is strictly a function of engine-propeller performance.

The airplane equipped with the reciprocating engine will experience very little, if any, variation of specific range with altitude at low altitudes. There is negligible variation of brake specific fuel consumption for values of ВНР below the maximum cruise power rating of the powerplant which is the auto-lean or manual lean range of engine operation. Thus, an increase in altitude will produce a decrease in specific range only when the increased power requirement exceeds the maximum cruise power rating of the powerplants. One advantage of supercharging is that the cruise power may be maintained at high altitude and the airplane may achieve the range at high altitude with the corresponding increase in TAS. The prin­cipal differences in the high altitude cruise and low altitude cruise are the true airspeeds and climb fuel requirements.

The airplane equipped with, the turboprop powerplant will exhibit a variation of specific range with altitude for two reasons. First, the specific fuel consumption (/) of the turbine engine improves with the lower inlet tem­peratures common to high altitudes. Also, the low power requirements to achieve opti­mum aerodynamic conditions at low altitude necessitate engine operation at low, inefficient output power. The increased power require­ments at high altitudes allow the turbine powerplant to operate in an efficient output range. Thus, while the airplane has no particular preference for altitude, the power – plants prefer the higher altitudes and cause an increase in specific range with altitude. Generally, the upper limit of altitude for efficient cruise operation is defined by airplane gross weight (and power required) or com­pressibility effects.

The optimum climb and descent for the propeller powered airplane is affected by many different factors and no general, all­inclusive relationship is applicable. Hand­book data for the specific airplane and various

operational factors will define operating pro­cedures.

RANGE, TURBOJET AIRPLANES. Many different factors influence the range of the turbojet airplane. In order to simplify the analysis of the overall range problem, it is convenient to separate airplane factors from powerplant factors and analyze each item independently. An analogy would be the study of “horsecart” performance by separat­ing “cart” performance from “horse” per­formance to distinguish the principal factors which affect the overall performance.

In the case of the turbojet airplane, the fuel flow is determined mainly by the thrust rather than power. Thus, the fuel flow could be most directly related to the thrust required to maintain the airplane in steady, level flight. This fact allows study of the turbojet powered airplane by analysis of the curves of thrust required versus velocity. Figure 2.25 illu­strates a typical curve of thrust required versus velocity which would be (somewhat) analo­gous to the variation of fuel flow versus veloc­ity. Maximum endurance condition would be obtained at (L/D)^ since this would incur the lowest fuel flow to keep the airplane in stead у, level flight. Maximum range condition would occur where the proportion between velocity and thrust required is greatest and this point is located by a straight line from the origin tangent to the curve.

The maximum range is obtained at the aero­dynamic condition which produces a maximum proportion between the square root of the lift coefficient^ (CL) and the drag coefficient (Cd), or_C^CL/CD’)max. In subsonic perform­ance, occurs at a particular value

angle of attack and lift coefficient and is un­affected by weight or altitude (within com­pressibility limits). At this specific aerody­namic condition, induced drag is approxi­mately 25 percent of the total drag so the turbojet airplane designed for long range does not have the strong preference for high aspect ratio planform like the propeller airplane.

On the other hand, since approximately 75 percent of the total drag is parasite drag, the turbojet airplane designed specifically for long range has the special requirement for great aerodynamic cleanness.

The effect of the variation of airplane gross weight is illustrated by the second graph of figure 2.25. The flight condition of (jIqjcbx« is achieved at one value of lift coefficient for a given airplane in subsonic flight. Hence, a variation of gross weight will alter the values of airspeed, thrust required, and specific range obtained at (_^СьіС^)тят. If a given configuration is operated at constant altitude and lift coefficient the following re­lationships will apply:

Thus, a 10 percent increase in gross weight would create:

a 5 percent increase in velocity a 10 percent increase in thrust required a 5 percent decrease in specific range when flight is maintained at the optimum con­ditions of (VC/Cd)™^ Since most jet airplanes

have a fuel weight which is a large part of the gross weight, cruise control procedures will be necessary to account for the changes in opti­mum airspeeds and power settings as fuel is consumed.

The effect of altitude on the range of the turbojet airplane is of great importance be­cause no other single item can cause such large variations of specific range. If a given con­figuration of airplane is operated at constant gross weight and the lift coefficient for a change in altitude will produce the following relationships:

Tr= constant (neglecting compressibility effects)

SR-i_ fir, (neglecting factors affecting en – TjRi у с2 gine performance)

where

condition (i) applies some known condition of velocity, thrust required, and specific range for (■}JCL! CD’)ma at some original, basic altitude.

condition (2) applies to some new values of velocity, thrust required, and specific range for (VcyCzOUz at some different altitude.

and

V— velocity, knots (TAS, of course)

Tr= thrust required, lbs.

SR= specific range, nmi/lb.

<r=altitude density ratio (sigma)

Thus, if flight is conducted at 40,000 ft. (<r = 0.246), the airplane will have: a 102 percent higher velocity the same thrust required a 102 percent higher specific range (even when the beneficial effects of altitude on engine performance are neglected)

than when operating at sea level. Of course, the greater velocity is a higher TAS and the same thrust required must be obtained with a greater engine RPM.

At this point it is necessary to consider the effect of the operating condition on pojverplant performance. An increase in altitude will im­prove powerplant performance in two respects. First, an increase in altitude when below the tropopause will provide lower inlet air tem­peratures which reduce the specific fuel con­sumption (c(). Of course, above the tropo­pause the specific fuel consumption tends to increase. At low altitude, the engine RPM necessary to produce the required thrust is low and, generally, well below the normal rated value. Thus, a second benefit of altitude on engine performance is due to the increased RPM required to furnish cruise thrust. An increase in engine speed to the normal rated value will reduce the specific fuel consumption.

The increase in specific range with altitude of the turbojet airplane can be attributed to these three factors:

(1) An increase in altitude will increase the proportion of (V/Tr) and provide a greater TAS for the same Tr.

(2) An increase in altitude in the tropo­sphere will produce lower inlet air temperature which reduces the specific fuel consumption.

(3) An increase in altitude requires in­creased engine RPM. to provide cruise thrust and the specific fuel consumption reduces as normal rated RPM is approached.

The combined effect of these three factors de­fines altitude as the one most important item affecting the specific range of the turbojet air­plane. As an example of this combined’effect, the typical turbojet airplane obtains a specific range at 40,000 ft. which is approximately 130 percent greater than that obtained at sea level. The increased TAS accounts for approxi­mately two-thirds of this benefit while in­creased engine performance (reduced q) ac­counts for the other one-third of the benefit. For example, at sea level the maximum spe­cific range of a turbojet airplane may be 0.1 nmi/lb. but at 40,000 ft. the maximum specific range would be approximately 0.25 nmi/lb.

From the previous analysis, it is apparent that the cruise altitude of the turbojet should be as high as possible within compressibility or thrust limits. Generally, the optimum alti­tude to begin cruise is the highest altitude at which the maximum continuous thrust can provide the optimum aerodynamic conditions. Of course, the optimum altitude is determined mainly by the gross weight at the begin of cruise. For the majority of turbojet airplanes this altitude will be at or above the tropopause for normal cruise configurations.

Most turbojet airplanes which have tran­sonic or moderate supersonic performance will obtain maximum range with a high subsonic cruise. However, the airplane designed spe­cifically for high supersonic performance will obtain maximum range with a supersonic cruise and subsonic operation will cause low lift-drag ratios, poor inlet and engine perform­ance and reduce the range capability.

The cruise control of the turbojet airplane is considerably different from that of the pro­peller driven airplane. Since the specific range is so greatly affected by altitude, the optimum altitude for begin of cruise should be attained as rapidly as is consistent with climb fuel re­quirements. The range-climb program varies considerably between airplanes and the per­formance section of the flight handbook will specify the appropriate procedure. The de­scent from cruise altitude will employ essen­tially the same feature, a rapid descent is necessary to minimize the time at low altitudes where specific range is low and fuel flow is high for a given engine speed.

During cruise flight of the turbojet airplane, the decrease of gross weight from expenditure of fuel can result in two types of cruise control. During a constant altitude cruise, a reduction in gross weight will require a reduction of air­speed and engine thrust to maintain the opti­mum lift coefficient of subsonic cruise. While such a cruise may be necessary to conform to the flow of traffic, it constitutes a certain in­efficiency of operation. If the airplane were not restrained to a particular altitude, main­taining the same lift coefficient and engine speed would allow the airplane to climb as the gross weight decreases. Since altitude gen­erally produces a beneficial effect on range, the climbing cruise implies a more efficient flight path.

The cruising flight of the turbojet airplane will begin usually at or above the tropopause in order to provide optimum range conditions. If flight is conducted at QjcL/CD’)max, optimum range will be obtained at specific values of lift coefficient and drag coefficient. When the air­plane is fixed at these values of CL and CD and the TAS is held constant, both lift and drag are directly proportional to the density ratio, a. Also, above the tropopause, the thrust is pro­portional to a when the TAS and RPM are con­stant. As a result, a reduction of gross weight by the expenditure of fuel would allow the airplane to climb but the airplane would re­main in equilibrium because lift, drag, and thrust all vary in the same fashion. This re­lationship is illustrated by figure 2.26.

The relationship of lift, drag, and thrust is convenient for, in part, it justifies the condi­tion of a constant velocity. Above the tropo­pause, the speed of sound is constant hence a constant velocity during the cruise-climb would produce a constant Mach number. In this case, the optimum values of Q^CJC^), CL and CD do not vary during the climb since the Mach number is constant. The specific fuel consumption is initially constant above the tropopause but begins to increase at altitudes much above the tropopause. If the specific fuel consumption is assumed to be constant during the cruise-climb, the following rela­tionships will apply:

V, M, CL and CD arc constant <H_W2 a, Wx

FF2_ <r2

FFi <гі

SRi_JWi (cruise climb above tropopause,

SR і W2 constant M, ci)

where

condition (X) applies to some known condi­tion of weight, fuel flow, and specific range at some original basic altitude during cruise climb.

condition (2) applies to some new values of weight, fuel flow, and specific range at some different altitude along a partic­ular cruise path.

and

V= velocity, knots

M=Mach number

W= gross weight, lbs.

RF=fuel flow, lbs./hr.

SR— specific range, nmi./lb.

<r= altitude density ratio

Thus, during a cruise-climb flight, a 10 percent decrease in gross weight from the consumption of fuel would create:

no change in Mach number or TAS a 5 percent decrease in EAS a 10 percent decrease in <r, i. e., higher altitude

a 10 percent decrease in fuel flow an 11 percent increase in specific range

An important comparison can be made between the constant altitude cruise and the cruise – climb with respect to the variation of specific range. From the previous relationships, a 2 percent reduction in gross weight during

SR,= IWi SRi W2

SR2_ Wi.

SRi W2

cruise would create a 1 percent increase in specific range in a constant altitude cruise but a 2 percent increase in specific range in a cruise – climb at constant Mach number. Thus, a higher average specific range cari. be maintained during the expenditure of a given increment of fuel. If an airplane begins a cruise at optimum conditions at or above the tropopause with a given weight of fuel, the following data

provide a comparison of the total range avail­able from a constant altitude or cruise-climb

flight path.

Hath of cruise fuel weight to airplane grots weight at beginning of cruise

0. 0

.1

.2

• 3 .4

• 5 .6 .7

For example, if the cruise fuel weight is 50 per­cent of the gross weight, the climbing cruise flight path will provide a range 18.2 percent greater than cruise at constant altitude. This comparison does not include consideration of any variation of specific fuel consumption dur­ing cruise or the effects of compressibility in defining the optimum aerodynamic conditions for cruising flight. However, the comparison is generally applicable for aircraft which have Subsonic cruise.

When the airplane has a supersonic cruise for maximum range, the optimum flight path is generally one of a constant Mach number. The optimum flight path is generally—but not necessarily—a climbing cruise. In this case of subsonic or supersonic cruise, a Machmeter is of principal importance in cruise control of the jet airplane.

The effect of wind on range is of considerable importance in flying operations. Of course, a headwind will always reduce range and a tailwind will always increase range. The selection of a cruise altitude with the most favorable (or least unfavorable) winds is a rel­atively simple matter for the case of the propeller powered airplane. Since the range of the. propeller powered airplane is relatively un­affected by altitude, the altitude with the most favorable winds is selected for range. However, the range of the turbojet airplane is greatly affected by altitude so the selection of an op­timum altitude will involve considering the wind profile with the variation of range with altitude. Since the turbojet range increases

TURBOJET CRUISE-CLIMB

greatly with altitude, the turbojet can tolerate less favorable (or more unfavorable) winds with increased altitude.

In some cases, large values of wind may cause a significant change in cruise velocity to maintain maximum ground nautical miles per lb. of fuel. As an example of an extreme con­dition, consider an airplane flying into a head­wind which equals the cruise velocity. In this case, any increase in velocity would improve range.

To appreciate the changes in optimum speeds with various winds, refer to the illustration of figure 2.26. When zero wind conditions exist, a straight line from the origin tangent to the curve of fuel flow versus velocity will locate maximum range conditions. When a head­wind condition exists, the speed for maximum ground range is located by a line tangent drawn from a velocity offset equal to the headwind velocity. This will locate maximum range at some higher velocity and fuel flow. Of course, the range will be less than when at zero wind conditions but the higher velocity and fuel flow will minimize the range loss due to the head­wind. In a similar sense, a tailwind will re­duce the cruise velocity to maximize the benefit of the tailwind.

The procedure of employing different cruise velocities to account for the effects of wind is necessary only at extreme values of wind velocity. It is necessary to consider the change in optimum cruise airspeed when the wind velocities exceed 25 percent of the zero wind cruise velocity.

CLIMB performance

During climbing flight, the airplane gains potential energy by virtue of elevation. This increase in potential energy during a climb is provided by one, or a combination, of two means-, (1) expenditure of propulsive energy above that required to maintain level flight or

(2) expenditure of airplane kinetic energy, i. e., loss of velocity by a zoom. Zooming for alti­tude is a transient process of trading kinetic energy for potential energy and is of considera­ble importance for airplane configurations which can operate at very high levels of kinetic energy. However, the major portions of climb performance for most airplanes is a near steady process in which additional propulsive energy is converted into potential energy. The funda­mental parts of airplane climb performance in­volve a flight condition where the airplane is in equilibrium but not at constant altitude.

The forces acting on the airplane during a climb are shown by the illustration of figure

2.21. When the airplane is in steady flight with moderate angle of climb, the vertical component of lift is very nearly the same as the actual lift. Such climbing flight would exist with the lift very nearly equal to the weight. The net thrust of the powerplant may be in­clined relative to the flight path but this effect will be neglected for the sake of simplicity. Note that the weight of the aircraft is vertical but a component of weight will act aft along the flight path.

If it is assumed that the aircraft is in a steady climb with essentially small inclination of the flight path, the summation of forces along the flight path resolves to the following:

Forces forward=Forces aft

T__ n і tj/ _: –

x УУ ЫН У

where

T= thrust available, lbs.

D=drag, lbs.

W= weight, lbs.

у = flight path inclination or angle of climb, degrees (“gamma’

This basic relationship neglects some of the factors which may be of importance for air­planes of very high climb performance. For example, a more detailed consideration would account for the inclination of thrust from the flight path, lift not equal to weight, subse­quent change of induced drag, etc. However, this basic relationship will define the principal factors affecting climb performance. With this relationship established by the condition of equilibrium, the following relationship exists to express the trigonometric sine of the climb angle, y:

This relationship simply states that, for a given weight airplane, the angle of climb (7) depends on the difference between thrust and drag (T— D), or excess thrust. Of course, when the excess thrust is zero (T— D=0 or Т-D’), the inclination of the flight path is zero and the airplane is in steady, level flight. When the thrust is greater than the drag, the excess thrust will allow a climb angle depend­ing on the value of excess thrust. Also, when the thrust is less than the drag, the deficiency of thrust will allow an angle of descent.

The most immediate interest in the climb angle performance involves obstacle clearance. The maximum angle of climb would occur where there exists the greatest difference be­tween thrust available and thrust required, i. e., maximum (T—Z>). Figure 2.21 illustrates the climb angle performance with the curves of thrust available and thrust required versus velocity. The thrust required, or drag, curve is assumed to be representative of some typical airplane configuration which could be powered by either a turbojet or propeller type power – plant. The thrust available curves included are for a characteristic propeller powerplant and jet powerplant operating at maximum output.

The thrust curves for the representative pro­peller aircraft show the typical propeller thrust which is high at low velocities and decreases with an increase in velocity. For the pro­peller powered airplane, the maximum excess thrust and angle of climb will occur at some speed just above the stall speed. Thus, if it is necessary to clear an obstacle after takeoff, the propeller powered airplane will attain maximum angle of climb at an airspeed con­veniently close to—if not at—the takeoff speed.

The thrust curves for the representative jet aircraft show the typical turbojet thrust which is very nearly constant with speed. If the thrust available is essentially constant with speed, the maximum excess thrust and angle of climb will occur where the thrust required

and*

e*8)

near the speed for (Z,/jD)met. There is no direct relationship which establishes this situation since the variation of propeller efficiency is the principal factor accounting for the variation of power available with velocity. In an ideal sense, if the propeller efficiency were constant, maximum rate of climb would occur at the speed for minimum power required. How­ever, in the actual case, the propeller efficiency of the ordinary airplane will produce lower power available at low velocity and cause the maximum rate of climb to occur at a speed greater than that for minimum power required.

The power curves for the representative jet aircraft show the near linear variation of power available with velocity. The maximum rate of climb for the typical jet airplane will occur at some speed much higher than that for max­imum rate of climb of the equivalent propeller powered airplane. In part, this is accounted for by the continued increase in power – avail­able with speed. Note that a 50 percent in­increase in thrust by use of an afterburner may cause an increase in rate of climb of approxi­mately 100 percent.

The climb performance of an airplane is affected by many various factors. The con­ditions of maximum climb angle or climb rate occur at specific speeds and variations in spe’ed will produce variations in climb performance. Generally, there is sufficient latitude that small variations in speed from the optimum do not produce large changes in climb performance and certain operational items may require speeds slightly different from the optimum. Of course, climb performance would be most critical at high weight, high altitude, or dur­ing malfunction of a powerplant. Then, opti­mum climb speeds are necessary. A change in airplane weight produces a twofold effect on climb performance. First, the weight, W, appears directly in denominator of the equa­tions for both climb angle and climb rate. In addition, a change in weight will alter the drag and power required. Generally, an in­crease in weight will reduce the maximum rate

of climb but the airplane must be operated at some increase of speed to achieve the smaller peak climb rate (unless the airplane is compres­sibility limited).

The effect of altitude on climb performance is illustrated by the composite graphs of figure

2.22. Generally, an increase in altitude will increase the power required and decrease the power available. Hence, the climb perform­ance of an airplane is expected to be greatly affected by altitude. The composite chart of climb performance depicts the variation with altitude of the speeds for maximum rate of climb, maximum angle of climb, and maximum and minimum level flight airspeeds. As alti­tude is increased, these various speeds finally converge at the absolute ceiling of the airplane. At the absolute ceiling, there is no excess of power or thrust and only one speed will allow steady level flight. The variation of rate of climb and maximum level flight speed with altitude for the typical propeller powered air­plane give evidence of the effect of supercharg­ing. Distinct aberrations in these curves take place at the supercharger critical altitudes and blower shift points. The curve of time to climb is the result of summing up the incre­ments of time spent climbing through incre­ments of altitude. Note that approach to the absolute ceiling produces tremendous increase in the time curve.

Specific reference points are established by these composite curves of climb performance. Of course, the absolute ceiling of the airplane produces zero rate of climb. The service ceiling is specified as the altitude which produces a rate of climb of 100 fpm. The altitude which produces a rate of climb of 500 fpm is termed the combat ceiling. Usually, these specific refer­ence points are provided for the airplane at the combat configuration or a specific design configuration.

The composite curves of climb performance for the typical turbojet airplane are shown in figure 2.22. One particular point to note is the more rapid decay of climb performance

ABSOLUTEJ CEILING

with altitude above the tropopause. This is due in great part to the more rapid decay of engine thrust in the stratosphere.

During a power off descent the deficiency of thrust and power define the angle of descent and rate of descent. Two particular points are of interest during a power off descent: minimum angle of descent and minimum rate of descent, The minimum angle of descent would provide maximum glide distance through the air. Since no thrust is available from the power plant, minimum angle of descent would be obtained at (LjD’)vua. At (L/D’)max the deficiency of thrust is a minimum and, as shown by figure 2.22, the greatest proportion between velocity and power required is ob­tained. The minimum rate of descent in power off flight is obtained at the angle of attack and airspeed which produce minimum power required. For airplanes of moderate aspect ratio, the speed for minimum rate of descent is approximately 75 percent of the speed for minimum angle of descent

ITEMS OF AIRPLANE. PERFORMANCE

The various items of airplane performance result from the combination of airplane and powerplant characteristics. The aerodynamic characteristics of the airplane generally define the power and thrust requirements at various conditions of flight while the powerplant characteristics generally define the power and thrust available at various conditions of flight. The matching of the aerodynamic configura­tion with the powerplant will be accomplished to provide maximum performance at the speci­fic design condition, e. g., range, endurance, climb, etc.

STRAIGHT AND LEVEL FLIGHT

When the airplane is in steady, level flight, the condition of equilibrium must prevail. The unaccelerated condition of flight, is achieved with the airplane trimmed for lift equal to weight and the powerplant set for a thrust to equal the airplane drag. In certain conditions of airplane performance it is con­venient to consider the airplane requirements by the thrust required (or drag) while in other cases it is more applicable to consider the power required. Generally, the jet airplane will require consideration of the thrust required and the propeller airplane will require consid­eration of the power required. Hence, the airplane in steady level flight will require lift equal to weight and thrust available equal to thrust required (drag) or power available equal to power required.

The variation of power required and thrust required with velocity is illustrated in figure

2.20. Each specific curve of power or thrust required is valid for a particular aerodynamic configuration at a given weight and altitude. These curves define the power or thrust re­quired to achieve equilibrium, lift-equal – weight, constant altitude flight at various airspeeds. As shown by the curves of figure

2.20, if it is desired to operate the airplane at the airspeed corresponding to point A, the power or thrust required curves define a par­ticular value of thrust or power that must be made available from the powerplant to achieve equilibrium. Some different airspeed such as that corresponding to point В changes the value of thrust or power required to achieve equilibrium. Of course, the change of air­speed to point В also would require a change in angle of attack to maintain a constant lift equal to the airplane weight. Similarly, to establish airspeed and achieve equilibrium at point C will require a particular angle of attack and powerplant thrust or power. In this case, flight at point C would be in the vicinity of the minimum flying speed and a major portion of the thrust or power required would be due to induced drag.

The maximum level flight speed for the air­plane will be obtained when the power .or thrust required equals the maximum power or thrust available from the powerplant. The minimum level flight airspeed is not usually defined by thrust or power requirement since conditions of stall or stability and control problems generally predominate.

AIRCRAFT PROPELLERS

The aircraft propeller functions to convert the powerplant shaft horsepower into propul­sive horsepower. The basic principles of pro­pulsion apply to the propeller in that thrust is produced by providing the airstream a mo­mentum change. The propeller achieves high propulsive efficiency by processing a relatively large mass flow of air and imparting a rela­tively small velocity change. The momentum change created by propeller is shown by the illustration of figure 2.18.

The action of the propeller can be idealized by the assumption that the rotating propeller is simply an actuating disc. As shown in fig­ure 2,18, the inflow approaching the propeller disc indicates converging streamlines with an increase in velocity and drop in pressure. The converging streamlines leaving the propeller disc indicate a drop in pressure and increase in velocity behind the propeller. The pressure change through the disc results from the distri­bution of thrust over the area of the propeller disc. In this idealized propeller disc, the pres­sure difference is uniformly distributed over the disc area but the actual case is rather different from this.

The final velocity of the propeller slipstream, Vt, is achieved some distance behind the pro­peller. Because of the nature of the flow pat­tern produced by the propeller, one half of the total velocity change is produced as the flow reaches the propeller disc. If the complete velocity increase amounts to la, the flow veloc­ity has increased by the amount a at the pro­peller disc. The propulsive efficiency, yp, of the ideal propeller could be expressed by the fol­lowing relationship:

Since the final velocity, V2, is the sum of total velocity change la and the initial velocity, Vl} the propulsive efficiency rearranges to a form identical to that for the turbojet.

2

So, the same relationship exists as with the turbojet engine in that high efficiency is de­veloped by producing thrust with th^ highest possible mass flow and smallest necessary velocity change.

The actual propeller must be evaluated in a more exact sense to appreciate the effect of nonuniform disc loading, propeller blade drag forces, interference flow between blades, etc. With these differences from the ideal propeller,

it is more appropriate to define propeller effi­ciency in the following manner:

r

Many different factors govern the efficiency of a propeller. Generally, a large diameter pro­peller favors a high propeller efficiency from the standpoint of large mass flow. However, a powerful adverse effect on propeller efficiency is produced by high tip speeds and compressi­bility effects. Of course, small diameter pro­pellers favor low tip speeds. In addition, the propeller and powerplant must be matched for compatibility of output and efficiency.

In order to appreciate some of the principal factors controlling the efficiency of a given propeller, figure 2.18 illustrates the distribu­tion of rotative velocity along the rotating propeller blade. These rotative velocities add to the local inflow velocities to produce a variation of resultant velocity and direction along the blade. The typical distribution of thrust along the propeller blade is shown with the predominating thrust being located on the outer portions of the blade. Note that the propeller producing thrust develops a tip vortex similar to the wing producing lift. Evidence of this vortex can be seen by the con­densation phenomenon occurring at this loca­tion under certain atmospheric conditions.

The component velocities at a given propeller blade section are shown by the diagram of figure 2.18. The inflow velocity adds vec­torially to the velocity due to rotation to pro­duce an inclination of the resultant wind with respect to the plane of rotation. This incli­nation is termed ф (phi), the effective pitch angle, and is a function of some proportion of the flight velocity, V, and the velocity due to rotation which is vnD at the tip. The pro­portions of these terms describe the propeller “advance ratio”, J.

where

J=propeller advance ratio V— flight velocity, ft. per sec.

»=propeller rotative speed, revolutions per sec.

D=propeller diameter, ft.

The propeller blade angle, 0 (beta), varies throughout the length of the blade but a representative value is measured at 75 percent of the blade length from the hub.

Note that the difference between the effec­tive pitch angle, Ф, and the blade angle, 0, determines an effective angle of attack for the propeller blade section. Since the angle of attack is the principal factor affecting the efficiency of an airfoil section, it is reasonable to make the analogy that the advance ratio, J, and blade angle, 0, are the principal factors affecting propeller efficiency. The perform­ance of a propeller is typified by the chart of figure 2.19 which illustrates the variation of propeller efficiency, i}p, with advance ratio, J, for various values of blade angle, 0. The value of rtp for each 0 increases with J until a peak is reached, then decreases. It is apparent that a fixed pitch propeller may be selected to provide suitable performance in a narrow range of advance ratio but efficiency would suffer considerably outside this range.

In order to provide high propeller efficiency through a wide range of operation, the pro­peller blade angle must be controllable. The most convenient means of controlling the propeller is the provision of a constant speed governing apparatus. The constant speed gov­erning feature is favorable from the standpoint of engine operation in that engine output and efficiency is positively controlled and governed.

The governing of the engine-propeller combi­nation will allow operation throughout a wide range of power and speed while maintaining efficient operation.

If the envelope of maximum propeller effi­ciency is available, the propulsive horsepower available will appear as shown in the second chart of figure 2.19- The propulsive power available, Pa, is the product of the propeller efficiency and applied shaft horsepower.

(ВНР)

The propellers used on most large reciprocating engines derive peak propeller efficiencies on the order of i7p=0.85 to 0.88. Of course, the peak values are designed to occur at some specific design condition. For example, the selection of a propeller for a long range transport would require matching of the engine-propeller com­bination for peak efficiency at cruise condition. On the other hand, selection of a propeller for a utility or liaison type airplane would require matching of the engine-propeller combination to achieve high propulsive power at low speed and high power for good takeoff and climb performance.

Several special considerations must be made for the application of aircraft propellers. In the event of a powerplant malfunction or failure, provision must be made to streamline the propeller blades and reduce drag so that flight may be continued on the remaining op­erating engines. This is accomplished by feathering the propeller blades which stops rotation and incurs a minimum of drag for the inoperative engine. The necessity for feather­ing is illustrated in figure 2.19 by the change in equivalent parasite area, Д/, with propeller blade angle, /3, of a typical installation. When the propeller blade angle is in the feathered position, the change in parasite drag is at a minimum and, in the case of a typical multi­engine aircraft, the added parasite drag from a single feathered propeller is a relatively small contribution to the airplane total drag.

At smaller blade angles near the flat pitch position, the drag added by the propeller is very large. At these small blade angles, the propeller windmilling at high RPM can create such a tremendous amount of drag that the airplane may be uncontrollable. The propel­ler windmilling at high speed in the low range of blade angles can produce an increase in para­site drag which may be as great as the parasite drag of the basic airplane. An indication of this powerful drag is seen by the helicopter in autorotation. The windmilling rotor is ca­pable of producing autorotation rates of descent which approach that of a parachute canopy with the identical disc area loading. Thus, the propeller windmilling at high speed and small blade angle can produce an effective dtag coefficient of the disc area which compares with that of a parachute canopy. The drag and yawing moment caused by loss of power at high engine-propeller speed is considerable aftd the transient yawing displacement of the aircraft may produce critical loads for the vertical tail. For this reason, automatic feathering may be a necessity rather than a luxury.

The large drag which can be produced by the rotating propeller can be utilized to im­prove the stopping performance of the air­plane. Rotation of the propeller blade to small positive values or negative values with applied power can produce large drag or re­verse thrust. Since the thrust capability of the propeller is quite high at low speeds, very high deceleration can be provided by reverse thrust alone.

The oprating limitations of the propeller are closely associated with those of the power- plant. Overspeed conditions are critical be­cause of the large centrifugal loads and blade twisting moments produced by an excessive rotative speed. In addition, the propeller blades will have various vibratory inodes and certain operating limitations may be necessary to prevent exciting resonant conditions.

PROPELLER EFFICIENCY

Figure 2.19. Propeller Operation

THE RECRIPROCATING ENGINE

The reciprocating engine is one of the most efficient powerplants used for aircraft power. The combination of the reciprocating engine and propeller is one of the most efficient means of converting the chemical energy of fuel into flying time or distance. Because of the in­herent high efficiency, the reciprocating engine is an important type of aircraft powerplant.

OPERATING CHARACTERISTICS. The function of the typical reciprocating engine in­volves four strokes of the piston to complete one operating cycle. This principal operating cycle is illustrated in figure 2.15 by the varia­tion of pressure and volume within the cylin­der. The first stroke of the operating cycle is the downstroke of the piston with the intake valve open. This stroke draws in a charge of fuel-air mixture along AB of the pressure – volume diagram. The second stroke accom­plishes compression of the fuel-air mixture along line BC. Combustion is initiated by a spark ignition apparatus and combustion takes place in essentially a constant volume. The combustion of the fuel-air mixture liberates heat and causes the rise of pressure along line CD. The power stroke utilizes the increased pressure through the expansion along line DE. Then the exhaust begins by the initial rejection along line EB and is completed by the upstroke along line BA.

The net work produced by the cycle of opera­tion is idealized by the area BCDE on the pressure-volume diagram of figure 2.15. Dur­ing the actual rather than ideal cycle of op­eration, the intake pressure is lower than the exhaust pressure and the negative work repre­sents a pumping loss. The incomplete expan­sion during the power stroke represents a basic loss in the operating cycle because of the re­jection of combustion products along line EB. The area EFB represents a basic loss in the operating cycle because of the rejection of combustion products along line EB. The area EFB represents a certain amount of energy of the exhaust gases, a part of which can be ex­tracted by exhaust turbines as additional shaft power to be coupled to the crankshaft (turbo­compound engine) or to be used in operating a supercharger (turbosupercharger). In addi­tion, the exhaust gas energy may be utilized to augment engine cooling flow (ejector exhaust) and reduce cowl drag.

Since the net work produced during the op­erating cycle is represented by the enclosed area of pressure-volume diagram, the output of the engine is affected by any factor which influences this area. The weight of fuel-air mixture will determine the energy released by combustion and the weight of charge can be altered by altitude, supercharging, etc. Mixture strength, preignition, spark timing, etc., can affect the energy release of a given airflow and alter the work produced during the operating cycle.

The mechanical work accomplished during the power stroke is the result of the gas pres­sure sustained on the piston. The linkage of the piston to a crankshaft by the connecting rod applies torque to the output shaft. During this conversion of pressure energy to mechani­cal energy, certain losses are inevitable because

COMPRESSION COMBUSTION

of friction and the mechanical output is less than the available pressure energy. The power output from the engine will be determined by the magnitude and rate of the power impulses. In order to determine the power output of the reciprocating engine, a brake or load device is attached to the output shaft and the operating characteristics are determined. Hence, the term “brake” horsepower, ВНР, is used to denote the output power of the powerplant, From the physical definition of "power" and the particular unit of “horsepower” (1 h. p.=

33,0 ft.-lbs. per min.), the brake horsepower can be expressed in the following form.

In this relationship, the output power is ap­preciated as some direct variable of torque, Tf and RPM. Of course, the output torque is some function of the combustion gas pressure during the power stroke. Thus, it is helpful to consider the mean effective gas pressure during the power stroke, the “brake mean effective pressure” or ВМЕР. With use of this term, the ВНР can be expressed in the following form.

СВМЕР)(РХЮ

792,000

where

ВНР = brake horsepower BMEP=brake mean effective pressure, psi D=engine displacement, cu. in.

N= engine speed, RPM

The ВМЕР is not actual pressure within the cylinder, but an effective pressure representing the mean gas load acting on the piston during
the power stroke. As such, ВМЕР is a con­venient index for a majority of items of recip­rocating engine output, efficiency, and operat­ing limitations.

The actual power output of any reciprocat­ing engine is a direct function of the combina­tion of engine torque and rotative speed. Thus, output brake horsepower can be related by the combination of ВМЕР and PPM or torque pressure and RPM. No other engine instruments can provide this immediate indi­cation of output power.

If all other factors are constant, the engine power output is directly related to the engine airflow. Evidence of this fact could be appre­ciated from the equation for ВНР in terms of ВМЕР.

This equation relates that, for a given ВМЕР, the ВНР is determined by the product of en­gine RPM, N, and displacement, D. In a sense, the reciprocating engine could be con­sidered primarily as an air pump with the pump capacity directly affecting the power output. Thus, any engine instruments which relate factors affecting airflow can provide some indirect reflection of engine power. The pres­sure and temperature of the fuel-air mixture decide the density of the mixture entering the cylinder. The carburetor air temperature will provide the temperature of the inlet air at the carburetor. While this carburetor inlet air is not the same temperature as the air in the cylinder inlet manifold, the carburetor inlet temperature provides a stable indication inde­pendent of fuel flow and can be used as a stand­ard of performance. Cylinder inlet manifold temperature is difficult to determine with the same degree of accuracy because of the normal variation of fuel-air mixture strength. The inlet manifold pressure provides an additional indication of the density of airflow entering the combustion chamber. The manifold absolute pressure, MAP, is affected by the carburetor

inlet pressure, throttle position, and super­charger or impeller pressure ratio. Of course, the throttle is the principal control of mani­fold pressure and the throttling action controls the pressure of the fuel-air mixture delivered to the supercharger inlet. The pressure re­ceived by the supercharger is magnified by the supercharger in some proportion depend­ing on impeller speed. Then the high pressure mixture is delivered to the manifold.

Of course, the engine airflow is a function of RPM for two reasons. A higher engine speed increases the pumping rate and the volume flow through the engine. Also, with the engine driven supercharger or impeller, an increase in engine speed increases the supercharger pres­sure ratio. With the exception of near closed throttle position, an increase in engine speed will produce an increase in manifold pressure.

The many variables affecting the character

w* HiV wv/iUMUJCXVfu £JL AIL All LA11L.

subject of reciprocating engine operation. Uniform mixtures of fuel and air will support combustion between fuel-air ratios of approxi­mately 0.04 and 0.20. The chemically correct proportions of air and hydrocarbon fuel would be 15 lbs. of air for each lb. of fuel, or a fuel – air ratio of 0.067. This chemically correct, or “stoichiometric,” fuel-air ratio would provide the proportions of fuel and air to produce maximum release of heat during combustion of a given weight of mixture. If the fuel-air ratio were leaner than stoichiometric, the ex­cess of air and deficiency of fuel would produce lower combustion temperatures and reduced heat release for a given weight of charge. If the fuel-air ratio were richer than stoichio­metric, the excess of fuel and deficiency of air would produce lower combustion temperatures and reduced heat release for a given weight of charge.

The stoichiometric conditions would pro­duce maximum heat release for ideal conditions of combustion and may apply quite closely for the individual cylinders of the low speed re­ciprocating engine. Because of the effects of flame propagation speed, fuel distribution, temperature variation, etc., the maximum power obtained with a fixed airflow occurs at fuel-air ratios of approximately 0.07 to 0.08. The first graph of figure 2.16 shows the varia­tion of output power with fuel-air ratio for a a constant engine airflow, i. e., constant RPM, MAP, and CAT (carburetor air temperature). Combustion can be supported by fuel-air ratios just greater than 0.04 but the energy released is insufficient to overcome pumping losses and engine mechanical friction. Essentially, the same result is obtained for the rich fuel-air ratios just below 0.20. Fuel-air ratios be­tween these limits produce varying amounts of output power and the maximum power output generally occurs at fuel-air ratios of approxi­mately 0.07 to 0.08. Thus, this range of fuel – air ratios which produces maximum power for a given airflow is termed the “best power” range. At some lower range of fuel-air ratios, a maximum of power per fuel-air ratio is ob­tained and this the “best economy” range. The best economy range generally occurs be­tween fuel-air ratios of 0,05 and 0.07. When maximum engine power is required for take­off, fuel-air ratios greater than 0.08 are neces­sary to suppress detonation. Hence, fuel-air ratios of 0.09 to 0.11 are typical during this operation.

The pattern of combustion in the cylinder is best illustrated by the second graph of figure 2.16. The normal combustion process begins by spark ignition toward the end of the com­pression stroke. The electric spark provides the beginning of combustion and a flame front is propagated smoothly through the com­pressed mixture. Such normal combustion is shown by the plot of cylinder pressure versus piston travel. Spark ignition begins a smooth rise of cylinder pressure to some peak value with subsequent expansion through the power stroke. The variation of pressure with piston travel must be controlled to achieve the great­est net work during the cycle of operation.

ENGINE AIRFLOW, LBS. PER HR.

Figure 2.16. Reciprocating Engine Operation


Obviously, spark ignition timing is an impor­tant factor controlling the initial rise of pres­sure in the combustion chamber. The ignition of the fuel mixture must begin at the proper time to allow flame front propagation and the release of heat to build up peak pressure for the power stroke.

The speed of flame front propagation is a major factor affecting the power output of the reciprocating engine since this factor controls the rate of heat release and rate of pressure rise in the combustion chamber. For this reason, dual ignition is necessary for powerplants of high specific power output. Obviously, nor­mal combustion can be accomplished more rapidly with the propagation of two flame fronts rather than one. The two sources of ignition are able to accomplish the combus­tion heat release and pressure rise in a shorter period of time. Fuel-air ratio is another factor affecting the flame propagation speed in the combustion chamber. The maximum flame propagation speed occurs near a fuel-air ratio of 0.08 and, thus, maximum power output for a given airflow will tend to occur at this value rather than the stoichiometric value.

Two aberrations of the combustion process are preignition and detonation. Preignition is simply a premature ignition and flame front propagation due to hot spots in the combustion chamber. Various lead and carbon deposits and feathered edges on metal surfaces can sup­ply a glow ignition spot and begin a flame propagation prior to normal spark ignition. As shown on the graph of figure 2.16, pre­ignition causes a premature rise of pressure during the piston travel. As a result, preignition combustion pressures and tempera­tures will exceed normal combustion values and are very likely to cause engine damage. Be­cause of the premature rise of pressure toward the end of the compression stroke, the net work of the operating cycle is reduced. Preignition is evidenced by a rise in cylinder head tempera­ture and drop in ВМЕР or torque pressure.

Denotation offers the possibility of immedi­ate destruction of the powerplant. The nor­mal combustion process is initiated by the spark and beginning of flame front propaga­tion. As the flame front is propagated, the combustion chamber pressure and temperature begin to rise. Under certain conditions of high combustion pressure and temperature, the mixture ahead of the advancing flame front may suddenly explode with considerable vi­olence and send strong detonation waves through the combustion chamber. The result is depicted by the graph of figure 2.16, where a sharp, explosive increase in pressure takes place with a subsequent reduction of the mean pres­sure during the power stroke. Detonation produces sharp explosive pressure peaks many times greater than normal combustion. Also, the exploding gases radiate considerable heat and cause excessive temperatures for many local tiarts of the engine. The effects of heavy

1 " " O" /

detonation are so severe that structural damage is the immediate result. Rapid rise of cylinder head temperature, rapid drop in ВМЕР, and loud, expensive noises are evidence of detona­tion.

Detonation is not necessarily confined to a period after the beginning of normal flame front propagation. With extremely low grades of fuel, detonation can occur before normal igni­tion. In addition, the high temperatures and pressure caused by preignition will mean that detonation is usually a corollary of preigniticn. Detonation results from a sudden, unstable de­composition of fuel at some critical combina­tion of high temperature and pressure. Thus, detonation is most likely to occur at any op­erating condition which produces high com­bustion pressures and temperatures. Gener­ally, high engine airflow and fuel-air ratios for maximum heat release will produce the critical conditions. High engine airflow is common to high MAP and RPM and the engine is most sensitive to CAT and fuel-air ratio in this region.

The detonation properties of a fuel are de­termined by the basic molecular structure of the fuel and the various additives. The fuel detonation properties are generally specified by the antidetonation or antiknock qualities of an octane rating. Since the antiknock proper­ties of a high quality fuel may depend on the mixture strength, provision must be made in. the rating of fuels. Thus, a fuel grade of 115/145 would relate a lean mixture antiknock rating of 115 and a rich mixture antiknock rating of 145. One of the most common opera­tional causes of detonation is fuel contamina­tion. An extremely small contamination of high octane fuel with jet fuel can cause a serious decrease in the antiknock rating. Also, the contamination of a high grade fuel with the next lower grade will cause a noticeable loss of antiknock quality.

The fuel metering requirements for an engine are illustrated by the third graph of figure 2.16 which is a plot of fuel-air ratio versus engine airflow. The carburetor must provide specific fuel-air ratios throughout the range of engine airflow to accommodate certain output power. Most modern engines equipped with auto­matic mixture control provide a scheduling of fuel-air ratio for automatic rich or automatic lean operation. The auto-rich scheduling usu­ally provides a fuel-air ratio at or near the maximum heat release value for the middle range of airflows. However, at high airflows a power enrichment must be provided to sup­press detonation. The auto-rich schedule gen­erally will provide an approximate fuel-air ratio of 0.08 which increases to 0.10 or 0.11 at the airflow for takeoff power. In addition, the low airflow and mixture dilution that oc­curs in the idle power range requires enrich­ment for satisfactory operation.

The schedule of fuel-air ratios with an auto­matic lean fuel-air ratio will automatically provide maximum usable economy. If manual leaning procedures are applicable a lower fuel – air ratio may be necessary for maximum possi­ble efficiency. The maximum continuous cruise power is the upper limit of power that can be utilized for this operation. Higher air­flows and higher power without a change in fuel-air ratio will intersect the knee of the detonation envelope.

The primary factor relating the efficiency of operation of the reciprocating engine is the brake specific fuel consumption, BSFC, or simply c.

Brake specific fuel consumption

_ engine fuel flow brake horsepower lbs. per hr.

C_ ВНР

Typical minimum values for c range from 0.4 to 0.6 lbs. per hr. per ВНР and most aircraft powerpiants average 0.5. The turbocompound engine is generally the most efficient because of the power recovery turbines and can ap­proach values of c=0.38 to 0.42, It should be noted that the minimum values of specific fuel consumption will be obtained only within the range of cruise power operation, 30 to 60 per­cent of the maximum power output. Gen­erally, the conditions of minimum specific fuel consumption are achieved with auto-lean or manual lean scheduling of fuel-air ratios and high ВМЕР and low RPM. The low RPM is the usual requirement to minimize friction horsepower and improve output efficiency.

The effect of altitude is to reduce the engine airflow and power output and supercharging is necessary to maintain high power output at high altitude. Since the basic engine is able to process air only by the basic volume displacement, the function of the supercharger is to compress the inlet air and provide a greater weight of air for the engine to process. Of course, shaft power is necessary to operate the engine driven supercharger and a tempera­ture rise occurs through the supercharger com­pression. The effect of various forms of super­charging on altitude performance is illustrated in figure 2.17-

The unsupercharged—or naturally aspi­rated—engine has no means of providing a

EFFECT OF SUPERCHARGING ON ALTITUDE
PERFORMANCE

manifold pressure any greater than the induc­tion system inlet pressure. As altitude is increased with full throttle and a governed RPM, the airflow through the engine is reduced and ВНР decreases. The first forms of supercharging were of relatively low pressure ratio and the added airflow and power could be handled at full throttle within detonation limits. Such a “ground boosted” engine would achieve higher output power at all altitudes but an increase in altitude would produce a decrease in manifold pressure, air­flow, and power output.

More advanced forms of supercharging with higher pressure ratios can produce very large engine airflow. In fact, the typical case of altitude supercharging will produce such high airflow at low altitude operation that full throttle operation cannot be utilized within detonation limits. Figure 2.17 illustrates this case for a typical two-speed engine driven altitude supercharging installation. At sea level, the limiting manifold pressure produces a certain amount of ВНР. Full throttle oper­ation could produce a higher MAP and ВНР if detonation were not the problem. In this case full throttle operation is unavailable because of detonation limits. As altitude is increased with the supercharger or "blower” at low speed, the constant MAP is maintained by opening the throttle and the ВНР increases above the sea level value because of the re­duced exhaust back pressure. Opening the throttle allows the supercharger inlet to re­ceive the same inlet pressure and produce the same MAP. Finally, the increase of altitude will require full throttle to produce the con­stant MAP with low blower and this point is termed the "critical altitude” or "full throttle height.” If altitude is increased beyond the critical altitude, the engine MAP, airflow, and ВНР decrease.

The critical altitude with a particular super­charger installation is specific to a given com­bination of MAP and RPM. Obviously, a lower MAP could be maintained to some higher altitude or a lower engine speed would produce less supercharging and a given MAP would require a greater throttle opening. Generally, the most important critical alti­tudes will be specified for maximum, rated, and maximum cruise power conditions.

A change of the blower to a high speed will provide greater supercharging but will require more shaft power and incur a greater tempera­ture rise. Thus, the high blower speed can produce an increase in altitude performance within the detonation limitations. The vari­ation of ВНР with altitude for the blower at high speed shows an increase in critical alti­tude and greater ВНР than is obtainable in low blower. Operation below the high blower critical altitude requires some limiting mani­fold pressure to remain within detonation limits. It is apparent that the shift to high blower is not required just past low blower critical altitude but at the point where the transition from low blower, full throttle to high blower, limit MAP will produce greater ВНР. Of course, if the blower speed is increased without reducing the throttle opening, an "overboost” can occur.

Since the exhaust gases have considerable energy, exhaust turbines provide a source of supercharger power. The turbosupercharger (TBS’) allows control of the supercharger speed and output to very high altitudes with a variable discharge exhaust turbine (VDT). The turbosupercharger is capable of providing the engine airflow with increasing altitude by increasing turbine and supercharger speed. Critical altitude for the turbosupercharger is usually defined by the altitude which produces the limiting exhaust turbine speed.

The minimum specific fuel consumption of the supercharged engine is not greatly affected by altitudes less than the critical altitude. At the maximum cruise power condition, specific fuel consumption will decrease slightly with an increase in altitude up to the critical altitude. Above critical altitude, maximum cruise power cannot be maintained but the

specific fuel consumption is not adversely affected as long as auto-lean or manual lean power can be used at the cruise power setting.

One operating characteristic of the recipro­cating engine is distinctly different from that of the turbojet. Wafer vapor in the air will cause a significant reduction in output power of the reciprocating engine but a negligible loss of thrust for the turbojet engine. This basic difference exists because the reciprocating engine operates with a fixed displacement and all air processed is directly associated with the combustion process. If water vapor enters the induction system of the reciprocating engine, the amount of air available for combustion is reduced and, since most carburetors do not distinguish water vapor from air, an enrich­ment of the fuel-air ratio takes place. The maximum power output at takeoff requires fuel-air ratios richer than that for maximum heat release so a further enrichment will take place with subsequent loss of power. The turbojet operates with such great excess of air that the combustion process essentially is unaffected and the reduction of air mass flow is the principal consideration. As an example, extreme conditions which would produce high specific humidity may cause a 3 percent thrust loss for a turbojet but a 12 percent loss of ВНР for a reciprocating engine. Proper accounting of the loss due to humidity is essential in the operation of the reciprocating engine.

OPERATING LIMITATIONS. Recipro­cating engines have achieved a great degree of refinement and development and are one of the most reliable of all types of aircraft power – plants. However, reliable operation of the re­ciprocating engine is obtained only by strict adherence to the specific operating limitations.

The most important operating limitations of the reciprocating engine are those provided to ensure that detonation and preignition do not take place. The pilot must ensure that proper fuel grades are used that limit MAP, ВМЕР, RPM, CAT, etc., are not exceeded. Since

fluid from fouling the plumbing.

When the fuel grades are altered during oper­ation and the engine must be operated on a next lower fuel grade, proper account must be made for the change in the operating limita­tions. This accounting must be made for the maximum power for takeoff and the maximum cruise power since both of these operating con­ditions are near the detonation envelope. In addition, when the higher grade of fuel again becomes available, the higher operating limits cannot be used until it is sure that no contamina­tion exists from the lower grade fuel remaining in the tanks.

Spark plug fouling can provide certain high as well as low limits of operating temperatures. When excessively low operating temperatures are encountered, rapid carbon fouling of the plugs will take place. On the other hand, excessively high operating temperatures will produce plug fouling from lead bromide de­posits from the fuel additives.

Generally, the limited periods of time at various high power settings arc set to mini­mize the accumulation of high rates of wear

Revised January 1965

and fatigue damage. By minimizing the amount of total time spent at high power setting, greater overhaul life of the powerplant can be achieved. This should not imply that the takeoff rating of the engine should not be used. Actually, the use of the full maximum power at takeoff will accumulate less total engine wear than a reduced power setting at the same RPM because of less time required to climb to a given altitude or to accelerate to a given speed.

The most severe rate of wear and fatigue damage occurs at high RPM and low MAP. High RPM produces high centrifugal loads and reciprocating inertia loads. When the large reciprocating inertia loads are not cush­ioned by high compression pressures, critical resultant loads can be produced. Thus, op­erating time at maximum RPM and MAP must be held to a minimum and operation at maxi­mum RPM and low MAP must be avoided.