You are on page 1of 58

Theory of Flight - Level 3

Why learn about Theory of Flight?

The pilot today has a large variety of airplanes from which to choose. Of these airplanes many may fly at less than 100 knots top speed while others are capable of speeds well into the hundreds of knots. Some are single seaters carrying only the pilot, while others, even in the single engine light airplane class may carry 10 or more passengers. Some airplanes have laminar flow airfoil sections; others have airfoils of conventional design. A few light airplanes fly at 3 1/2 times their stalling speed; others do well to cruise at 1 1/3 times their stalling speed. Every one of these airplanes has different flight characteristics. If a pilot has a good grasp of the fundamentals of flight, he will understand what to expect of each different airplane that he may have the opportunity to fly. He will understand how best to handle each airplane as a result of his knowledge of the theory of flight and of the airplanes design. He will comprehend the various loads to which an airplane of a particular design may be exposed while flying under abnormal or adverse conditions of flight. Not only to get the best performance but also to ensure the safety of each flight, an understanding of Theory of Flight is very essential. The study of theory of flight and aerodynamics can be a lifetime proposition. New theories are forever being put forward. Some questions have answers that are difficult to find. Others perhaps do not yet have adequate answers. The information that comprises this section can only be considered an introduction to a substantial, but fascinating, topic in aviation.

Sections in this Chapter:


Forces Acting on an Airplane in Flight How Airplanes Fly: Physical Description of Lift Wing Design
CONVENTIONAL AIRFOILS and LAMINAR FLOW AIRFOILS PLANFORM DESIGN and OTHER ISSUES OTHER WING ADDITIONS

WING BOUNDARY LAYER

Fluid Dynamics
FLOW EQUATIONS INCOMPRESSIBLE FLOW AROUND A CYLINDER and A WING SECTION FOILSIM - BASIC AERODYNAMICS SOFTWARE

Axes of an Airplane
THE AXES OF AN AIRPLANE BALANCED CONTROLS STABILITY OF AN AIRPLANE

Propeller Aircraft Performance


WHAT IS AIRCRAFT PERFORMANCE? THE BOOTSTRAP APPROACH: BACKGROUND THE BOOTSTRAP APPROACH: FORMULAS AND GRAPHS CONCLUSION & REFERENCES

Flight Performance Flight Instruments Navigation Systems Basic Instrument Flying Flight Environment-weather

Forces Acting on an Airplane in Flight - Level 3


There are four forces acting on an airplane in flight. These are thrust, drag, lift and weight(gravity). 1. THRUST. The force exerted by the engine and its propeller(s), which pushes

air backward with the object of causing a reaction, or thrust, of the airplane in the forward direction. 2. DRAG. The resistance of the airplane to forward motion directly opposed to thrust. 3. LIFT. The upward force created by the wings moving through the air, which sustains the airplane in flight. 4. WEIGHT. The downward force due to the weight(gravity) of the airplane and its load, directly opposed to lift. When thrust and drag are equal and opposite, the airplane is said to be in a state of equilibrium. That is to say, it will continue to move forward at the same uniform speed. (Equilibrium refers to steady motion and not to a state of rest, in this context) If either of these forces becomes greater than the force opposing it, the state of equilibrium will be lost. If thrust is greater than drag, the airplane will accelerate or gain speed. If drag is greater than thrust, the airplane will decelerate or lose speed and consequently, the airplane will descend. Similarly, when lift and weight are equal and opposite, the airplane will be in equilibrium. If lift, however, is greater than weight, the airplane will climb. If weight is greater than lift, the airplane will sink.

Forces Acting on an Airplane in Flight - Level 3


Thrust

Thrust is the force that provides the forward motion of the airplane through the air. There are several ways to produce this forcejets, propellers or rockets but they all depend on the principle of pushing air backward with the object of causing a reaction, or thrust, in the forward direction. The effect is the same whether the thrust is produced by a propeller moving a large mass of air backward at a relatively slow speed or by a jet moving a small mass of air backward at a relatively high speed. For jet aircraft, the means of thrust is the gas turbine engine. The figure below shows the inlet and exhaust flows of the turbojet. The negative thrust due to bringing the freestream air almost to rest just ahead of the engine is called momentum drag or ram drag. The resulting thrust is given by following equation,

Schematic of a turbojet engine.

where: = is weight flow rate of the air passing through the engine. = jet stream velocity = static pressure across propelling nozzle = atmospheric pressure = propelling nozzle area = aircraft speed

Drag - Level 3

Drag is the resistance an airplane experiences in moving forward through the air. For an airplane to maintain steady flight, there must be sufficient lift to balance the weight of the airplane, and there must be sufficient thrust to overcome drag. Fig. 1 shows an airfoil moving forward through the air and depicts the principle known as the resolution of forces. The vertical component (OL) is the lift and is used to support the weight of the airplane. The horizontal component (OD) is the drag. OR is the resultant reaction of these two components.

Since drag is a force directly opposed to the motion of the airfoil and, as the work of overcoming it is performed by the engine, it is desirable to have it as small as possible, to afford the engine to be more efficient. Drag is of two principal types. 1. PARASITE DRAG is the term given to the drag of all those parts of the airplane which do not contribute to lift, that is, the fuselage, landing gear, struts, antennas, wing tip fuel tanks, etc. In addition, any loss of momentum of the airstream caused by openings, such as those in the cowling and those between the wing and the ailerons and the flaps, add to parasite drag. Parasite drag may be divided into two components.
1. Form drag refers to the drag created by the form or shape of a body as it resists motion through the air. 2. Skin friction refers to the tendency of air flowing over a body to cling to its surface. 3. Although parasite drag can never be completely eliminated, it can be substantially reduced. One method is to eliminate altogether those parts of the airplane that cause it. For this reason, retractable landing gear has been developed. Wing struts have been eliminated in favour of fully cantilevered wings. Another method is to streamline those parts that cannot be eliminated. Skin friction can be reduced substantially by the removal of dust, dirt, mud or ice that has collected on the airplane.

Even the most carefully designed individual parts must, however, be joined together to create the total airplane. Resistance caused by the effect of one part on another (i.e. where the wing is attached to the fuselage, or the struts to the wings) is called interference drag and can be reduced by careful design in the fairing of one shape into another. 2. INDUCED DRAG is caused by those parts of an airplane which are active in producing lift (i.e. the wing). It is the result of the wing's work in sustaining the airplane in flight and is, therefore, a part of the lift and can never be eliminated. It increases as the angle of attack increases and decreases as the- angle of attack decreases.

Induced drag can be reduced only during the initial designing of an airplane. A wing with a high aspect ratio, that is, with a very long span and a narrow chord, produces less induced drag than does a wing with a short span and a wide chord. Gliders and sail planes are therefore commonly designed with high aspect ratio wings. The phenomenon, known as wing tip vortices, is testimony to the existence of induced drag. As the decreased pressure over the top of the wing is less than the atmospheric pressure around it, the air flowing over the top surface tends to flow inward. The air flowing over the lower surface, due to the relatively higher pressure around it, tends to flow outward and curl upward over the wing tips. When the two airflows unite at the trailing edge of the wing, they are flowing contra-wise. Eddies and vortices are formed which tend to unite into one large eddy at each wing tip. These are called wing-tip vortices. This disturbed air exerts a resistant force against the forward motion of the wing. This resistant force is known as induced drag. In order to support the weight of an airplane, a large amount of air must be displaced downward. This displaced air must have somewhere to go, and tends to flow spanwise outwards, as explained above. It is seeking to escape around the wing tips and flow into the low pressure area created over the upper surface of the wing. It will be become very clear that the heavier the airplane and the higher the span loading on the wing, the more air it will displace downward, therefore the greater will be the circulation of air, and the greater the magnitude of the wing tip vortex created and the greater the induced drag. Induced drag does not increase as the speed increases. On the contrary, it is greatest when the airplane is flying slowly, a few knots above the stalling speed when maximum lift is being realized at minimum speed. The induced drag characteristics of a wing are not the same very near the ground as they are at altitude. During landing and take-off, the ground interferes with the formation of a large wing-tip vortex. Induced drag is, therefore, reduced when an airplane is flown very near the ground. This phenomenon is known as ground effect. Although induced drag cannot be eliminated, it can be reduced by certain design features. As has been stated earlier, less induced drag is generated by a long, narrow wing than by a short, broad one. It has also been found that winglets are effective in reducing induced drag. Attached to the wing tip, the winglet, a small, vertical surface of airfoil section, is effective in producing side forces that diffuse the wind-tip vortex flow. See Lift and Drag Curves

Aileron Drag
In banking to make an airplane turn, one aileron is depressed and the other is raised. The downgoing aileron, being depressed into the compressed airflow on the underside of the wing,

causes drag. The upgoing aileron, moving up into a more streamlined position, causes less drag. The drag on the downgoing aileron is known as aileron drag and if not corrected for in the design of the aileron, tends to cause a yaw in the opposite direction to which the bank is applied.

The Boundary Layer


The boundary layer is a very thin layer of air lying over the surface of the wing (and, for that matter, all other surfaces of the airplane). Because air has viscosity, this layer of air tends to adhere to the wing. As the wing moves forward through the air, the boundary layer at first flows smoothly over the streamlined shape of the airfoil. Here the flow is called the laminar layer. As the boundary layer approaches the center of the wing, it begins to lose speed due to skin friction and it becomes thicker and turbulent. Here it is called the turbulent layer. The point at which the boundary layer changes from laminar to turbulent is called the transition point (Fig. 3). Where the boundary layer becomes turbulent, drag due to skin friction is relatively high. As speed increases, the transition point tends to move forward. As the angle of attack increases, the transition point also tends to move forward.

Fig. 3

Various methods have been developed to control the boundary layer in order to reduce skin friction drag. Suction Method. One method uses a series of thin slots in the wing running out from the wing root towards the tip. A vacuum sucks the air down through the slots, preventing the airflow from breaking away from the wing and forcing it to follow the curvature of the wing surface. The air, which is sucked in, siphons through the ducts inside the wing and is exhausted backwards to provide a little extra thrust. The laminar flow airfoil is itself a structural design intended to make possible better boundary layer control. The thickest part of a laminar flow wing occurs at 50% chord. The transition point at which the laminar flow of air breaks down into turbulence is at or near the thickest part. The transition point at which the laminar flow of air becomes turbulent on a laminar flow airfoil is rearward of that same point on a conventional designed airfoil (Fig. 4).

Fig. 4

Vortex generators are small plates about an inch deep standing on edge in a row spanwise along the wing. They are placed at an angle of attack and (like a wing airfoil section) generate vortices. These tend to prevent or delay the breakaway of the boundary layer by re-energizing it. They are lighter and simpler than the suction boundary layer control system described above. Streamlining is a design feature by which a body is so shaped that drag is minimized as the body moves forward through the air. A flat plate or a round ball moving through the air disturb the smooth flow of air and set up eddies behind them. The effect of streamlining an object can be seen in Fig. 5.

Fig. 5

COUPLES
The principle of equilibrium has already been discussed in earlier discussions of the forces that act on an airplane in flight. When two forces, such as thrust and drag, are equal and opposite, but displaced parallel to each other rather than passing through the same point, they are said to form a couple. A couple will cause a turning moment about a given axis. If the center of gravity of weight is ahead of lift on a wing, the couple created will turn the nose of the airplane down. Conversely, if lift is ahead of weight, the couple created will turn the nose of the airplane up. If drag is above thrust, the couple formed will turn the nose of the airplane up. Conversely, if thrust is above drag, the couple formed will turn the nose of the airplane down.

Lift - Level 3
If you consider the definitions cited by air authorities, a boy flying a kite could be construed to be a pilot in charge of an airplane! Ponder the idea a moment and it may not appear quite as absurd as it seems at first glance. A kite is an inclined plane, the weight of which is supported in the air by the reaction of the wind flowing against it. If we substitute for the string, which holds the kite against the wind, the engine and propeller of an airplane, which move the wings forward against the airflow, we will see that the analogy of the kite is not without some validity. The wings of an airplane are so designed that when moved through the air horizontally, the force exerted on them produces a reaction as nearly vertical as possible. It is this reaction that lifts the weight of the airplane.

Airfoils
An airfoil, or airfoil section, may be defined as any surface designed to obtain a reaction from the air through which it moves, that is, to obtain lift. It has been found that the most suitable shape for producing lift is a curved or cambered shape.

The camber of an airfoil is the curvature of the upper and lower surfaces. Usually the upper surface has a greater camber than the lower.

How Is Lift Created


What, then, causes this lift, you may ask. Air flowing around an airfoil is subject to the Laws of Motion discovered by Sir Isaac Newton. Air, being a gaseous liquid possesses inertia and, therefore, according to Newton's First Law, when in motion tends to remain in motion. The introduction of an airfoil into the streamlined airflow alters the uniform flow of air. Newton's Second Law states that a force must be applied to alter the state of uniform motion of a body. The airfoil is the force that acts on the body (in this case, the air) to produce a change of direction. The application of such a force causes an equal and opposite reaction, in compliance with Newtons Third Law, called, in this case, lift.

As the air passes over the wing towards the trailing edge, the air flows not only rearward but downward as well. This flow is called downwash. At the same time, the airflow passing under the wing is deflected downward by the bottom surface of the wing. Think of a water ski or surfboard planing over the water. In exerting a downward force upon the air, the wing receives an upward counterforce. Remember Newton's Third Lawfor every action there is an equal and opposite reaction. Therefore, the more air deflected downward, the more lift is created. Air is heavy; its weight exerts a pressure of 14.7 Ibs per square inch at sea level. The reaction produced by the downwash is therefore significant.

The phenomenon defined by Bernoulli's Principle also has an effect in the production of lift by the wing moving through the air. Scientist Daniel Bernoulli discovered that the total energy in any system remains constant. In other words, if one element of an energy system is increased, another decreases to counter balance it. Take the example of water flowing through a venturi tube. Being incompressible, the water must speed up to pass through the constricted space of the venturi. The moving water has energy in the form of both pressure and speed. Within the venturi tube, pressure is sacrificed (decreased) to accelerate the speed of the flow. Air is a fluid, just like water, and can be assumed incompressible insofar as speed aerodynamics is concerned. As such, it acts exactly the same way as water in a venturi tube. Picture the curved upper surface of a wing as the bottom half of a venturi tube. The upper half of this imaginary tube is the undisturbed airflow above the wing. Air flowing over the wing's upper surface accelerates as it passes through the constricted area just as it does in the venturi tube. The result is a decrease in pressure on the upper surface of the wing that results in the phenomenon known as lift.

Relative Airflow
Relative airflow is a term used to describe the direction of the airflow with respect to the wing. In other texts, it is sometimes called relative wind. If a wing is moving forward and downward, the relative airflow is upward and backward. If the wing is moving forward horizontally, the relative airflow moves backward horizontally. The flight path and the relative airflow are, therefore, always parallel but travel in opposite directions.

Relative airflow is created by the motion of the airplane through the air. It is also created by the motion of air past a stationary body or by a combination of both. Therefore, on a take-off roll, an airplane is subject to the relative airflow created by its motion along the ground and also by the moving mass of air (wind). In flight, however, only the motion of the airplane produces a relative airflow. The direction and speed of the wind have no effect on relative airflow.

Angle of Attack and Center of Pressure


The angle at which the airfoil meets the relative airflow is called the angle of attack. As the angle of attack is increased, the changes in pressure over the upper and lower surfaces and the amount of downwash, that is air deflected downward, increase up to a point (the stalling angle). Beyond this angle, these changes in pressure decreases. If we consider all the distributed pressure on the wing to be equivalent to a single force, this force will act through a straight line. The point where this line cuts the chord of an airfoil is called the center of pressure. Thus, it will be seen that as the angle of attack of an airfoil is increased up to the point of stall, the center of pressure will move forward. Beyond this point of stall, the center of pressure will move back. The movement of the center of pressure causes an airplane to be unstable. See Lift and Drag Curves

Weight - Level 3

The weight of an airplane is the force, which acts vertically downward toward the center of the earth and is the result of gravity on the airplane. Just as the lift of an airplane acts through the center of pressure, the weight of an airplane acts through the center of gravity (C.G.). This is the point through which the resultant of the weights of all the various parts of the airplane passes, in every attitude that the airplane can assume.

How Airplanes Fly: A Physical Description of Lift Level 3


Almost everyone today has flown in an airplane. Many ask the simple question "what makes an airplane fly"? The answer one frequently gets is misleading and often just plain wrong. We hope that the answers provided here will clarify many misconceptions about lift and that you will adopt our explanation when explaining lift to others. We are going to show you that lift is easier to understand if one starts with Newton rather than Bernoulli. We will also show you that the popular explanation that most of us were taught is misleading at best and that lift is due to the wing diverting air down. Let us start by defining three descriptions of lift commonly used in textbooks and training manuals. The first we will call the Mathematical Aerodynamics Description which is used by aeronautical engineers. This description uses complex mathematics and/or computer simulations to calculate the lift of a wing. These are design tools which are powerful for computing lift but do not lend themselves to an intuitive understanding of flight. The second description we will call the Popular Explanation which is based on the Bernoulli principle. The primary advantage of this description is that it is easy to understand and has been taught for many years. Because of its simplicity, it is used to describe lift in most flight training manuals. The major disadvantage is that it relies on the "principle of equal transit times" which is wrong. This description focuses on the shape of the wing and prevents one from understanding such important phenomena as inverted flight, power, ground effect, and the dependence of lift on the angle of attack of the wing. The third description, which we are advocating here, we will call the Physical Description of lift. This description is based primarily on Newtons laws. The physical description is useful for understanding flight, and is accessible to all that are curious. Little math is needed to yield an

estimate of many phenomena associated with flight. This description gives a clear, intuitive understanding of such phenomena as the power curve, ground effect, and high-speed stalls. However, unlike the mathematical aerodynamics description, the physical description has no design or simulation capabilities.

The popular explanation of lift


Students of physics and aerodynamics are taught that airplanes fly as a result of Bernoullis principle, which says that if air speeds up the pressure is lowered. Thus a wing generates lift because the air goes faster over the top creating a region of low pressure, and thus lift. This explanation usually satisfies the curious and few challenge the conclusions. Some may wonder why the air goes faster over the top of the wing and this is where the popular explanation of lift falls apart. In order to explain why the air goes faster over the top of the wing, many have resorted to the geometric argument that the distance the air must travel is directly related to its speed. The usual claim is that when the air separates at the leading edge, the part that goes over the top must converge at the trailing edge with the part that goes under the bottom. This is the so-called "principle of equal transit times". As discussed by Gale Craig (Stop Abusing Bernoulli! How Airplanes Really Fly., Regenerative Press, Anderson, Indiana, 1997), let us assume that this argument were true. The average speeds of the air over and under the wing are easily determined because we can measure the distances and thus the speeds can be calculated. From Bernoullis principle, we can then determine the pressure forces and thus lift. If we do a simple calculation we would find that in order to generate the required lift for a typical small airplane, the distance over the top of the wing must be about 50% longer than under the bottom. Figure 1 shows what such an airfoil would look like. Now, imagine what a Boeing 747 wing would have to look like!

Fig 1 Shape of wing predicted by principle of equal transit time. If we look at the wing of a typical small plane, which has a top surface that is 1.5 - 2.5% longer than the bottom, we discover that a Cessna 172 would have to fly at over 400 mph to generate enough lift. Clearly, something in this description of lift is flawed. But, who says the separated air must meet at the trailing edge at the same time? Figure 2 shows the airflow over a wing in a simulated wind tunnel. In the simulation, colored smoke is introduced periodically. One can see that the air that goes over the top of the wing gets to the

trailing edge considerably before the air that goes under the wing. In fact, close inspection shows that the air going under the wing is slowed down from the "free-stream" velocity of the air. So much for the principle of equal transit times.

Fig 2 Simulation of the airflow over a wing in a wind tunnel, with colored "smoke" to show the acceleration and deceleration of the air. The popular explanation also implies that inverted flight is impossible. It certainly does not address acrobatic airplanes, with symmetric wings (the top and bottom surfaces are the same shape), or how a wing adjusts for the great changes in load such as when pulling out of a dive or in a steep turn? So, why has the popular explanation prevailed for so long? One answer is that the Bernoulli principle is easy to understand. There is nothing wrong with the Bernoulli principle, or with the statement that the air goes faster over the top of the wing. But, as the above discussion suggests, our understanding is not complete with this explanation. The problem is that we are missing a vital piece when we apply Bernoullis principle. We can calculate the pressures around the wing if we know the speed of the air over and under the wing, but how do we determine the speed? Another fundamental shortcoming of the popular explanation is that it ignores the work that is done. Lift requires power (which is work per time). As will be seen later, an understanding of power is key to the understanding of many of the interesting phenomena of lift.

Newton s laws and lift


So, how does a wing generate lift? To begin to understand lift we must return to high school physics and review Newtons first and third laws. (We will introduce Newtons second law a little later.) Newtons first law states a body at rest will remain at rest, or a body in motion will continue in straight-line motion unless subjected to an external applied force. That means, if one sees a bend in the flow of air, or if air originally at rest is accelerated into motion, there is a force acting on it. Newtons third law states that for every action there is an equal and opposite reaction. As an example, an object sitting on a table exerts a force on the table (its weight) and the table puts an equal and opposite force on the object to hold it up. In order to generate lift a

wing must do something to the air. What the wing does to the air is the action while lift is the reaction. Lets compare two figures used to show streams of air (streamlines) over a wing. In figure 3 the air comes straight at the wing, bends around it, and then leaves straight behind the wing. We have all seen similar pictures, even in flight manuals. But, the air leaves the wing exactly as it appeared ahead of the wing. There is no net action on the air so there can be no lift! Figure 4 shows the streamlines, as they should be drawn. The air passes over the wing and is bent down. The bending of the air is the action. The reaction is the lift on the wing.

Fig 3 Common depiction of airflow over a wing. This wing has no lift.

Fig 4 True airflow over a wing with lift, showing upwash and downwash.

The wing as a pump


As Newtons laws suggests, the wing must change something of the air to get lift. Changes in the airs momentum will result in forces on the wing. To generate lift a wing must divert air down; lots of air. The lift of a wing is equal to the change in momentum of the air it is diverting down. Momentum is the product of mass and velocity. The lift of a wing is proportional to the amount of air diverted down times the downward velocity of that air. Its that simple. (Here we have used an alternate form of Newtons second law that relates the acceleration of an object to its mass and to the force on it; F=ma) For more lift the wing can either divert more air (mass) or increase its downward velocity. This downward velocity behind the wing is called "downwash". Figure 5 shows how the downwash appears to the pilot (or in a wind tunnel). The figure also shows how the downwash appears to an observer on the ground watching the wing go by. To the pilot the air is coming off the wing at roughly the angle of attack. To the observer on the ground, if he or she

could see the air, it would be coming off the wing almost vertically. The greater the angle of attack, the greater the vertical velocity. Likewise, for the same angle of attack, the greater the speed of the wing the greater the vertical velocity. Both the increase in the speed and the increase of the angle of attack increase the length of the vertical arrow. It is this vertical velocity that gives the wing lift.

Fig 5 How downwash appears to a pilot and to an observer on the ground. As stated, an observer on the ground would see the air going almost straight down behind the plane. This can be demonstrated by observing the tight column of air behind a propeller, a household fan, or under the rotors of a helicopter; all of which are rotating wings. If the air were coming off the blades at an angle the air would produce a cone rather than a tight column. If a plane were to fly over a very large scale, the scale would register the weight of the plane. If we estimate that the average vertical component of the downwash of a Cessna 172 traveling at 110 knots to be about 9 knots, then to generate the needed 2,300 lbs of lift the wing pumps a whopping 2.5 ton/sec of air! In fact, as will be discussed later, this estimate may be as much as a factor of two too low. The amount of air pumped down for a Boeing 747 to create lift for its roughly 800,000 pounds takeoff weight is incredible indeed. Pumping, or diverting, so much air down is a strong argument against lift being just a surface effect as implied by the popular explanation. In fact, in order to pump 2.5 ton/sec the wing of the Cessna 172 must accelerate all of the air within 9 feet above the wing. (Air weighs about 2 pounds per cubic yard at sea level.) Figure 6 illustrates the effect of the air being diverted down from a wing. A huge hole is punched through the fog by the downwash from the airplane that has just flown over it.

Fig 6 Downwash and wing vortices in the fog. (Photographer Paul Bowen, courtesy of Cessna Aircraft, Co.) So how does a thin wing divert so much air? When the air is bent around the top of the wing, it pulls on the air above it accelerating that air down, otherwise there would be voids in the air left above the wing. Air is pulled from above to prevent voids. This pulling causes the pressure to become lower above the wing. It is the acceleration of the air above the wing in the downward direction that gives lift. (Why the wing bends the air with enough force to generate lift will be discussed in the next section.) As seen in figure 4, a complication in the picture of a wing is the effect of "upwash" at the leading edge of the wing. As the wing moves along, air is not only diverted down at the rear of the wing, but air is pulled up at the leading edge. This upwash actually contributes to negative lift and more air must be diverted down to compensate for it. This will be discussed later when we consider ground effect. Normally, one looks at the air flowing over the wing in the frame of reference of the wing. In other words, to the pilot the air is moving and the wing is standing still. We have already stated that an observer on the ground would see the air coming off the wing almost vertically. But what is the air doing above and below the wing? Figure 7 shows an instantaneous snapshot of how air molecules are moving as a wing passes by. Remember in this figure the air is initially at rest and it is the wing moving. Ahead of the leading edge, air is moving up (upwash). At the trailing edge, air is diverted down (downwash). Over the top the air is accelerated towards the trailing edge. Underneath, the air is accelerated forward slightly, if at all.

Fig 7 Direction of air movement around a wing as seen by an observer on the ground. In the mathematical aerodynamics description of lift this rotation of the air around the wing gives rise to the "bound vortex" or "circulation" model. The advent of this model, and the complicated mathematical manipulations associated with it, leads to the direct understanding of forces on a wing. But, the mathematics required typically takes students in aerodynamics some time to master. One observation that can be made from figure 7 is that the top surface of the wing does much more to move the air than the bottom. So the top is the more critical surface. Thus, airplanes can carry external stores, such as drop tanks, under the wings but not on top where they would interfere with lift. That is also why wing struts under the wing are common but struts on the top of the wing have been historically rare. A strut, or any obstruction, on the top of the wing would interfere with the lift.

Air has viscosity


The natural question is "how does the wing divert the air down?" When a moving fluid, such as air or water, comes into contact with a curved surface it will try to follow that surface. To demonstrate this effect, hold a water glass horizontally under a faucet such that a small stream of water just touches the side of the glass. Instead of flowing straight down, the presence of the glass causes the water to wrap around the glass as is shown in figure 8. This tendency of fluids to follow a curved surface is known as the Coanda effect. From Newtons first law we know that for the fluid to bend there must be a force acting on it. From Newtons third law we know that the fluid must put an equal and opposite force on the object which caused the fluid to bend.

Fig 8 Coanda effect. Why should a fluid follow a curved surface? The answer is viscosity; the resistance to flow which also gives the air a kind of "stickiness". Viscosity in air is very small but it is enough for the air molecules to want to stick to the surface. At the surface the relative velocity between the surface and the nearest air molecules is exactly zero. (That is why one cannot hose the dust off of a car and why there is dust on the backside of the fans in a wind tunnel.) Just above the surface the fluid has some small velocity. The farther one goes from the surface the faster the fluid is moving until the external velocity is reached (note that this occurs in less than an inch). Because the fluid near the surface has a change in velocity, the fluid flow is bent towards the surface. Unless the bend is too tight, the fluid will follow the surface. This volume of air around the wing that appears to be partially stuck to the wing is called the "boundary layer".

Lift as a function of angle of attack


There are many types of wing: conventional, symmetric, conventional in inverted flight, the early biplane wings that looked like warped boards, and even the proverbial "barn door". In all cases, the wing is forcing the air down, or more accurately pulling air down from above. What each of these wings have in common is an angle of attack with respect to the oncoming air. It is this angle of attack that is the primary parameter in determining lift. The inverted wing can be explained by its angle of attack, despite the apparent contradiction with the popular explanation involving the Bernoulli principle. A pilot adjusts the angle of attack to adjust the lift for the speed and load. The popular explanation of lift which focuses on the shape of the wing gives the pilot only the speed to adjust. To better understand the role of the angle of attack it is useful to introduce an "effective" angle of attack, defined such that the angle of the wing to the oncoming air that gives zero lift is defined to be zero degrees. If one then changes the angle of attack both up and down one finds that the lift is proportional to the angle. Figure 9 shows the coefficient of lift (lift normalized for the size of the wing) for a typical wing as a function of the effective angle of attack. A similar lift versus angle of attack relationship is found for all wings, independent of their design. This is true for the wing of a 747 or a barn door. The role of the angle of attack is more important than the details of the airfoils shape in understanding lift.

Fig 9 Coefficient of lift versus the effective angle of attack. Typically, the lift begins to decrease at an angle of attack of about 15 degrees. The forces necessary to bend the air to such a steep angle are greater than the viscosity of the air will support, and the air begins to separate from the wing. This separation of the airflow from the top of the wing is a stall.

The wing as air "scoop"


We now would like to introduce a new mental image of a wing. One is used to thinking of a wing as a thin blade that slices though the air and develops lift somewhat by magic. The new image that we would like you to adopt is that of the wing as a scoop diverting a certain amount of air from the horizontal to roughly the angle of attack, as depicted in figure 10. The scoop can be pictured as an invisible structure put on the wing at the factory. The length of the scoop is equal to the length of the wing and the height is somewhat related to the chord length (distance from the leading edge of the wing to the trailing edge). The amount of air intercepted by this scoop is proportional to the speed of the plane and the density of the air, and nothing else.

Fig 10 The wing as a scoop. As stated before, the lift of a wing is proportional to the amount of air diverted down times the vertical velocity of that air. As a plane increases speed, the scoop diverted more air. Since the load on the wing, which is the weight of the plane, does not increase the vertical speed of the

diverted air must be decreased proportionately. Thus, the angle of attack is reduced to maintain a constant lift. When the plane goes higher, the air becomes less dense so the scoop diverts less air for the same speed. Thus, to compensate the angle of attack must be increased. The concepts of this section will be used to understand lift in a way not possible with the popular explanation.

Lift requires power


When a plane passes overhead the formerly still air ends up with a downward velocity. Thus, the air is left in motion after the plane leaves. The air has been given energy. Power is energy, or work, per time. So, lift must require power. This power is supplied by the airplanes engine (or by gravity and thermals for a sailplane). How much power will we need to fly? The power needed for lift is the work (energy) per unit time and so is proportional to the amount of air diverted down times the velocity squared of that diverted air. We have already stated that the lift of a wing is proportional to the amount of air diverted down times the downward velocity of that air. Thus, the power needed to lift the airplane is proportional to the load (or weight) times the vertical velocity of the air. If the speed of the plane is doubled the amount of air diverted down doubles. Thus the angle of attack must be reduced to give a vertical velocity that is half the original to give the same lift. The power required for lift has been cut in half. This shows that the power required for lift becomes less as the airplane's speed increases. In fact, we have shown that this power to create lift is proportional to one over the speed of the plane. But, we all know that to go faster (in cruise) we must apply more power. So there must be more to power than the power required for lift. The power associated with lift, described above, is often called the "induced" power. Power is also needed to overcome what is called "parasitic" drag, which is the drag associated with moving the wheels, struts, antenna, etc. through the air. The energy the airplane imparts to an air molecule on impact is proportional to the speed squared. The number of molecules struck per time is proportional to the speed. Thus the parasitic power required to overcome parasitic drag increases as the speed cubed. Figure 11 shows the power curves for induced power, parasitic power, and total power which is the sum of induced power and parasitic power. Again, the induced power goes as one over the speed and the parasitic power goes as the speed cubed. At low speed the power requirements of flight are dominated by the induced power. The slower one flies the less air is diverted and thus the angle of attack must be increased to maintain lift. Pilots practice flying on the "backside of the power curve" so that they recognizes that the angle of attack and the power required to stay in the air at very low speeds are considerable.

Fig 11 Power requirements versus speed. At cruise, the power requirement is dominated by parasitic power. Since this goes as the speed cubed an increase in engine size gives one a faster rate of climb but does little to improve the cruise speed of the plane. Since we now know how the power requirements vary with speed, we can understand drag, which is a force. Drag is simply power divided by speed. Figure 12 shows the induced, parasitic, and total drag as a function of speed. Here the induced drag varies as one over speed squared and parasitic drag varies as the speed squared. Taking a look at these curves one can deduce a few things about how airplanes are designed. Slower airplanes, such as gliders, are designed to minimize induced drag (or induced power), which dominates at lower speeds. Faster airplanes are more concerned with parasite drag (or parasitic power).

Fig 12 Drag versus speed.

Wing efficiency
At cruise, a non-negligible amount of the drag of a modern wing is induced drag. Parasitic drag, which dominates at cruise, of a Boeing 747 wing is only equivalent to that of a 1/2-inch cable of the same length. One might ask what effects the efficiency of a wing. We saw that the induced power of a wing is proportional to the vertical velocity of the air. If the length of a wing were to be doubled, the size of our scoop would also double, diverting twice as much air. So, for the same lift the vertical velocity (and thus the angle of attack) would have to be halved. Since the induced power is proportional to the vertical velocity of the air, it too is reduced by half. Thus, the lifting efficiency of a wing is proportional to one over the length of the wing. The longer the wing the less induced power required to produce the same lift, though this is achieved with and increase in parasitic drag. Low speed airplanes are effected more by induced drag than fast airplanes and so have longer wings. That is why sailplanes, which fly at low speeds, have such long wings. High-speed fighters, on the other hand, feel the effects of parasite drag more than our low speed trainers. Therefore, fast airplanes have shorter wings to lower parasite drag. There is a misconception by some that lift does not require power. This comes from aeronautics in the study of the idealized theory of wing sections (airfoils). When dealing with an airfoil, the picture is actually that of a wing with infinite span. Since we have seen that the power necessary for lift is proportional to one over the length of the wing, a wing of infinite span does not require power for lift. If lift did not require power airplanes would have the same range full as they do empty, and helicopters could hover at any altitude and load. Best of all, propellers (which are rotating wings) would not require power to produce thrust. Unfortunately, we live in the real world where both lift and propulsion require power.

Power and wing loading


Let us now consider the relationship between wing loading and power. Does it take more power to fly more passengers and cargo? And, does loading affect stall speed? At a constant speed, if the wing loading is increased the vertical velocity must be increased to compensate. This is done by increasing the angle of attack. If the total weight of the airplane were doubled (say, in a 2g turn) the vertical velocity of the air is doubled to compensate for the increased wing loading. The induced power is proportional to the load times the vertical velocity of the diverted air, which have both doubled. Thus the induced power requirement has increased by a factor of four! The same thing would be true if the airplanes weight were doubled by adding more fuel, etc. One way to measure the total power is to look at the rate of fuel consumption. Figure 13 shows the fuel consumption versus gross weight for a large transport airplane traveling at a constant speed (obtained from actual data). Since the speed is constant the change in fuel consumption is due to the change in induced power. The data are fitted by a constant (parasitic power) and a term that goes as the load squared. This second term is just what was predicted in our Newtonian discussion of the effect of load on induced power.

Fig 13 Fuel consumption versus load for a large transport airplane traveling at a constant speed. The increase in the angle of attack with increased load has a downside other than just the need for more power. As shown in figure 9 a wing will eventually stall when the air can no longer follow the upper surface. That is, when the critical angle is reached. Figure 14 shows the angle of attack as a function of airspeed for a fixed load and for a 2-g turn. The angle of attack at which the plane stalls is constant and is not a function of wing loading. The stall speed increases as the square root of the load. Thus, increasing the load in a 2-g turn increases the speed at which the wing will stall by 40%. An increase in altitude will further increase the angle of attack in a 2-g turn. This is why pilots practice "accelerated stalls" which illustrates that an airplane can stall at any speed. For any speed there is a load that will induce a stall.

Fig 14 Angle of attack versus speed for straight and level flight and for a 2-g turn.

Wing vortices
One might ask what the downwash from a wing looks like. The downwash comes off the wing as a sheet and is related to the details on the load distribution on the wing. Figure 15 shows, through condensation, the distribution of lift on an airplane during a high-g maneuver. From the figure one can see that the distribution of load changes from the root of the wing to the tip. Thus, the amount of air in the downwash must also change along the wing. The wing near the root is "scooping" up much more air than the tip. Since the root is diverting so much air the net effect is that the downwash sheet will begin to curl outward around itself, just as the air bends around the top of the wing because of the change in the velocity of the air. This is the wing vortex. The tightness of the curling of the wing vortex is proportional to the rate of change in lift along the wing. At the wing tip the lift must rapidly become zero causing the tightest curl. This is the wing tip vortex and is just a small (though often most visible) part of the wing vortex. Returning to figure 6 one can clearly see the development of the wing vortices in the downwash as well as the wing tip vortices.

Fig 15 Condensation showing the distribution of lift along a wing. The wingtip vortices are also seen. (from Patterns in the Sky, J.F. Campbell and J.R. Chambers, NASA SP-514.) Winglets (those small vertical extensions on the tips of some wings) are used to improve the efficiency of the wing by increasing the effective length of the wing. The lift of a normal wing must go to zero at the tip because the bottom and the top communicate around the end. The winglets blocks this communication so the lift can extend farther out on the wing. Since the efficiency of a wing increases with length, this gives increased efficiency. One caveat is that winglet design is tricky and winglets can actually be detrimental if not properly designed.

Ground effect
Another common phenomenon that is misunderstood is that of ground effect. That is the increased efficiency of a wing when flying within a wing length of the ground. A low-wing airplane will experience a reduction in drag by 50% just before it touches down. There is a great deal of confusion about ground effect. Many pilots (and the FAA VFR Exam-O-Gram No. 47) mistakenly believe that ground effect is the result of air being compressed between the wing and the ground. To understand ground effect it is necessary to have an understanding of upwash. For the pressures involved in low speed flight, air is considered to be non-compressible. When the air is accelerated over the top of the wing and down, it must be replaced. So some air must shift around the wing (below and forward, and then up) to compensate, similar to the flow of water around a canoe paddle when rowing. This is the cause of upwash. As stated earlier, upwash is accelerating air in the wrong direction for lift. Thus a greater amount of downwash is necessary to compensate for the upwash as well as to provide the necessary lift. Thus more work is done and more power required. Near the ground the upwash is reduced because the ground inhibits the circulation of the air under the wing. So less downwash is necessary to provide the lift. The angle of attack is reduced and so is the induced power, making the wing more efficient. Earlier, we estimated that a Cessna 172 flying at 110 knots must divert about 2.5 ton/sec to provide lift. In our calculations we neglected the upwash. From the magnitude of ground effect, it is clear that the amount of air diverted is probably more like 5 ton/sec.

Conclusions
Let us review what we have learned and get some idea of how the physical description has given us a greater ability to understand flight. First what have we learned:
y y y y

The amount of air diverted by the wing is proportional to the speed of the wing and the air density. The vertical velocity of the diverted air is proportional to the speed of the wing and the angle of attack. The lift is proportional to the amount of air diverted times the vertical velocity of the air. The power needed for lift is proportional to the lift times the vertical velocity of the air.

Now let us look at some situations from the physical point of view and from the perspective of the popular explanation.
y

The plane s speed is reduced. The physical view says that the amount of air diverted is reduced so the angle of attack is increased to compensate. The power needed for lift is also increased. The popular explanation cannot address this.

The load of the plane is increased. The physical view says that the amount of air diverted is the same but the angle of attack must be increased to give additional lift. The power needed for lift has also increased. Again, the popular explanation cannot address this. A plane flies upside down. The physical view has no problem with this. The plane adjusts the angle of attack of the inverted wing to give the desired lift. The popular explanation implies that inverted flight is impossible.

As one can see, the popular explanation, which fixates on the shape of the wing, may satisfy many but it does not give one the tools to really understand flight. The physical description of lift is easy to understand and much more powerful.

Wing Design - Level 3


CONVENTIONAL AIRFOILS and LAMINAR FLOW AIRFOILS

The type of operation for which an airplane is intended has a very important bearing on the selection of the shape and design of the wing for that airplane. If the airplane is designed for low speed, a thick airfoil is most efficient. A thin airfoil is best for high speed.

CONVENTIONAL AIRFOILS and LAMINAR FLOW AIRFOILS , are in common use in airplane design. Laminar flow airfoils were originally developed for the purpose of making an airplane fly faster. The laminar flow wing is usually thinner than the conventional airfoil, the leading edge is more pointed and its upper and lower surfaces are nearly symmetrical. The major and most important difference between the two types of airfoil is this, the thickest part of a laminar wing occurs at 50% chord while in the conventional design the thickest part is at 25% chord. The effect achieved by this type of design of a wing is to maintain the laminar flow of air throughout a greater percentage of the chord of the wing and to control the transition point. Drag is therefore considerably reduced since the

laminar airfoil takes less energy to slide through the air. The pressure distribution on the laminar flow wing is much more even since the camber of the wing from the leading edge to the point of maximum camber is more gradual than on the conventional airfoil. However, at the point of stall, the transition point moves more rapidly forward. NACA AIRFOIL NUMBERING SYSTEM Many times you will see airfoils described as NACA xxxx or NACA xxxxx or NACA xxy-xxx series. From the book Airplane Aerodynamics, by Dommasch, Sherby and Connally, Pitman Press, 1967, the following definitions are given to this nomenclature.

The NACA 4-digit airfoils mean the following: The first digit expresses the camber in percent chord, the second digit gives the location of the maximum camber point in tenths of chord, and the last two digits give the thickness in percent chord. Thus 4412 has a maximum camber of 4% of chord located at 40% chord back from the leading edge and is 12% thick, while 0006 is a symmetrical section of 6% thickness. The NACA 5 digit series airfoil means the following: The first digit designates the approximate camber in percent chord, the second digit indicates twice the position of the maximum camber in tenths chord, the third (either 0 or 1) distinguishes the type of mean-camber line, and the last two digits give the thickness in percent chord. Thus, the 23012 airfoil has a maximum camber of about 2% of the chord located at 15% of the chord from the leading edge (3 tenths divided by 2) and is 12% thick. The NACA six, seven and even eight series were designed to highlight some aerodynamic characteristic. For example, NACA 653-421 is a 6-series airfoil for which the minimum pressure's position in tenths chord is indicated by the second digit (here, at the 50% chord location), the subscript 3 means that the drag coefficient is near its minimum value over a range of lift coefficients of 0.3 above and below the design lift coefficient, the next digit indicates the lift coefficient in tenths (here, 0.4) and the last two digits give the maximum thickness in percent chord (here, 21% of chord). The description for this example comes from Foundations of Aerodynamics, Kuethe and Schetzer, 2nd Edition, 1959, John Wiley and Sons, New York. There are formulas that define all the stations of the airfoil section from these digits and you can probably find those in your library in any good aerodynamics book. Also, you are referred to

two other references listed below for more information on these classifications. HOWEVER, in all cases, the last two digits of the classification gives the thickness in percent chord. Summary of Airfoil Data, NACA Report 824, 1945, by Abbott, von Doenhoff and Stivers. It was originally issued as ACR L5C05. Theory of Wing Sections, Including a Summary of Airfoil Data, by Abbott and von Doenhoff, Mc-Graw Hill, New York, 1949, in which families of airfoils constructed according to a certain plan were tested and their characteristics recorded.

Wing Design - Level 3


PLANFORM DESIGN and OTHER ISSUES

The type of operation for which an airplane is intended has a very important bearing on the selection of the shape and design of the wing for that airplane. Planform design and other issues are discussed in this section. Planform determines the stall characteristics of the wing. Angle of incidence of the wing improves flight visibility, enhances take-off and landing characteristics and reduces drag in level flight. Wing washout helps to control the aircraft near its stall angle, the angle at which the lift coefficient of the wing drops drastically.

PLANFORM
Planform refers to the shape of the wing as seen from directly above. Wings may be rectangular or elliptical or delta shaped. Some wings taper from wing root to wing tip, with the taper along the leading edge or along the trailing edge or, in some cases, with a taper along both edges. The aspect ratio of a wing is the relationship between the length or span of the wing and its width or chord. It is computed by dividing the span by the average chord. A wing, for example, that has a span of 24 feet and a chord of 6 feet has an aspect ratio of 4. A wing with a span of 36 feet and a chord of 4 feet has an aspect ratio of 9. The actual size, in area, of both wings is identical (144 sq. ft.) but their flight performance is quite different because of their differing aspect ratios. A wing with a high aspect ratio will generate more lift and less induced drag than a wing with a low aspect ratio. For this reason, gliders have wings with high aspect ratios.

ANGLE OF INCIDENCE
The angle of incidence is the angle at which the wing is permanently inclined to the airplanes longitudinal axis. Choosing the right angle of incidence can improve flight visibility, enhance take-off and landing characteristics and reduce drag in level flight.

The angle of incidence that is usually chosen is the angle of attack at which the lift-drag ratio is optimum. In most modern airplanes, there is a small positive angle of incidence so that the wing has a slight angle of attack when the airplane is in level cruising flight.

WASHOUT
To reduce the tendency of the wing to stall suddenly as the stalling angle is approached, designers incorporate in wing design a feature known as washout. The wing is twisted so that the angle of incidence at the wing tip is less than that at the root of the wing. As a result, the wing has better stall characteristics, in that the section towards the root will stall before the outer section of the wing. The ailerons, located towards the wing tips, are still effective even though part of the wing has stalled. The same improved stall characteristics are achieved by the device of changing the airfoil shape from the root to the tip. The manufacturer incorporates a wing shape at the tip, which has the characteristic of stalling at a slightly higher angle of attack.

Wing Design - Level 3


OTHER WING ADDITIONS

The type of operation for which an airplane is intended has a very important bearing on the selection of the shape and design of the wing for that airplane. Wing fences, slots, slats,

spoilers, speed brakes and flaps are additions to the wing that perform a variety of functions related to control of the boundary layer, increase of the planform area (thus affecting lift and drag) and reduction of aircraft velocity during landing and stopping.

WING FENCES
Wing fences are fin-like vertical surfaces attached to the upper surface of the wing, that are used to control the airflow. On swept wing airplanes, they are located about two-thirds of the way out towards the wing tip and prevent the drifting of air toward the tip of the wing at high angles of attack. On straight wing airplanes, they control the airflow in the flap area. In both cases, they give better slow speed handling and stall characteristics.

SLOTS AND SLATS


Slats are auxiliary airfoils fitted to the leading edge of the wing. At high angles of attack, they automatically move out ahead of the wing. The angle of attack of the slat being less than that of the mainplane, there is a smooth airflow over the slat which tends to smooth out the eddies forming over the wing. Slats are usually fitted to the leading edge near the wing tips to improve lateral control. Slots are passageways built into the wing a short distance from the leading edge in such a way that, at high angles of attack, the air flows through the slot and over the wing, tending to smooth out the turbulence due to eddies.

SPOILERS
Spoilers are devices fitted to the wing which increase drag and decrease lift. They usually consist of a long narrow strip of metal arranged spanwise along the top surface of the airfoil. In some airplanes, they are linked to the ailerons and work in unison with the ailerons for lateral control. As such, they open on the side of the upgoing aileron, spoil the lift on that wing and help drive the wing down and help the airplane to roll into a turn. In some airplanes, spoilers have replaced ailerons as a means of roll control. The spoiler moves only upward in contrast to the aileron that moves upward to decrease lift and downward to increase lift. The spoiler moves only up, spoiling the wing lift. By using spoilers for roll control, full span flaps can be used to increase low speed lift.

Spoilers can also be connected to the brake controls and. when so fitted, work symmetrically across the airplane for producing drag and destroying lift after landing, thereby transferring all the weight of the airplane to the wheels and making braking action more effective.

SPEED BRAKES
Speed brakes are a feature on some high performance airplanes. They are a device designed to facilitate optimum descent without decreasing power enough to shock cool the engine and are especially advantageous in airplanes with high service ceilings. They are also of use in setting up the right approach speed and descent pattern in the landing configuration. The brakes, when extended, create drag without altering the curvature of the wing and are usually fitted far enough back along the chord so as not to disrupt too much lift and in a position laterally where they will not disturb the airflow over the tailplane. They are usually small metal blades housed in a fitting concealed in the wing that, when activated from the cockpit, pivot up to form a plate. On some types of aircraft, speed brakes are incorporated into the rear fuselage and consist of two hinged doors that open into the slipstream.

FLAPS
Flaps are high lift devices which, in effect, increase the camber of the wing and, in some cases, as with the Fowler Flap, also increase the effective wing area. Their use gives better take-off performance and permits steeper approach angles and lower approach and landing speeds. When deflected, flaps increase the upper camber of the wing, increasing the negative pressure on the top of the wing. At the same time, they allow a build up of pressure below the wing. During take-off, flap settings of 10 degrees to 20 degrees are used to give better take-off performance and a better angle of climb, especially valuable when climbing out over obstacles. However, not all airplane manufacturers recommend the use of flaps during take-off. They can be used only on those airplanes, which have sufficient take-off power to overcome the extra drag that extended flaps produce. The recommendations of the manufacturer should, therefore, always be followed. Flaps do indeed increase drag. The greater the flap deflection. the greater the drag. At a point of about half of their full travel, the increased drag surpasses the increased lift and the flaps become air brakes. Most flaps can be extended to 40 degrees from the chord of the wing. At settings between 20 degrees and 40 degrees, the essential function of

the flaps is to improve the landing capabilities, by steepening the glide without increasing the glide speed. In an approach over obstacles, the use of flaps permits the pilot to touch down much nearer the threshold of the runway. Flaps also permit a slower landing speed and act as air brakes when the airplane is rolling to a stop after landing, thus reducing the need for excessive braking action. As a result, there is less wear on the undercarriage, wheels and tires. Lower landing speeds also reduce the possibility of ground looping during the landing roll. Plain and split flaps increase the lift of a wing, but at the same time, they greatly increase the drag. For all practical purposes, they are of value only in approach and landing. They should not normally be employed for take-off because the extra drag reduces acceleration. Slotted flaps, on the other hand, including such types as Fowler and Zap, produce lift in excess of drag and their partial use is therefore recommended for take-off. From the standpoint of aerodynamic efficiency, the Fowler Flap is generally considered to offer the most advantages and the fewest disadvantages, especially on larger airplanes, while double slotted flaps have won wide approval for smaller types. On STOL airplanes, a combination of double slotted flaps and leading edge slats are common. Changes in flap setting affect the trim of an airplane. As flaps are lowered, the center of pressure moves rearward creating a nose down, pitching moment. However, in some airplanes, the change in airflow over the tailplane as flaps are lowered, is such that the total moment created is nose up and it becomes necessary to trim the airplane "nose down". The airplane is apt to lose considerable height when the flaps are raised. At low altitudes, therefore, the flaps should be raised cautiously. Most airplanes are placarded to show a maximum speed above which the flaps must not be lowered. The flaps are not designed to withstand the loads imposed by high speeds. Structural failure may result from severe strain if the flaps are selected "down" at higher than the specified speed. When the flaps have been lowered for a landing, they should not ordinarily be raised until the airplane is on the ground. If a landing has been missed, the flaps should not be raised until the power has been applied and the airplane has regained normal climbing speed. It is then advisable to raise the flaps in stages. How much flap should be used in landing? Generally speaking, an airplane should be landed as slowly as is consistent with safety. This usually calls for the use of full flaps. The use of flaps affects the wing airfoil in two ways. Both lift and drag are increased. The Increased lift results in a lower stalling speed and permits a lower touchdown speed. The increased drag permits a steeper approach angle without increasing airspeed. The extra drag of full flaps results in a shorter landing roll.

An airplane that lands at 50 knots with full flaps selected may have a landing speed as fast as 70 knots with flaps up. If a swerve occurs during the landing roll, the centrifugal force unleashed at 70 knots is twice what it would be at 50 knots, since centrifugal force increases as the square of the speed. It follows then, that a slower landing speed reduces the potential for loss of control during the landing roll. It also means less strain on the tires, brakes and landing gear and reduces fatigue on the airframe structure. There are, of course, factors, which at times call for variance from the procedure of using full flaps on landing. These factors would include the airplane's all-up-weight, the position of the C.G., the approach path to landing, the desired rate of descent and any unfavourable wind conditions, such as a strong cross wind component, gusty winds and extreme turbulence. With experience, a pilot learns to assess these various factors as a guide to flap selection. In some airplanes, in a crosswind condition, the use of full flap may be inadvisable. Flaps present a greater surface for the wind to act upon when the airplane is rolling on the ground. The wing on the side from which the wind is blowing will tend to rise. In addition, cross wind acting on full flaps increases the weather vaning tendencies, although in an airplane with very effective rudder control even at slow speeds, the problem is not so severe. However, in many airplanes, the selection of full flaps deflects the airflow from passing over the empennage, making the elevator and rudder surfaces ineffective. Positive control of the airplane on the ground is greatly hampered. Since maintaining control of the airplane throughout the landing roll is of utmost importance, it may be advisable to use less flaps in cross wind conditions. In any case, it is very important to maintain the crosswind correction throughout the landing roll.

Wing Design - Level 3


WING BOUNDARY LAYER

The type of operation for which an airplane is intended has a very important bearing on the selection of the shape and design of the wing for that airplane. Boundary layer effects play a very important part in determining the drag for the aircraft. Thus, the wing should be designed to minimize the drag.

THE BOUNDARY LAYER


The boundary layer is a very thin layer of air lying over the surface of the wing (and, for that matter, all other surfaces of the airplane). Because air has viscosity, this layer of air tends to adhere to the wing. As the wing moves forward through the air, the boundary layer at first flows smoothly over the streamlined shape of the airfoil. Here the flow is called the laminar layer.

As the boundary layer approaches the center of the wing, it begins to lose speed due to skin friction and it becomes thicker and turbulent. Here it is called the turbulent layer. The point at which the boundary layer changes from laminar to turbulent is called the transition point (Fig. 3). Where the boundary layer becomes turbulent, drag due to skin friction is relatively high. As speed increases, the transition point tends to move forward. As the angle of attack increases, the transition point also tends to move forward.

Fig. 3 Various methods have been developed to control the boundary layer in order to reduce skin friction drag. Suction Method. One method uses a series of thin slots in the wing running out from the wing root towards the tip. A vacuum sucks the air down through the slots, preventing the airflow from breaking away from the wing and forcing it to follow the curvature of the wing surface. The air, which is sucked in, siphons through the ducts inside the wing and is exhausted backwards to provide a little extra thrust. The laminar flow airfoil is itself a structural design intended to make possible better boundary layer control. The thickest part of a laminar flow wing occurs at 50% chord. The transition point at which the laminar flow of air breaks down into turbulence is at or near the thickest part. The transition point at which the laminar flow of air becomes turbulent on a laminar flow airfoil is rearward of that same point on a conventional designed airfoil.

Vortex generators are small plates about an inch deep standing on edge in a row spanwise along the wing. They are placed at an angle of attack and (like a wing airfoil section) generate vortices. These tend to prevent or delay the breakaway of the boundary layer by re-energizing it. They are lighter and simpler than the suction boundary layer control system described above.

Aeronautics - Fluid Dynamics - Level 3 Flow Equations


Equations Describing Fluid Flow
The flow of most fluids may be analyzed mathematically by the use of two equations. The first, often referred to as the Continuity Equation, requires that the mass of fluid entering a fixed control volume either leaves that volume or accumulates within it. It is thus a "mass balance" requirement posed in mathematical form, and is a scalar equation. The other governing equation is the Momentum Equation, or Navier-Stokes Equation, and may be thought of as a "momentum balance." As will be seen later, the Navier-Stokes equations are the fluid dynamic equivalent of Newtons second law, force equals mass times acceleration. The Navier-Stokes equations are vector equations, meaning that there is a separate equation for each of the coordinate directions (usually three).

There are many methods to derive these equations. One of the simplest, a control volume approach, is used here to demonstrate the origin of each term. These equations may be used to analyze the flow of most common fluids in internal (pipes) or external (wings) flow situations. Mathematically speaking, these equations are extremely difficult to solve in their raw form. The Navier-Stokes equations are second order, non-homogenous, non-linear partial differential equations that require at least two boundary conditions for solution. Most solutions that exist are for highly simplified flow situations where certain terms in the equations have been eliminated through some rational process.

Derivation of the Continuity Equation


Lets start with a small, fixed volume of fluid somewhere in the middle of a flow stream. This elemental volume has sides of lengths (x, (y and (z (see Figure 1).

Figure 1. Illustration of the elemental volume used to derive the equations.

These lengths are short enough so that changes in all fluid properties across the volume may be well approximated with linear functions. On the other hand, these dimensions must me large enough so that the fluid may be considered as a continuum (i.e., much larger than the molecular scale). The mass of fluid in this elemental volume depends on the amount of fluid entering and leaving through the faces. The difference between these two is the rate of mass that accumulates in the volume. The rate of mass entering a face is the product of the density, the fluid velocity and the face area. For example, on the side facing the reader, the density (V) is multiplied by the velocity in the x direction (u) and the area of the face (y (z. Thus, the mass flux entering the volume through this face is

The mass leaving the volume on the opposite side of the volume is again the product of density, velocity and area, but the density and velocity may have changed as the fluid passed through the volume. We will express these changes as small quantities (since our volume is small enough), i.e., V+ (V and u + (u. The mass flux leaving that face is thus

The negative sign tells us that the mass is leaving, rather than entering, the control volume. Performing the same analysis on the mass entering the volume through the other faces of the volume gives us

where v is the y direction velocity and w is the z direction velocity. Similarly, the mass fluxes leaving the volume on the opposite faces are

All of these added together must equal the mass of fluid accumulating in the volume,

Putting all of these together, we have

Multiplying out the quantities in parentheses results in the cancellation of some terms and the appearance of higher-order terms such as (V (u (y (z. Since the quantities preceded by ( are very small, products of these quantities will be extremely small, depending on the number of ( terms included in the product. The terms with four of these will be much smaller than the terms with only three ( terms. Thus, all higher order terms are neglected. This leaves

which, when divided by

and rearranged, yields

The application of basic calculus (taking the limit as (t tends to 0) allows us to write this equation as

where the symbol / t, for example, is a "partial derivative" with respect to time. Partial derivatives are used when the function (velocity or density in this case) depends on several variables (3 position or spatial variables and time, in this case). The Continuity Equation may be simplified for some common flow situations as follows. If the fluid may be treated as incompressible (as is the case with water or in low velocity air flows), the density will be constant. The Continuity Equation then becomes

In the case when the flow is steady (all time derivatives are zero), then

Note that in this equation, the density and velocities are still functions of the spatial coordinates x,y and z.

Derivation of the Momentum (Navier Stokes) Equations


Again we start with a small, fixed volume of fluid somewhere in the middle of a flow stream with sides of lengths (x, (y and (z (see Figure 1). The law of the conservation of momentum states that the rate of change of momentum in the control volume must equal the net momentum flux into the control volume plus any external forces acting on the control volume. We will first deal with the momentum change and flux, then with the external forces. Recall that momentum, the product of mass and velocity, is a vector quantity. This derivation will be based on the momentum in the x direction in Cartesian coordinates. Similar derivations may be demonstrated for the y and z direction. This would make a good exercise to better understand this material. There are thus three different momentum equations that together comprise the Navier-Stokes Equations.

Momentum Change and Flux

The time rate of change of momentum within the control volume is

where / t is the partial derivative operator with respect to time presen

where (A is the surface area. For the side facing the reader, the momentum flux is thus

The momentum flux out the opposite side is

The mass flux into the face with normal vector in the negative y direction is as derived for the continuity equation above, or Vv. The momentum flux into this face is thus

For the face opposite, the momentum flux out is

Using a similar analysis, it is easily shown that the momentum flux into and out of the faces with normal vectors in the z direction are

Our first expression of the momentum equation comes from adding all of these terms together as expressed in the law of the conservation of momentum

where 7Fx is the sum of the external forces on the control volume. The momentum fluxes into the control volume cancel with the first part of the momentum fluxes out of the control volume. Performing this cancellation and moving the momentum fluxes to the left hand side of the equation gives

Using the product rule, the momentum change and fluxes can be expanded to

where it is noted that the last four terms in parentheses are the continuity equation times u. Since this must be zero, that leaves

In the same method we can calculate the force in the y and z directions:

Derivation of Forces

We now turn our attention to the right hand side of the equation, where the forces on the control volume are represented. There are two types of forces to be included: body forces and surface forces. Body forces act on the entire control volume. The most common body force is that due to gravity. Electromagnetic phenomena may also create body forces, but this is a rather specialized situation. The body force due to gravity is the component of the acceleration due to gravity in the x-direction (gx) times the mass of the fluid control volume (density times volume), or <="" p="" height="15" width="70"> Surface forces act on only one particular surface of the control volume at a time, and arise due to pressure or viscous stresses. The stress on a surface of the control volume acts in the outward direction, and is given the symbol ij, with two subscripts. The first subscript i indicates the normal direction of the face on which the stress acts, while the second subscript j indicates the

direction of the stress. For example, using the cube above, the x axis is the normal direction to the back y-z face of the cube, the y axis is the normal direction to the left x-z face of the cube and the z axis is the normal direction to the top x-y face of the cube. Also, the -x axis is the normal direction to the front y-z face of the cube, the -y axis is the normal direction to the right x-z face of the cube and the -z axis is the normal direction to the bottom x-y face of the cube. The force due to the stress is the product of the stress and the area over which it acts. Thus, on the faces with normals in the x-direction ((y(z), the forces acting in the x-direction due to the direct stresses are <="" p="" height="33" width="227"> The sum of these two forces is

Similarly, on the faces with normals in the y-direction ((x(z), the forces in the x-direction due to shear stresses sum to

and on the faces with normals in the z-direction ((x(y), the forces in the x-direction due to shear stresses sum to

The sum of all surface forces in the x-direction is thus

The stress Wxx includes the pressure p (acting inward, and, hence, has a negative sign) and the normal viscous stress Xxx. The stresses Wyx and Wzx include only viscous shearing stresses Xyx and Xzx. This gives the force in the x-direction as:

Newtonian/Non-Newtonian Fluids Most fluids may be classified as Newtonian fluids. A Newtonian fluid is one in which the viscous stress is linearly proportional to the rate of deformation ( b du/dy). The constant of proportionality is the viscosity, . Air would be considered a low viscosity Newtonian fluid, while water would be a medium viscosity Newtonian fluid. Motor oil and Maple Syrup are high viscosity Newtonian fluids. Fluids that do not follow the Newtonian behavior law include toothpaste, blood and paints. For an incompressible Newtonian fluid the viscous stresses are

Some of these terms can be cancelled out using the continuity equation. The remaining terms, combined with the body force term and put into the equation for the force in the x-direction, give

This gives, as the final expression of the x-momentum equation,

Recall that there are corresponding momentum equations in the y and z directions, namely

and

The derivation of these equations would be a good exercise for the viewer.

Derivation of the Energy Equation


In situations where the fluid may be treated as incompressible and temperature differences are small, the continuity and momentum equations are sufficient to specify the velocities and pressure (that is, four equations [continuity+3 momentum] and four unknown quantities [u,v,w and p]). If the flow is compressible (V is not constant), or if heat flux occurs (temperature not constant), at least one additional equation is required. In some of these instances, the energy equation may be used. In the derivation, we use the fact that work is the dot product of velocity and force and the fact that work and energy are related to each other. The energy equation is, of course, a scalar equation, meaning that it has no particular direction associated with it. The procedure for deriving the energy equation is similar to those presented for the continuity and momentum equations. In this case, the change in energy of the fluid within the control volume is equal to the net thermal energy transferred into the control volume plus the rate of work done by external forces. The energy of the fluid is expressed in this case as the sum of the absolute thermodynamic internal energy per unit mass, e, and the kinetic energy per unit mass, 1/2 V2, (V is the magnitude of the velocity vector). The change in total energy per unit volume of the fluid in the control volume is

As was found above for the momentum transfer into and out of the control volume, the net transfer of energy per unit volume through the control volume is

This equation is obtained by replacing the momentum term (density times velocity) by the energy term (density times the sum of the internal and kinetic energies). The net thermal energy transferred into the control volume is determined by the heat flux qi , positive for heat going from within the control volume to the surroundings in the ith-direction (that is, the x-,y- or z-direction). The total heat per unit volume transferred into the control volume is

The rate of work per unit volume being done by the surface forces is found by multiplying the stress, Wij, by the velocity in the j-direction for each i face. Similar to the procedure above for the stresses in the momentum equation, the net rate of work being done from all sides is

Lastly, the rate of work per unit volume done by the gravity force vector (g = gx i + gy j + gz k ), is

Putting all of these terms together, we have

Fourier's Heat Conduction

We will now use Fourier's Law of Heat Conduction that relates the heat flow in the ith direction, qi, to the rate of change of temperature in the ith direction, namely, qi = -kiA T/ xi x,y,z directions where ki is the heat conduction coefficient in the ith-direction, A represents the surface area perpendicular to the ith-direction, and T represents the temperature of the flow. From the Zeroth Law of Thermodynamics, heat flows from a location of higher temperature to that of a lower temperature. If, for example, this is in the x-direction, then T/ x is negative. But since heat flow is considered positive when flowing from the control volume to the surroundings (meaning, in this case, in the positive x-direction), then for qx to be positive, we need the minus sign as indicated. Thus the heat flow rate per unit volume terms i=1,2,3 represents the

become (kx T/ x)/ x + (ky T/ y)/ y + (kz T/ z)/ z which, in the case of constant heat conduction coefficient (kx = ky = kz = k), become

<="" font="" height="40" width="110">

The rate of work per unit volume being done by the surface forces is found by multiplying the stress Wij by the velocity in the j direction for each i face. Similar to the procedure above for the

stresses in the momentum equation, the net rate of work per unit volume being done from all sides is

Using the relationship between surface stresses and velocity gradients for incompressible Newtonian fluids, this becomes

Lastly, the rate of work done per unit volume by gravity is

Putting all of these terms together, we have

This equation demonstrates that, per unit volume, the change in energy of the fluid moving through a control volume is equal to the rate of heat transferred into the control volume plus the rate of work done by surface forces plus the rate of work done by gravity. This expression of the energy equation is valid for most applications. However, some specialized situations may require additional terms representing the contributions of other sources (electromagnetic forces, etc.).

Aeronautics - Fluid Dynamics - Level 3

Incompressible flow around a cylinder and a wing section


Netscape users: Please be advised that this page contains JAVA script to show the animations. In order to view the animations effectively, please move your mouse over the "hot" words embedded in the text but do NOT click on them, otherwise it will take you back to our homepage. If that happens, use the "BACK" button to return to this page.
Incompressible flow around a stationary cylinder:

Inviscid (ideal) flow:

Viscous flow:

Inviscid (Ideal) flow around a stationary cylinder:


The impingement of flow on the cylinder creates a stagnation point on the approaching surface. The departure of the flow away from the cylinder creates another stagnation point on the trailing surface.

In the idealized situation where viscosity is neglected, the no-slip condition at the surface of the cylinder does not apply. Also in the absence of vorticity (inviscid flow) flow separation cannot occur.

Viscous flow around a stationary cylinder:


One stagnation point is created in front of the cylinder.

Because of the viscosity, a no-slip condition exists everywhere on the surface of the cylinder, i.e., the velocity must vanish everywhere on the surface. Consequently a boundary layer is created where the velocity transitions from a value of zero at the surface to the free stream value some distance away from the cylinder surface. The inertia of the fluid as it rounds the top and bottom of the trailing surface causes the flow to separate at these locations. This creates a disturbed wake (VonKarman vortex street) downstream from the cylinder.

Incompressible flow around a stationary wing section:

Inviscid (ideal) flow:

Viscous flow:

Inviscid (Ideal) flow around a stationary wing section:


As with the cylinder, stagnation points are created by the impingement of the flow on the approaching surface and the departure of the flow on the trailing surface.

In the idealized situation where viscosity is neglected, the no-slip condition at the surface of the wing section does not apply. Also in the absence of vorticity (inviscid flow) flow separation cannot occur.

Viscous flow around a stationary wing section:


One stagnation point is created in the wing section's leading edge.

Because of the viscosity a no-slip condition exists everywhere on the surface of the profile i.e. the velocity must vanish everywhere on the surface.

Consequently a boundary layer is created. The inertia of the fluid as it rounds the top and bottom of the trailing surface causes the flow to separate at these locations. This creates a disturbed wake that induces vortices downstream from the wing section.

Aeronautics - Fluid Dynamics - Level 3 FoilSim


Basic Aerodynamics Software is educational software designed at the NASA Lewis Research Center to instruct students in the basics of aerodynamics. The software contains two main parts, a baseball pitch and a wing-airflow simulator. The software also includes lessons which prompt students to engage in problem solving and discovery. The software was created to satisfy an objective to cultivate a more thorough understanding of the research being done at the NASA Lewis Research Center, while also filling a critical need for additional intuitive tools that supplement and enhance math and science curricula. This program is an interactive flow simulator determines the airflow around various basic shapes of airfoils, including a flat plate. Several lessons on basic aerodynamics are included with this simulator package. The lessons are displayed in a separate window while the program is operating.

Here's a screen shot from the program showing the controls and displays:

The program will run under Windows 3.x, 95 or Windows NT. Click here to download a copy of FoilSim {foilsim.zip} {1.09 Mb} in zip format. To install, 1. 2. 3. 4. Use PKUNZIP to expand this zip file to a temporary directory Run the program "setup.exe" and follow the instructions You can delete the files in the temporary directory when the installation is complete Once installed, Double click on the FoilSim icon to start the program

If you don't have PKUNZIP, you can download this self-extracting file of FoilSim {foilsimzp.exe} {1.10 Mb} 1. Run "foilsimzp.exe" and put the files in a TEMPORARY directory (as prompted, e.g. C:\TEMP) 2. Change directory to this temporary directory

3. Run the program "setup.exe" and follow the instructions to install FoilSim in the final directory (e.g., C:/FOILSIM) 4. Note: You can delete the files in the temporary directory when the installation is complete 5. Once installed, Double click on the FoilSim icon or the "foilsim.exe" file to start the program

The Axes of an Airplane - Level 3


There are three axes around which the airplane moves. These axes all pass through the airplane's center of gravity, which is that point which is the center of the airplane's total weight. The longitudinal axis extends lengthwise through the fuselage from the nose to the tail. Movement of the airplane around the longitudinal axis is known as roll and is controlled by movement of the ailerons. To move the ailerons, the pilot turns the control wheel either clockwise or counter clockwise (or moves the control stick either right or left). This action lowers the aileron on one wing and raises the aileron on the other wing. The downgoing aileron increases the camber of its wing, producing more lift and the wing rises. The upgoing aileron spoils the airflow on its wing, decreases the lift and the wing descends. The airplane rolls into a turn.

The lateral axis extends crosswise from wingtip to wing tip. Movement of the airplane around the lateral axis is known as pitch and is controlled by movement of the elevators. To effect a nose down attitude, the pilot pushes forward on the control wheel or stick. The elevator deflects downward, increasing the camber of the horizontal tail surface and thereby increasing the lift on the tail. To effect a nose up attitude of the airplane, the pilot pulls the wheel toward him. The

elevators are deflected upwards decreasing the lift on the tail, with a resultant downward movement of the tail. The vertical or normal axis passes vertically through the center of gravity. Movement of the airplane around the vertical axis is yaw and is controlled by movement of the rudder. Pressure applied to the left rudder pedal, for example deflects the rudder to the left into the airflow. The pressure of the airflow against the rudder pushes the tail to the right. The nose of the airplane yaws to the left. There is a distinct relationship between movement around the vertical and longitudinal axes of an airplane (i.e. yaw and roll). When rudder is applied to effect a yaw, for example, to the right, the left wing (on the outside of the turn) moves faster than the inside wing, meets the relative airflow at a greater angle of attack and at greater speed and produces more lift.. The use of rudder, therefore, along with aileron can help to raise the wing and produce a better coordinated turn. In a roll, the airplane has a tendency to yaw away from the intended direction of the turn. This tendency is the result of aileron drag and is called adverse yaw. The upgoing wing, as well as gaining more lift from the increased camber of the downgoing aileron, also experiences more induced drag. The airplane, as a result, skids outward on the turn. Use of rudder in the turn corrects this tendency.

Press YAW

to see Yaw Animation

Rudder rotates the airplane around vertical axis.

Press ROLL

to see Roll Animation

Ailerons rotate the airplane around longitudinal axis.

Press PITCH

to see Pitch Animation

Elevators rotate airplane around lateral axis.

BALANCED CONTROLS - Level 3

Controls are sometimes dynamically balanced to assist the pilot to move them. By having some of the control surface in front of the hinge, the air striking the forward portion helps to move the control surface in the required direction. The design also helps to counteract adverse yaw when used in aileron design. Control surfaces are sometimes balanced by fitting a mass (usually of lead) of streamline shape in front of the hinge of the control surface. This is called mass balance and is incorporated to prevent flutter of the control surface, which is liable to occur at high speeds. The exact distribution of weight on a control surface is very important. For this reason, when a control surface is repainted, repaired or component parts replaced, it is essential to check for proper balance and have it rebalanced if necessary. To do this, the control surface is removed, placed in a jig and the position of the center of gravity checked against the manufacturer's specifications. Without any airflow over the control surface, it must balance about its specified C.G. This is known as static balance. For example, the aileron of the Bonanza is designed for a static nose heavy balance of 0.2 inch pounds. The C.G. of the aileron is forward of the hinge centerline causing the control surface to be nose heavy.

Stability of an Airplane - Level 3


An airplane in flight is constantly subjected to forces that disturb it from its normal horizontal flight path. Rising columns of hot air, down drafts gusty winds, etc., make the air bumpy and the airplane is thrown off its course. Its nose or tail drops or one wing dips. How the airplane reacts to such a disturbance from its flight attitude depends on its stability characteristics. Stability is the tendency of an airplane in flight to remain in straight, level, upright flight and to return to this attitude, if displaced, without corrective action by the pilot. Static stability is the initial tendency of an airplane, when disturbed, to return to the original position. Dynamic stability is the overall tendency of an airplane to return to its original position, following a series of damped out oscillations. Stability may be (a) positive, meaning the airplane will develop forces or moments which tend to restore it to its original position; (b) neutral, meaning the restoring forces are absent and the airplane will neither return from its disturbed position, nor move further away; (c) negative, meaning it will develop forces or moments which tend to move it further away. Negative stability is, in other words, the condition of instability.

A stable airplane is one that will fly "hands off" and is pleasant and easy to handle. An exceedingly stable airplane, on the other hand, may lack maneuverability. An airplane which, following a disturbance, oscillates with increasing up and down movements until it eventually stalls or enters a dangerous dive would be said to be unstable, or to have negative dynamic stability. An airplane that has positive dynamic stability does not automatically have positive static stability. The designers may have elected to build in, for example, negative static stability and positive dynamic stability in order to achieve their objective in maneuverability. In other words, negative and positive dynamic and static stability may be incorporated in any combination in any particular design of airplane. An airplane may be inherently stable, that is, stable due to features incorporated in the design, but may become unstable due to changes in the position of the center of gravity (caused by consumption of fuel, improper disposition of the disposable load, etc.). Stability may be (a) longitudinal, (b) lateral, or (c) directional, depending on whether the disturbance has affected the airframe in the (a) pitching, (b) rolling, or (c) yawing plane.

LONGITUDINAL STABILITY
Longitudinal stability is pitch stability, or stability around the lateral axis of the airplane. To obtain longitudinal stability, airplanes are designed to be nose heavy when correctly loaded. The center of gravity is ahead of the center of pressure. This design feature is incorporated so that, in the event of engine failure, the airplane will assume a normal glide. It is because of this nose heavy characteristic that the airplane requires a tailplane. Its function is to resist this diving tendency. The tailplane is set at an angle of incidence that produces a negative lift and thereby, in effect, holds the tail down. In level, trimmed flight, the nose heavy tendency and the negative lift of the tailplane exactly balance each other. Two principal factors influence longitudinal stability: (1) size and position of the horizontal stabilizer, and (2) position of the center of gravity.
The Horizontal Stabilizer

The tail plane, or stabilizer, is placed on the tail end of a lever arm (the fuselage) to provide longitudinal stability. It may be quite small. However, being situated at the end of the lever arm, it has great leverage. When the angle of attack on the wings is increased by a disturbance, the center of pressure moves forward, tending to turn the nose of the airplane up and the tail down. The tailplane, moving down, meets the air at a greater angle of attack, obtains more lift and tends to restore the balance. On most airplanes, the stabilizer appears to be set at an angle of incidence that would produce an upward lift. It must, however, be remembered that the tailplane is in a position to be in the

downwash from the wings. The air that strikes the stabilizer has already passed over the wings and been deflected slightly downward. The angle of the downwash is about half the angle of attack of the main airfoils. The proper angle of incidence of the stabilizer therefore is very important in order for it to be effective in its function.
Center of Gravity

The center of gravity is very important in achieving longitudinal stability. If the airplane is loaded with the center of gravity too far aft, the airplane may assume a nose up rather than a nose down attitude. The inherent stability will be lacking and, even though down elevator may correct the situation, control of the airplane in the longitudinal plane will be difficult and perhaps, in extreme cases, impossible.

LATERAL STABILITY
Lateral stability is stability around the longitudinal axis, or roll stability. Lateral stability is achieved through (1) dihedral, (2) sweepback, (3) keel effect, and (4) proper distribution of weight.
Dihedral

The dihedral angle is the angle that each wing makes with the horizontal. The purpose of dihedral is to improve lateral stability. If a disturbance causes one wing to drop, the unbalanced force produces a sideslip in the direction of the downgoing wing. This will, in effect, cause a flow of air in the opposite direction to the slip. This flow of air will strike the lower wing at a greater angle of attack than it strikes the upper wing. The lower wing will thus receive more lift and the airplane will roll back into its proper position. Since dihedral inclines the wing to the horizontal, so too will the lift reaction of the wing be inclined from the vertical. Hence an excessive amount of dihedral will, in effect, reduce the lift force opposing weight. Some modern airplanes have a measure of negative dihedral or anhedral, on the wings and/or stabilizer. The incorporation of this feature provides some advantages in overall design in certain type of airplanes. However, it does have an effect, probably adverse, on lateral stability.
Keel Effect

Dihedral is more usually a feature on low wing airplanes although some dihedral may be incorporated in high wing airplanes as well. Most high wing airplanes are laterally stable simply because the wings are attached in a high position on the fuselage and because the weight is therefore low. When the airplane is disturbed and one wing dips, the weight acts as a pendulum returning the airplane to its original attitude.

Sweepback

A sweptback wing is one in which the leading edge slopes backward. When a disturbance causes an airplane with sweepback to slip or drop a wing, the low wing presents its leading edge at an angle that is perpendicular to the relative airflow. As a result, the low wing acquires more lift, rises and the airplane is restored to its original flight attitude. Sweepback also contributes to directional stability. When turbulence or rudder application causes the airplane to yaw to one side, the right wing presents a longer leading edge perpendicular to the relative airflow. The airspeed of the right wing increases and it acquires more drag than the left wing. The additional drag on the right wing pulls it back, yawing the airplane back to its original path.

DIRECTIONAL STABILITY
Directional stability is stability around the vertical or normal axis. The most important feature that affects directional stability is the vertical tail surface, that is, the fin and rudder. Keel effect and sweepback also contribute to directional stability to some degree.
The Fin

An airplane has the tendency always to fly head-on into the relative airflow. This tendency which might be described as weather vaning is directly attributable to the vertical tail fin and to some extent also the vertical side areas of the fuselage. If the airplane yaws away from its course, the airflow strikes the vertical tail surface from the side and forces it back to its original line of flight. In order for the tail surfaces to function properly in this weather vaning capacity, the side area of the airplane aft of the center of gravity must be greater than the side area of the airplane forward of the C.G. If it were otherwise, the airplane would tend to rotate about its vertical axis.

You might also like