You are on page 1of 57

Wings and Lift How does a wing work?

Any object, whether it be a book sitting on a table or a rocket blasting into space, is acted upon by some number of forces. The primary force most of us experience in our lives is our own weight, the force holding us to the surface of the Earth. An aircraft in flight is acted upon by four basic forces: weight that pulls the aircraft toward the ground, thrust that pushes the aircraft forward, drag or the force of the air pushing against the motion of the aircraft, and lift that pulls the aircraft up.

Four forces acting on an aircraft in flight It is this lift force that is produced (primarily) by the wing, and I believe what the questioner means to ask is "How does a wing generate lift?" Though this seems like a simple enough question, the general public would probably be amazed to find out that engineers and scientists still debate just how lift is produced even 100 years after flight became a reality. In fact, it is quite easy to be drawn into charged debates on the subject, as I was when trying to answer this question. So, to be fair to the proponents of each theory, I will discuss each in turn. But first, let us simplify our discussion slightly by thinking of the wing as only a twodimensional shape. Consider the cross-section of a wing created by a plane cutting through the wing. This two-dimensional cross-sectional shape is called an airfoil (or aerofoil to our British friends). An example of a common airfoil shape is the Clark Y.

Clark Y airfoil Bernoulli theory: The most common explanation of the concept of lift is based upon the Bernoulli equation, an equation that relates the pressures and velocites acting along the surface of a wing. What this equation says, in simple terms, is that the sum of the pressures acting on a body is a constant. This sum consists of two types of pressures: 1) the static pressure, or the atmospheric pressure at any point in a flowfield, and 2) the dynamic pressure, or the pressure created by the motion of a body through the air. Since dynamic pressure is a function of the velocity of the flow, the Bernoulli equation relates the sum of pressures to the velocity of the flow past the body. So what this equation tells us is that as velocity increases, pressure decreases and vice versa. To understand why the flow velocity changes, we introduce a second relation called the Continuity equation. What this relationship tells us is that the velocity at which a flow passes through an area is directly related to the size of that area. For example, if you blow through a straw, the air will come out at a certain speed. If you then blow in with the same strength but now squeeze the end of the straw, the air will come out faster. So how do these equations relate to our two-dimensional airfoil? Look again at the Clark Y and notice that an airfoil is a curved shape. While the bottom is relatively flat, the top surface is thicker and more curved. Thus, when air passes over an airfoil, that flow over the top is squeezed into a smaller area than that airflow passing the lower surface. The Continuity equation tells us that a flow squeezed into a smaller area must go faster, and the Bernoulli equation tells us that when a flow moves faster, it creates a lower pressure. Thus, a higher pressure exists on the lower surface of an airfoil and a lower pressure on the upper surface. Whenever such a pressure difference exists in nature, a force is created in the direction of the lower pressure (since pressure is defined as force per unit area). Think of it as the upper surface being sucked upward. This upward force, of course, is lift. It is this theory that appears in most aerodynamic textbooks, albeit sometimes with incorrect assumptions applied and conclusions drawn. Newtonian theory: A theory currently gaining in popularity and arguably more "fundamental" in origin is the Newtonian theory, so named because it is said to follow from Newton's third law of motion (for every action there is an equal and opposite reaction). First, one most realize that any airfoil generating lift deflects the air flow behind it. Positive lift deflects the air downward, towards the ground. Thus, the motion of any lifting surface through a flow accelerates that flow in a new direction. Newton's second law tells us that force is directly proportional to acceleration (F=ma). Therefore, we must conclude from Newton's third law that the force accelerating the air downward must be accompanied by an equal and opposite force pushing the airfoil upward. This upward force is lift. Circulation theory: The most mathematical explanation for lift is the circulation theory. Circulation can be thought of as a component of velocity that rotates or swirls around an airfoil or any other shape. In a branch of aerodynamics called incompressible flow, we can use potential flow relationships to solve for this circulation for a desired shape. Once this quantity is known, the force of lift can be solved for using the Kutta-Joukowski theorem that directly relates lift and circulation. This approach tends to be more mathematically intense than I wish to get into here, and it's really more of a method of calculating lift in an ideal flowfield than an explanation of the physical origins of lift.

Conclusion: So the reader may be asking which of these theories is correct? In truth, each is valid in some respect and useful for certain applications, but the ultimate question is which is the most fundamental explanation. Mathematicians would surely prefer the circulation theory, which is certainly a very elegant approach firmly based on mathematical principles, but it fails to explain what force of nature creates circulation or lift. Many would argue that the Newtonian explanation is most fundamental since it is "derived" from Newtonian laws of motion. While this is true to some degree, the theory lacks an explanation as to why an airfoil deflects the flow downward in the first place. Even accepting this principle, the idea that an airfoil deflects the flow and therefore experiences lift also fails to capture the fundamental tools of nature (pressure and friction) that create and exert that force on the body. Proponents of this explanation generally deride the Bernoulli theory because it relies on less fundamental concepts, like the Bernoulli and Continuity equations. There is some truth to this complaint, and the theory may be more difficult for the novice to understand as a result. However, both equations are derived from Newtonian physics, and I would argue from more fundamental and more mathematically sound premises than the Newtonian theory. In the end, I leave it up to the reader to decide. The following are some sources for further reading on the subject: ALLSTAR Principles of Flight -- A much simplified description of the Bernoulli theory Airfoil Misconception -- A site favoring the Newton theory A Physical Description of Flight -- The champions of the Newton theory Bernoulli vs. Newton -- Good discussion of the two approaches Remember that in answering this question, I considered only the two-dimensional crosssection of a wing, the airfoil, also known as an infinite wing. However, I did not adress any issues regarding the design of the airfoil itself, like camber, thickness, or symmetrical airfoils. In addition, this discussion hasn't even touched on the aerodynamic properties of a finite wing, or a wing with tips. As one might expect, such a wing has far different characteristics than does the infinite wing. A discussion of finite wings involves addressing concepts like induced drag, tip vortices, and aspect ratio as well as wing design issues like taper, sweep, twist, and high-lift devices like flaps and slats. Such a discussion goes far beyond the scope of this question, but we look forward to receiving questions about these topics in the future. - answer by Jeff Scott, 17 December 2000

How Airplanes Fly: A Physical Description of Lift


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 highspeed 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. Newtons 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 formally 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 2g 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: 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. The planes 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.

Chapter 1 - Structure of an Airplane


At the end of this block of study, you should be able to: Label the parts of an airplane. Describe the five types of stress which act on an aircraft in flight and give an example of where each applies to an airplane. Describe both truss and semimonocoque types of fuselages. Describe the basic structure of a wing. Explain the structure and function of the empennage. Identify the three types of landing gear.

Most aircraft are composed of the fuselage (body), wings, empennage (tail assembly), landing gear, and power plant (see figure 1-1). Locate these parts in the diagram as they are discussed. Sections in this Chapter: Section 1.1 - THE FUSELAGE STRUCTURE Section 1.2 - WINGS Section 1.3 - EMPENNAGE Section 1.4 - LANDING GEAR Section 1.5 - POWER PLANT Section 1.6 - REVIEW EXERCISE

Section 1.1 - The Fuselage Structure

The word fuselage is based on the French word fuseler, which means "to streamline." The fuselage must be strong and streamlined since it must withstand the forces that are created in flight. It houses the flight crew, passengers, and cargo. Fuselages are classified according to the arrangement of their force-resisting structure. The types of fuselages we will study are the truss and the semimonocoque. Five types of stress act on an aircraft in flight: tension, compression, bending, shear, and torsion. Let's look at each one individually

Tension:

Tension is the stress which tends to pull things apart. When you try to break a length of rope, you exert a type of stress which is called tension. (see animation or figure 1-2a)

Compression: Compression is the opposite of tension. It is the stress which tends to push materials together. When you grasp a football at both ends and push, the ball is subject to compression. The landing gear struts of an aircraft are also subject to compression. (see animation or figure 1-2b) Bending: This type of stress combines tension and compression. You put a bending stress on a bar when you grasp it with both hands and push the ends together or when you bend a paper clip. The wing spars (interior structural members) are subjected to bending while the aircraft is in flight. The lower side of the spar is subjected to tension, while the upper side is subjected to compression. Obviously, some materials will break before they bend and often are unacceptable for aircraft construction. (see animation or figure 1-2c which shows the upper side as the tensile side) Shear: Shear stress is caused by forces tending to slip or slide one part of a material in respect to another part. This is the stress that is placed on a piece of wood clamped in a vise and you Chip away at it with a hammer and chisel. This type of stress is also exerted when two pieces of metal, bolted together, are pulled apart by sliding one over the other or when you sharpen a pencil with a knife. The rivets in an aircraft are intended to carry only shear. Bolts, as a rule, carry only shear, but sometimes they carry both shear and tension. (see animation or figure 1-2d) Torsion: Torsion is the stress which tends to distort by twisting. You produce a torsional force when you tighten a nut on a bolt. The aircraft engine exerts a torsional force on the crankshaft or turbine axis. All the members (or major portions) of an aircraft are subjected to one or more of these stresses. Sometimes a member has alternate stresses, such as compression one instant and tension the next. Some members can carry only

one type of stress. Wire and cables, for example, normally carry only tension.(see animation or figure 1-2e) Since any member is stronger in compression or tension than in bending, members carry end loads better than side loads. In order to do this, designers arrange the members in the form of a truss, or rigid framework (see figure 1-3). In order for a truss to be rigid, it must be composed entirely of triangles. When the load on a truss acts in one direction, every alternate member carries tension while the other members carry compression. When the load is reversed, the members which were carrying compression now are subjected to tension and those which were carrying tension are under compression. The truss itself consists of a welded tubular steel structure with longerons (horizontal members) and diagonal braces. These features make it rigid, strong, and light. The truss is covered with a metal or fabric cover so that less drag will be generated. To produce a smooth surface, the fabric cover is put on fairing strips, which are thin flat strips of wood or metal. These fairing strips run the length of the fuselage in line with the direction of flight. The semimonocoque is the most often used construction for modern, high-performance aircraft. Semimonocoque literally means half a single shell. Here, internal braces as well as the skin itself carry the stress (see figure 1-4). The internal braces include longitudinal (lengthwise) members called stringers and vertical bulkhead. The semimonocoque structure is easier to streamline than the truss structure. Since the skin of the semimonocoque structure must carry much of the fuselage's strength, it will be thicker in some places than at other places. In other words, it will be thicker at those points where the stress on it is the greatest.

Section 1.2 - The Wings

Wing construction is basically the same in all types of aircraft. Most modern aircraft have all metal wings, but many older aircraft had wood and fabric wings. Ailerons and flaps will be studied later in this chapter.

To maintain its all-important aerodynamic shape, a wing must be designed and built to hold its shape even under extreme stress. Basically, the wing is a framework composed chiefly of spars, ribs, and (possibly) stringers (see figure 1-5). Spars are the main members of the wing. They extend lengthwise of the wing (crosswise of the fuselage). All the load carried by the wing is ultimately taken by the spars. In flight, the force of the air acts against the skin. From the skin, this force is transmitted to the ribs and then to the spars. Most wing structures have two spars, the front spar and the rear spar. The front spar is found near the leading edge while the rear spar is about two-thirds the distance to the trailing edge. Depending on the design of the flight loads, some of the all-metal wings have as many as five spars. In addition to the main spars, there is a short structural member which is called an aileron spar. The ribs are the parts of a wing which support the covering and provide the airfoil shape. These ribs are called forming ribs. and their primary purpose is to provide shape. Some may have an additional purpose of bearing flight stress, and these are called compression ribs. The most simple wing structures will be found on light civilian aircraft. High-stress types of military aircraft will have the most complex and strongest wing structure.

Three systems are used to determine how wings are attached to the aircraft fuselage depending on the strength of a wing's internal structure. The strongest wing structure is the full cantilever which is attached directly to the fuselage and does not have any type of external, stress-bearing structures. The semicantilever usually has one, or perhaps two, supporting wires or struts attached to each wing and the fuselage. The externally braced wing is typical of the biplane (two wings placed one above the other) with its struts and flying and landing wires (see figure 16).

Section 1.3 - Empennage

The empennage, commonly called the tail assembly (see figure 1-7), is the rear section of the body of the airplane. Its main purpose is to give stability to the aircraft. The fixed parts are the horizontal stabilizer and the vertical stabilizer or fin. The front, fixed section is called the horizontal stabilizer and is used to prevent the airplane from pitching up or down. The rear section is called the elevator and is usually hinged to the horizontal stabilizer. The elevator is a movable airfoil that controls the up-and-down motion of the aircraft's nose. The vertical tail structure is divided into the vertical stabilizer and the rudder. The front section is called the vertical stabilizer and is used to prevent the aircraft from yawing back and forth. The principle behind its operation is much like the principle of a deep keel on a sailboat. In light, single-engine aircraft, it also serves to offset the tendency of the aircraft to roll in the opposite direction in which the propeller is rotating. The rear section of the vertical structure is the rudder. It is a movable airfoil that is used to turn the aircraft.

Section 1.4 - Landing Gear


Airplanes require landing gear for taxiing, takeoff, and landing. The earliest airplane of all--the Wright Flyer--used skids as its landing gear. Soon, wheels were attached to the skids. Since that time, various arrangements have been used for wheels and structures to connect them to the airplane. Today, there are three common types of landing gear: conventional, tricycle, and tandem (see figure 1-8).

Conventional landing gear consists of two wheels forward of the aircraft's center of gravity and a third small wheel at the tail. This type of landing gear is most often seen in older general aviation airplanes. The two main wheels are fastened to the fuselage by struts. Without a wheel at the nose of the plane, it easily pitches over if brakes are applied too soon. Because the tailwheel is castered--free to move in any direction--the plane is very difficult to control when landing or taking off.

The tricycle landing gear, as you can guess from its name, has three wheels--two main wheels and a nosewheel (see figure 1-9). This type of landing gear makes the aircraft easier to handle on the ground and it also makes landings much safer. An aircraft equipped with tricycle landing gear is less apt to pitch forward. The tandem landing gear is used for very large aircraft like the B-52 bomber and the U-2 reconnaissance/research aircraft. The main landing gear is in two sets that are located one behind the other on the fuselage. The tandem landing gear allows the use of a highly flexible wing, but it may also require the use of small wheels on the tips of the wings to keep the wings from scraping the ground.

Section 1.6 - Review Exercise

1. Label the parts of the aircraft shown below.

A. C. E. G. I. K. M.

B. D. F. H. J. L.

2. Complete the chart of landing gear types by filling in the first column with the correct term: conventional, tricycle, or tandem. Use the description to guide you. A. B. C.

Easiest for pilot to make a safe landing. The landing gear is in two sets on the fuselage. Allows the use of highly flexible wing. Most difficult for pilot to control the plane. Used in earlier aircraft; would often pitch over.

3. Match the following statements with the BEST response. Choose the appropriate answer from the drop down menu. A. B. C. D. E. F. G. H. I. J.

Structure on a wing that supports the skin and provides the airfoil shape. The type of fuselage where the skin and internal braces carry the flight stresses. Movable section of the horizontal stabilizer that pitches nose up or down. Type of stress that pulls things apart. Vertical structure that prevents an airplane from rolling over. Movement up and down, back and forth. Type of stress which is caused by twisting. Structure bearing the major part of the wing's load. Body of a plane that must be strong and streamlined. Strongest type of system for attaching the wings to the fuselage.

Chapter 2-Characteristics of the Flight Atmosphere

At the end of this block of study, you should be able to: Describe the composition of the atmosphere. Describe why air has weight and how this is related to atmospheric pressure. Describe how changes in air temperature affect atmospheric pressure and density. Define relative humidity. Convert temperatures between the Fahrenheit and the Celsius scales. The atmosphere is composed of a mixture of gases. Nitrogen is by far the largest component of air, accounting for 78 percent; the next largest component is oxygen which consists of 21 percent. The remainder consists of small portions of other gases. All gases, regardless of where they are found, have certain characteristics such as weight, density, temperature, pressure, and mass. To understand how an airplane flies within the atmosphere, you must first understand some of the features of the atmosphere.

Sections in this Chapter:


Section 2.1 - WEIGHT Section 2.2 - DENSITY Section 2.3 - PRESSURE Section 2.4 - HUMIDITY Section 2.5 - TEMPERATURE Section 2.6 - REVIEW EXERCISE

Section 2.1 - WEIGHT

Because we cannot see air we may think of it as nothing, but you need only to stand in front of a fan or pedal your bicycle into the wind to know that air has substance. Air has weight because it is matter and is attracted by the Earth's gravity. Weight is explained in Figure 2-1a.

Notice that the direction of the atmosphere's weight is toward Earth's center, which is the center of gravity. Refer to the 1 inch x 1 inch x 60 miles column of air shown in Figure 2-1a. This column of air has a weight of 14.7 pounds. This is caused by gravity pulling all air molecules within the column toward Earth's center with that many pounds of force. It is the weight of the air which produces atmospheric pressure. Always remember that WEIGHT is in one direction, while PRESSURE is in all directions. Mass: Mass is the amount of material as shown in Figure 2-1b.

Section 2.2 - DENSITY

Density is illustrated in Figure 2-2a and is the amount of material contained in a unit of volume. Density is constant in solids because more material cannot be forced into a given volume of a solid. Solids will remain almost unchanged in density as long as the temperature doesn't become high enough to cause the solids to melt or burn. With air, however, the story is quite different.

Density: Amount of material PER UNIT VOLUME. Figure 2-2a shows two small one-inch cubes on the right side. One of them weighs only one pound, and the other, with tightly packed molecules, weighs ten pounds Although the volume is the same, the weight is very different because the molecules are packed closer together. The density of the air and the changes that temperature produce upon the density of air play an important role in an aircraft's flight.

Temperature: ENERGY of motion.


Since air is a gas, it is free to expand or contract as its temperature changes. Notice in Figure 2-2b how the five molecules of air increase their range of motion with each 20-degree increase in temperature. Since the molecules are not confined in a container, the air expands as the temperature increases. This also means that the range of particle motion decreases with decreases in temperature. Comparing the temperature and density of a unit of air, you can see that as temperature increases the density will decrease, and as the temperature decreases the density increases. As you will understand later, this constantly changing air density is very significant to flight.

Section 2.3 - PRESSURE

The atmosphere extends upward for hundreds of miles. The pull of the Earth's gravity on air molecules (see figure 2-3a) creates a pressure that pushes in all directions and amounts to about 14.7 pounds per square inch (psi) at sea level. This is air pressure on a standard day at sea level. This pressure will support 29.92 inches of mercury in a barometer, which is the instrument that provides the local atmospheric pressure that you hear on your daily television weather bulletins. As you know, pressure changes take place at Earth's surface as high- and low-pressure cells form and move across the surface. However, changes in atmospheric pressure occur for aviators not only as they fly into or out of high- and low-pressure cells but also as they climb or descend in their airplanes. This is because atmospheric pressure changes with altitude (higher altitude - less pressure, lower altitude - more pressure). Let us consider Earth's total atmosphere. We find that it extends from the surface outward into what we call space. As we travel outward (upward) there is less pressure because there are fewer air molecules above us. At 18,000 feet the pressure is about one-half as that at the surface. If we continue to travel outward, the pressure continues to reduce until there is no measurable pressure and, for all practical purposes, we are in space. Few aviators are concerned with high altitudes and consequent extreme low pressures, because the maximum operating altitude (ceiling) for most small air planes is 20,000 feet or less. Even at the relatively low altitude of 12,500 feet, pilots are required by regulation to breathe supplemental oxygen. The factors of weight, density, temperature, and pressure of the atmosphere interact and one does not change without affecting the others. However, temperature change is the main reason the atmospheric pressure and density change. Let's look at how the temperature causes these changes. To do this, we must use a starting point and the accepted starting point is known as the standard atmosphere. The standard atmosphere is based on average conditions at 40 north latitude where the average pressure is 29.92 inches of mercury and the average temperature is 59 Fahrenheit (F) or 15 Celsius (C). Under standard conditions, temperature decreases about 2 C or 3.5 F. for each 1,000foot increase in altitude. Variations to this standard are common and the standard decrease is not always found. For example, what is called a temperature inversion may actually cause an increase rather than a decrease in temperature at some locations. Section 2.4 - HUMIDITY The air contains water vapor in addition to the gases we have mentioned. This will vary from a very small amount up to a maximum of four or five percent. The air gets this moisture by evaporation from bodies of water and from vegetation. The capacity of air for holding water vapor is directly related to the temperature of the air; the warmer the air, the more water vapor it can hold. When the air contains all the vapor it can for its temperature and pressure, it is said to be saturated. Weather forecasts provide us with the relative humidity which is the ratio of amount of water vapor which a sample of air holds to the amount it can hold when saturated (see figure 2-5). It may be surprising to you that water vapor (composed mainly of hydrogen) is lighter than air (composed mainly of nitrogen). Thus, air which feels damp and heavy is actually lighter than dry air. Its density is also less.

Temperature: ENERGY of motion.


Temperature also changes horizontally as well as vertically in the atmosphere (see figure 2-2b). How much it changes in any direction is partially based on the extent to which Earth's surface area is exposed to the Sun, our primary source of heat. The Sun's rays penetrate the atmosphere and heat Earth's surface, which in turn heats the atmosphere next to it. Atmospheric temperature is based on the amount of heat absorbed or reflected by the Earth's surface. Sand, water, and forested surfaces differ in the amount of heat they absorb and reflect so the air over them will be heated differently.

Pressure: Omnidirectional. FORCE of motion PER UNIT AREA. The relationship of temperature to density and pressure can be confusing (see figure 2-3b). If a cubic foot of air on the Earth's surface is heated, the air will expand and rise in the atmosphere. This cubic foot of air may then occupy two cubic feet. Both its pressure and density will have decreased. As the air rises and expands, the heat it contains is spread over a larger area, and its temperature decreases with altitude. Also, as the air rises it receives less and less additional heat from the Earth's surface. Section 2.5 - TEMPERATURE We have mentioned Fahrenheit and Celsius temperatures. The different figures result merely from using different scales or units of measurement. The military services and scientists usually use the Celsius scale, while civilian aviation primarily uses the Fahrenheit scale. The Celsius scale runs from 0 for freezing to 100 for boiling water; these same events occur at 32 and 212 on the Fahrenheit scale. The average standard day temperature at sea level for most of the United States is 15 C and 59 F. To convert from one to the other, you can use the appropriate formula:

C = (F-32) x 5/9 F = (9/5 x C) + 32 For example: 15C = (59F - 32) x 5/9 or 59F=(9/5 x 15C)+32 Section 2.6 - REVIEW EXERCISE Choose the BEST option that completes the statement. 1. On the Celsius scale water boils at 2. 41 degrees F is equal to about

3. What percentage of the Earth's atmosphere is made up of oxygen? 4. As the altitude increases the atmospheric pressure

5. As the temperature of a parcel of air decreases, its density

6. As the temperature of a parcel of air increases, its pressure 7. Relative humidity is the measurement of the total percentage of water vapor in the air.

Chapter 3 - Principles of Flight - Level 2 At the end of this block of study, you should be able to:

Define airfoil, camber, and chord. Identify the parts of an airfoil. Describe Bernoulli's principle and tell how it relates to lift on an airfoil. Define relative wind, angle of incidence, and angle of attack. Aeronautics is the term applied to the flight of an aircraft through the atmosphere. As defined in Webster's New Collegiate Dictionary, aeronautics is "the science dealing with the operation of aircraft" or "the art or science of flight." We will begin the study of aeronautics in this section by discussing airfoils, relative wind, angle of attack, and the four forces of flight. these are the basics of aeronautics. Sections in this Chapter: Section 3.1 - AIRFOILS Section 3.2 - BERNOULLI'S PRINCIPLE Section 3.3 - RELATIVE WIND Section 3.4 - REVIEW EXERCISE

Section 3.1 - AIRFOILS An airfoil is any part of an airplane that is designed to produce lift. Those parts of the airplane specifically designed to produce lift include the wing and the tail surface. In modern aircraft, the designers usually provide an airfoil shape to even the fuselage. A fuselage may not produce much lift, and this lift may not be produced until the aircraft is flying relatively fast, but every bit of lift helps. Figure 3-1 shows a cross section of a wing, but it could be a tail surface or a propeller because they are all essentially the same. Locate the leading edge, the trailing edge, the chord, and the upper and lower camber on Figure 3-1.

Leading Edge:
The leading edge of an airfoil is the portion that meets the air first. The shape of the leading edge depends upon the function of the airfoil. If the airfoil is designed to operate at high speed, its leading edge will be very sharp, as on most current fighter aircraft. If the airfoil is designed to produce a greater amount of lift at a relatively low rate of speed, as in a Cessna 150 or a Cherokee 140, the leading edge will be thick and fat. Actually, the supersonic fighter aircraft

and the light propeller-driven aircraft are virtually two ends of a spectrum. Most other aircraft lie between these two.

Trailing Edge:
The trailing edge is the back of the airfoil, the portion at which the airflow over the upper surface joins the airflow over the lower surface. The design of this portion of the airfoil is just as important as the design of the leading edge. This is because the air flowing over the upper and lower surfaces of the airfoil must be directed to meet with as little turbulence as possible, regardless of the position of the airfoil in the air.

Chord:
The chord of an airfoil is an imaginary straight line drawn through the airfoil from its leading edge to its trailing edge. We might think of this chord line as the starting point for drawing or designing an airfoil in cross section. It is from this baseline that we determine how much upper or lower camber there is and how wide the wing is at any point along the wingspan. The chord also provides a reference for certain other measurements as we shall see.

Camber:
The camber of an airfoil is the characteristic curve of its upper or lower surface. The camber determines the airfoil's thickness. But, more important, the camber determines the amount of lift that a wing produces as air flows around it. A high-speed, low-lift airfoil has very little camber. A low-speed, high-lift airfoil, like that on the Cessna 150, has a very pronounced camber. You may also encounter the terms upper camber and lower camber. Upper camber refers to the curve of the upper surface of the airfoil, while lower camber refers to the curve of the lower surface of the airfoil. In the great majority of airfoils, upper and lower cambers differ from one another. Bernoullis Principle: Daniel Bernoulli, an eighteenth-century Swiss scientist, discovered that as the velocity of a fluid increases, its pressure decreases. How and why does this work, and what does it have to do with aircraft in flight? Bernoulli's principle can be seen most easily through the use of a venturi tube (see Animation or Figure 3-2). The venturi will be discussed again in the unit on propulsion systems, since a venturi is an extremely important part of a carburetor. A venturi tube is simply a tube which is narrower in the middle than it is at the ends. When the fluid passing through the tube reaches the narrow part, it speeds up. According to Bernoulli's principle, it then should exert less pressure. Let's see how this works.

As the fluid passes over the central part of the tube, shown in Animation or Figure 3-2, more energy is used up as the molecules accelerate. This leaves less energy to exert pressure, and the pressure thus decreases. One way to describe this decrease in pressure is to call it a differential pressure. This simply means that the pressure at one point is different from the pressure at another point. For this reason, the principle is sometimes called Bernoulli's Law of Pressure Differential. Bernoulli's principle applies to any fluid, and since air is a fluid, it applies to air. The camber of an airfoil causes an increase in the velocity of the air passing over the airfoil. This results in a decrease in the pressure in the stream of air moving over the airfoil. This decrease in pressure on the top of the airfoil causes lift. Section 3.3 - RELATIVE WIND In order to discuss how an airfoil produces lift or why it stalls, there are three terms we must understand. These are relative wind, angle of incidence, and angle of attack. There is a noticeable motion when an object moves through a fluid or as a fluid moves around an object. If a thick stick is moved through still water or the same stick is held still in a moving creek, relative motion is produced. It does not matter whether the stick or the water is moving. This relative motion has a speed and direction. Now let's replace the water with air as our fluid and the stick with an airplane as our object. Here again, it doesn't matter whether the airplane or the air is moving, there is a relative motion called relative wind. The relative wind will be abbreviated with the initials RW (see figure 3-3). Since an airplane is a rather large object, we will use a reference line to help in explaining the effects of relative wind. This reference is the aircraft's longitudinal axis, an imaginary line running from the center of the propeller, through the aircraft to the center of the tail cone.

Note in Figure 3-4 that the relative wind can theoretically be at any angle to the longitudinal axis. However, to maintain controlled flight, the relative wind must be from a direction that will produce lift as it flows over the wing. The relative wind, therefore, is the airflow produced by the aircraft moving through the air. The relative wind is in a direction parallel with and opposite to the direction of flight. Let's look a little closer at how relative wind affects an airplane and its wings. As shown in Figure 3-3, the chord line of the wing is not parallel to the longitudinal axis of the aircraft. The wing is attached so that there is an angle between the chord line and the longitudinal axis. (We call this difference the angle of incidence.) Since we describe relative wind (relative motion) as having velocity (speed and direction), the relative wind's direction for the wing is different from that of the fuselage. It should be easy to see that the direction of the relative wind can also be different for the other parts of the airplane. Very briefly, angle of attack is a term used to express the relationship between an airfoil's chord and the direction of its encounter with the relative wind. This angle can be either positive, negative, or zero. When speaking of the angle of attack, we normally think of the relative wind striking the airfoil from straight ahead. In practice, however, this is true only during stabilized flight which is in a constant direction. Section 3.4 - REVIEW EXERCISE

1. The part of an airplane that produces lift is called a(n)

2. The baseline for designing an airfoil in cross section is determined from the

3. The amount of lift that a wing can produce is determined by the

4. Bernoulli's principle states that as the velocity of a fluid increases, its pressure

. in pressure on top of an airfoil as the air speeded up.

5. Lift is produced by the passing over the top of the airfoil is

6. Name the parts of the airfoil shown by the arrows.

A. B. C. D. E. F.

7. The relative wind is airflow produced by the 8. The relative wind flows in a direction direction of flight.

moving through the air. with and

to the

9. The angle between the chord of the wing and the longitudinal axis of the aircraft is called the

10. The wing and the relative wind.

is the angle between the chord of the

Chapter 4 - The four forces of flight - Level 2

At the end of this block of study, you should be able to: Define lift. State the relationship between airspeed, camber, angle of attack, and lift. Give the four forces of flight and tell which of these forces oppose each other. Describe maximum gross weight, empty weight, center of gravity, center of lift, and useful load with relation to an airplane. Define induced drag and parasite drag and give two examples of each.

Four forces of flight in balance. Sections in this Chapter: Section 4.1 - LIFT Section 4.2 - AIRSPEED, CAMBER, AND LIFT Section 4.3 - LIFT AND WEIGHT Section 4.4 - DRAG Section 4.5 - INDUCED DRAG Section 4.6 - PARASITE DRAG Section 4.7 - REVIEW EXERCISE

Section 4.1 - LIFT

We know that we can cause reduced pressure in a fluid if the velocity of its flow is increased (Based on Bernoulli's principle - section 3-2 ). The camber of an airfoil's upper surface makes the air flowing over it move faster than the air flowing under the wing. This increase in velocity reduces the pressure (P1>P2) on the top of the wing so lift is produced. (See Figure 4-1). Lift is also called airfoil lift or Bernoulli's lift. Lift will continue as long as the airfoil is moving through the air and the air remains smooth rather than turbulent. Every airfoil, no matter what its camber or chord, will lose its smooth flow at some point along the upper surface. The perfect airfoil, if there were such a thing, would have turbulent flow at its trailing edge where the divided airstream comes together again. The regular airfoil has turbulence somewhere forward of the trailing edge even though level flight is maintained. With every increase in angle of attack (See Figure 4-2), this turbulent flow moves farther and farther toward the leading edge. The increase in angle of attack increases lift. This is true up to a point because we also must consider the power needed to force the craft through the air. If we had unlimited power, angle of attack would be of no concern, but this is not the normal situation so the turbulent flow continues forward until there is no more lift available. Dynamic lift It may interest you to know, at this point, that lift can also be created by an airfoil without any camber at all. This lift, however, is completely different from the lift we have been talking about. This type of lift is called dynamic lift and is caused by the pressure of impact air against the lower surface of the airfoil. A kite flying on a balmy spring day is an excellent example of an airfoil without camber being sustained in flight by dynamic lift. Similar to the airfoil in the wind tunnel, it makes no difference to the kite whether it is moving forward through the air or the air is moving past it. It simply goes on and hangs up there in the spring sky. (If you have flown a kite, however, you know there's a difference. You know that when the wind is light, you have to run your legs off at times to get the kite airborne.) This same kind of lift also helps hold the aircraft in the air and can be explained by Newton's third law of motion. Newton's third law of motion states that for every action there is an equal and opposite reaction. A popular example of this law is the gun and the bullet shown in Figure 4-3. When the trigger is pulled and the gun fires, the bullet leaving the barrel is the action and the recoil of the gun is the reaction. If we can ignore friction and air resistance, the force of the bullet striking the wall and the force of the gun striking the opposite wall will be equal.

Section 4.2 AIRSPEED, CAMBER, AND LIFT

The energy factors at the upper surface of a wing, as we have said, are velocity and pressure higher velocity, lower pressure. If the velocity of the relative wind is normally very high during cruising flight of an airplane, it is not necessary for its wings to have much camber. This is one of the reasons why fighter-type military aircraft have thin wings. At slower speeds, such as during takeoffs or landings, the loss of induced lift because of the low camber is compensated for by using a high angle of attack. As you can see, this high angle of attack causes an increase in the dynamic lift. Even so, the airplane with low-camber airfoils must use much higher takeoff and landing speeds than the more conventional airplane. To further illustrate these points, note in the top portion of Figure 4-4 that we have two examples of airfoils with the same relative wind velocity and the same dynamic lift. However, by thickening and increasing the camber of the wing, wing B's total lift is increased because of the increased induced lift. In the lower portion of Figure 4-4, you are looking at two wings which are producing the same mount of total lift even though one wing has less amber than the other. Both wings are at the same angle of attack so they have the same amount of dynamic lift for any given airspeed (velocity of the relative wind). The only way to make the thin wing produce as much lift as the thick wing is to speed it up, and this is what we attempt to show in the figure. Wing C's relative wind is ten miles per hour faster than D's relative wind, this additional speed is needed to increase both the dynamic and induced lift so that its total lift can equal that of Wing D. We want you to understand that the examples in Figure 4-4 are just that. We have discussed the atmosphere and how airfoils produce lift because of their movement through the atmosphere. We also mentioned that lift is the force that counteracts the force of gravity to allow flight. At this point, you may have concluded that lift and gravity are the only forces involved with flight. Actually there are two others, thrust and drag, which complete the threedimensional forces acting upon an aircraft in flight. Figure 4-5 shows the basic directions of all four forces when an aircraft is in straight and level flight at a constant speed. Now, you should be able to see that, in this situation, the four forces are in balance. The force of total lift equals the force of total weight, so there is no upward or downward movement. The force of thrust equals the force of drag, so there is no increase or decrease in the speed of the airplane. You should also be able to see that the moment one of these forces becomes stronger or weaker than the others, some type of reaction must take place.

Section 4.3 - LIFT AND WEIGHT

With these two forces in opposition to each other, it is obvious that increased lift and decreased weight are objectives in both the designing and flying of aircraft. Induced lift can be increased, as has been mentioned before, by changing the camber, or curvature, of the airfoil. Work continues in an effort to achieve the most efficient designs possible. But more important, at least to the person who is flying an airplane, is the angle of the airfoil as it encounters the relative wind (angle of attack). As indicated earlier, lift is increased as the angle of attack is increased because there is more relative wind striking the airfoil's bottom surface, creating higher pressure. There is also an increase in the induced lift, because at a higher angle of attack the air has to travel even farther over the top surface of the wing. There is a point in this relationship of airfoil to angle of attack where lift is destroyed and the force of gravity (weight) takes command. This is called stall. The air can no longer flow smoothly over the wing's upper surface. Instead, the air burbles over the wing and lift is lost. You might wonder why the force of power from the engine can't take the place of the loss of lift from the airfoil. Very simply, there just isn't enough of this force available from a conventional aircraft's engine. Some of the more powerful jet fighters and acrobatic sport airplanes can, for a short time and distance, climb straight up without any significant help from their airfoils. However, these airplanes will eventually stall and start to fall toward Earth. The stalled condition is one from which recovery (and continued flight) is fairly easy. At this point, we should mention another situation where lift can no longer overcome weight. No matter how efficient the airfoils and power plant of an aircraft may be, there is still a limit as to how high in the atmosphere it can go. This limit is called the aircraft's ceiling. At its ceiling, the aircraft's power plant is producing all possible power, and the airfoils are producing all possible lift just to equal the force of the aircraft's weight. Why? The atmosphere, you will remember, becomes less and less dense as altitude increases. The aircraft's ceiling is that point in the atmosphere where the air is too thin to allow further increase in lift. Since weight is a problem to be overcome when we speak of lift, how do we manage the weight problem? First of all, the airplane must be constructed of the lightest-weight materials that can be used according to the type flying for which the airplane is designed. Today, most airplanes are built of metal, with aluminum alloy being used extensively because of its strength and light weight. The weight of the load the airplane carries also receives very careful consideration. Each airplane has a total weight limitation called maximum gross weight, above which the airplane is unsafe for flight. It is possible to keep putting luggage or other cargo into an airplane until it is so heavy it will not fly. Since the pilot cannot put the airplane on a scale to make sure it doesn't exceed the maximum gross weight, another approach must be used. This involves computations that were begun during the design and testing of the airplane since the maximum gross weight is established by the manufacturer. Then, as each airplane is completed, it is weighed at the assembly plant and this weight is entered in certain documents (which must remain with the airplane at all times) as the empty weight. Thus, the difference between empty weight and maximum gross weight tells the pilot how much weight can be put in the airplane without overloading. Incidentally, this amount of weight is called useful load.

Where the weight is placed in the airplane is another factor that has a pronounced effect on how well the airplane will fly. This is because the center of gravity (CG) of the airplane must be maintained within certain limits prescribed by the manufacturer. The CG of an aircraft is the point where all of the weight of the aircraft is considered to be located (see figure 4-6). In order for the aircraft to be flown safely, the CG must be kept within certain limits with relationship to the center of lift (CL). The center of lift is the point at which all of the lift on the aircraft is considered to be concentrated. Notice in Figure 4-6 that there is a forward and an aft CG limit. If the CG gets too far forward or too far aft of the CL the aircraft will be out of balance and difficult, if not impossible, to control.

Section 4.4 - DRAG


Thrust is the force that propels the aircraft forward. Thrust for aircraft is obtained from the different types of engines discussed earlier. An airplane cannot gain altitude or maintain straight and level flight unless its engine is producing enough thrust to propel (pull or push) the airfoils fast enough to produce the needed amount of lift. Without this thrust, the airplane will continue to fly (it will not "drop out of the sky,' as many people think) but its flight becomes a gradual descent toward the ground. Without the needed thrust, weight has more influence than lift and pulls the airplane toward the ground. Helping the force of weight is drag. Drag is present at all times and can be defined as the force which opposes thrust, or, better yet, it is the force which opposes all motion through the atmosphere and is parallel to the direction of the relative wind.(See Figure 4-7)

But, what causes drag? It is caused, purely and simply, by the resistance of air. Air, you will remember, is a fluid and has mass. When you stick your hand out of the window of a moving automobile, you do several things. First, you may violate the law in some sections of the country-in addition to possibly getting your hand clipped off by a tree! Second, you experience (or feel) the relative wind created by the car's forward movement. Remember the relative wind? It is the wind moving past an object and the object, in this case, is the car. Your hand, in effect, becomes an extension of the car in experiencing the relative wind. Third, you may possibly create lift. If you arch your hand slightly (you're really giving it some camber), your hand may tend to rise. If you place your hand at a slight angle to the relative wind, the impact air will cause your hand to rise. But fourth and for sure, you will encounter the resistance and experience drag. This drag will tend, then, to push your hand backwards. Aircraft in flight encounter the same force as your hand, but aircraft are designed to fly, rather than to do all the things that your hand can do. Aeronautical engineers realize that drag, like the other forces acting on an aircraft in flight, is made up of a number of components. One way to look at the total drag is to divide it up into two fairly broad divisions, induced drag and parasite drag. Let's look at these two, one at a time.

Section 4.5 - INDUCED DRAG


Induced drag is the unavoidable by-product of lift and increases as the angle of attack increases. Remember, the greater the angle of attack, up to a critical angle, the greater the amount of lift developed and the greater the induced drag. Since there are two different ways that lift is produced, there are also two different types of induced drag: dynamic drag (Newtonian) and pressure drag (Bernoulli). First, let's consider dynamic induced drag, shown in Figure 4-8. If you hold your hand out of the window of a moving car, with the front edge tipped up at an angle to the relative wind to give it an angle of attack, you will feel a force pushing your hand back, but also slightly upward. In other words, depending on the angle of attack, there will be a force backward (induced drag) and a force upward (lift). The amount of force in each direction will depend on the angle of attack. If the angle of attack is small, the drag and lift are comparatively small. Any increase in angle of attack, up to a certain point, will increase drag and lift. However, at very high angles of attack, approaching the stall point, lift will decrease and the drag will overcome lift and thrust with an accompanying loss of speed and attitude. If you were to hold your hand vertical to the relative wind, the only force would be backward; that is, all dynamic drag and no lift. Now let's consider pressure-induced drag, which can be divided into the two types. You will remember that the thin layer of air over the upper surface of the wing will break away from the wing at high angles of attack and that the flow will become turbulent as the flow of air breaks away from the wing. This turbulence results in pressure drag and loss of lift. Turbulence and pressure drag also result from the flow of air around the wingtip as the comparatively high-pressure air under the wing flows over the wingtip to the low-pressure area on top of the wing.

Section 4.6 - PARASITE DRAG


So far, we have discussed only induced drag. There are also skin-friction drag and form drag, which are referred to as parasite drag. All drag other than induced drag is parasite drag. Skin-friction drag is caused by the friction between outer surfaces of the aircraft and the air through which it moves. It will be found on all surfaces of the aircraft: wing, tail, engine, landing gear, and fuselage. Form drag is also a resistance to the smooth flow of air. The shape of something may create low-pressure areas and turbulence which retard the forward movement of the aircraft (see figure 4-9). Streamlining the aircraft will help eliminate form drag. Parts of an aircraft which do not lend themselves to streamlining are enclosed either partially or wholly in covers called fairings which have a streamlined shape.

Chapter 5 - Flight Controls - Level 2


We pointed out earlier that aircraft are able to sustain flight because of a balance among the four forces we've just discussed: lift, weight. thrust, and drag. When the aircraft is in the air and flying straight and level, we can assume that these four forces are in balance. But you may say, the aircraft has to get there somehow. So. what other types of aircraft motion are there besides straight, level, and unaccelerated flight?

At the end of this block of study, you should be able to: Identify the three axes of rotation. Describe the surfaces and devices used to control the flight of an aircraft. Explain how aircraft flight is controlled. Describe the basic flight maneuvers an airplane can make. Identify the actions needed for carrying out basic flight maneuvers. Explain the cause and effect of a stall. Sections in this Chapter: Section 5.1 - THE AXES OF ROTATION Section 5.2 - BASIC FLIGHT MANEUVERS Section 5.3 - TAKEOFF AND CLIMB Section 5.4 - LANDING Section 5.5 - STALLS

Section 5.1 - THE AXES OF ROTATION


Aircraft fly in three dimensions, and they move in directions other than straight and level. In order to examine these other directions, we have to take another look at our aircraft. In addition to moving forward, an aircraft in flight may move about three axes. See Figure 5-1 and you will understand what we mean. The simplest way to understand the axes is to think of them as long

rods passing through the aircraft where each will intersect the other two. At this point of intersection, called the center of gravity, each of these axes is also perpendicular to the other two.

The axis that extends lengthwise (nose through tail) is called the longitudinal axis, and rotation about this axis is called roll. The axis that extends crosswise (wingtip through wingtip) is called the lateral axis, and rotation about this axis is called pitch. The axis that passes vertically through the center of gravity (when the aircraft is in level night) is called the vertical axis, and rotation about this axis is called yaw. There is apparently no real rationale for these names; you simply have to memorize them (longitudinal axis-roll, lateral axis-pitch, and vertical axis-yaw). If you hold a model in your hand in a straight and level manner and roll it so as to dip a wing, you are demonstrating roll around the longitudinal axis. If you hold it level and move the nose up and down, you are demonstrating movement around the lateral axis-that is, you are changing pitch. If you hold the model straight, and move it through the air while turning the nose to one side or the other, you have it yawing on its vertical axis; in other words, it is not moving forward in the direction in which the nose is pointing. An aircraft may move about one or all of these axes at the same time. With various types of turns and maneuvers, all three may be used. As you begin to perceive, the axes of rotation control an aircraft's maneuverability. Still holding your model airplane, try the same movements again, but carry them further. Dip a wing and roll 180 degrees and you are flying upside down. Roll another 180 degrees and you are again straight and level. Your aircraft has executed a barrel roll. Change pitch 90 degrees downward and you are diving toward the ground. Continue the rotation another 90 degrees and you have reversed your direction. You are also upside down. In other words, if you were flying east, upright, you are now flying west, upside down. Continue the rotation another 90 degrees and you are climbing straight up. Complete a 360-degree circle, and you are again flying east, straight and level. Your aircraft has executed an outside loop. To do an inside loop, change pitch in an upward direction, again making a complete circle. Yaw, of course, is involved in every turn. You can describe turns of numerous degrees and angles with your model. Now that we understand the axes of rotation, let's look at how the aircraft is controlled (moved) around these axes.

The Longitudinal Axis


Running from the nose to the tail of an aircraft is the longitudinal axis (see figure 5-2). This axis can be thought of as a skewer which runs the length of the fuselage, and movement around the longitudinal axis is called roll. The cause of movement or roll about the axis is the action of the ailerons. Ailerons are attached to the wing and to the control column in a manner that ensures one aileron will deflect downward when the other is deflected upward. How is it that deflecting an aileron causes the wing to move? Very simply, when an aileron is not in perfect alignment with the total wing, it changes the wing's lift characteristics. To make a wing move upward, the aileron on that wing must move downward. When this happens, the total lift being produced by that wing is increased. At the same time, the lift on the other wing is reduced. This causes the aircraft to roll. The ailerons are attached to the cockpit control column by mechanical linkage. When the control wheel is turned to the right (or the stick is moved to the right), the aileron on the right wing is raised and the aileron on the left wing is lowered. This action increases the lift on the left wing and decreases the lift on the right wing, thus causing the aircraft to roll to the right. Moving the control wheel or stick to the left reverses this and causes the aircraft to roll to the left.

The Lateral Axis


The Lateral Axis is another name for the lateral axis is the pitch axis. This name makes sense because the airplane is actually caused to pitch its nose upward or downward around the lateral axis which runs from wingtip to wingtip. What causes this pitching movement? It is the elevator which is attached to the horizontal stabilizer. The elevator can be deflected up or down as the pilot moves the control column (or stick) backward or forward. Movement backward on the control column moves the elevator upward. As shown in Figure 5-3 , the relative wind striking the top surface of the raised elevator pushes the tail downward. Since this motion is around the lateral axis, as the tail moves (pitches) downward, the nose moves (pitches) upward and the aircraft climbs. Deflection of the elevator downward creates the opposite effect and the relative wind striking the lower surface of the elevator causes the tail to pitch up. This, as you can see in Figure 5- , pitches the nose of the aircraft downward and the airplane dives. Before leaving the horizontal stabilizer, we should introduce the term stabilizer. Most aircraft designs no longer use a stabilizer with elevator arrangement. Instead the entire horizontal tail surface is hinged so that the surface's angle of attack is changed as the pilot pulls or pushes on the control column. This type of design is doing the job of both a stabilizer and elevator so it is called a stabilator.

The Vertical Axis


The third axis which passes through the meeting point of the longitudinal and lateral axes from the top of the aircraft to the bottom is called the vertical or yaw axis. The aircraft's nose moves about this axis in a side-to-side direction. In other words, the airplane's nose is made to point in a different direction when the airplane turns about this particular axis. The airplane's rudder, which is moved by pressing on the rudder pedals, is responsible for movement about this axis (see figure 5-4 ). The rudder is a movable control surface attached to the vertical fin of the tail assembly. By pressing the proper rudder pedal, the pilot moves the rudder of the aircraft in the direction of the pedal he or she presses (right pedal moves the rudder to the right, and left pedal moves the rudder to the left). What happens then? When the pilot pushes the left rudder pedal, he or she then sets the rudder so that it deflects the relative air to the left. This then creates a force on the tail, causing it to move to the right and the nose of the aircraft to yaw to the left. At this point, it is probably well to point out that the rudder does not steer the aircraft in normal flight. The rudder does not turn the aircraft; rather, its primary purpose is to offset the drag produced by the lowered aileron.

Section 5.2 BASIC FLIGHT CONTROLS


Basic flight maneuvers include climbs, descents, turns, and combinations of these. generally, the basic flight maneuvers are started from what is called straight and level flight. Straight and level flight (also called controlled flight) is a flight condition where the wings are kept level and the altitude and heading constant. Power setting is maintained at 55 percent to 75 percent of available power, using a higher setting within this range if speed is desired and a lower setting if fuel economy is desired. Straight and level flight is really a series of slight adjustments or corrections in pitch, yaw, and roll to keep the wings level and heading and altitude constant. A very good pilot can hold a constant heading and altitude so smoothly that a passenger can scarcely see the control movements. Climbs are a combination of power and "up elevator." The amount of power used determines whether the climb is steep or shallow. If, for example, a pilot is taking off and must clear trees near the end of the runway, all available power must be used and the climb angle must be as steep as possible. This is called the best angle of climb, but it is a short-term climb. A sustained climb at this angle can overheat the engine because there is too little cooling air flowing around the engine's cylinders. The reason the airflow is reduced is the relatively low airspeed resulting from the steep climb angle. Normal descents are a combination of reducing power and adjusting to maintain the desired airspeed. The airspeed is maintained by varying pressure on the control wheel. This, as you know, varies the angle of attack and, consequently, airspeed.

The third basic maneuver is the turn. Turns are either gentle, medium, or steep; and they may be made when climbing, descending, or while not gaining or losing altitude. Causing the airplane to turn requires smooth coordination of aileron, rudder, and elevator controls; in other words, pressure on the control wheel and the rudder pedal should be applied simultaneously. Why? The moment a wing begins to rise in a banked turn, it experiences more drag because of the lowered aileron and its higher angle of attack. A simultaneous application of rudder compensates for this additional drag by making the airplane also rotate about its vertical axis (see figure 5-5 ).

Section 5.3 - TAKEOFF AND CLIMB


After taxiing to the runway, a pre-takeoff check list is accomplished. This check is to ensure that all systems are working normally. When this is completed, the airplane is taxied to the center of the runway and aligned with it. The throttle is opened fully to start the takeoff run (also called take off roll).During this takeoff run, the control wheel, or stick, is usually held in the neutral position, but the rudder pedals are used to keep the airplane on the runway's centerline. As takeoff airspeed is approached, gentle back pressure on the control wheel raises the elevator which causes the airplane's nose to pitch upward slightly. This lifts the nose wheel off the runway (see fig. 5-6 ).

Once the nose wheel is off the runway, the right rudder is applied to counteract the left-turning tendency which is present under low airspeed and high-power flight conditions. As the airplane lifts clear of the runway, the pilot varies the pressure on the control wheel. First, pressure is relaxed slightly to gain airspeed while still in ground effect (additional lift provided by compression of air between the airplane's wings and the ground). As airspeed increases to the best rate-of-climb airspeed, back pressure on the control wheel is adjusted to maintain that

airspeed until the first desired altitude is reached. (Best rate-of-climb airspeed provides the most altitude for a given unit of time.) Climbs to other and higher altitudes are made at airspeeds determined by the pilot, until the desired cruising altitude is reached. Upon reaching cruising altitude, the airplane's pitch attitude is reduced and the airplane accelerates to cruising speed. The power is reduced and adjusted to maintain the selected cruising speed. Almost simultaneously, the pilot adjusts the elevator and possibly the rudder to keep the airplane at the desired altitude and heading (direction). If the flight is to go to a distant airport, the airplane will be kept in its cruising flight configuration until the destination is near. If the pilot wants only to perform basic flight maneuvers in a practice area, the cruising flight configuration will necessarily be changed rather soon.

Section 5.4 - LANDING


A good landing begins with a good approach (see figure 5-7 ). Before the final approach is begun, the pilot performs a landing checklist to ensure that critical items such as fuel flow, landing gear down, and carburetor heat on are not forgotten. Flaps are used for most landings because they permit a lower- approach speed and a steeper angle of descent. This gives the pilot a better view of the landing area. The airspeed and rate of descent are stabilized, and the airplane is aligned with the runway centerline as the final approach is begun. When the airplane descends across the approach end (threshold) of the runway, power is reduced further (probably to idle). At this time, the pilot slows the rate of descent and airspeed by progressively applying more back pressure to the control wheel. The airplane is kept aligned with the center of the runway mainly by use of the rudder.

Continuing back pressure on the control wheel, as the airplane enters ground effect and gets closer and closer to the runway, further slows its forward speed and rate of descent. The pilot's objective is to keep the airplane safely flying just a few inches above the runway's surface until it loses flying speed. In this condition, the airplane's main wheels will either "squeak on" or strike the runway with a gentle bump. With the wheels of the main landing gear firmly on the runway, the pilot applies more and more back pressure on the control wheel. This holds the airplane in a nose-high attitude which keeps the nose wheel from touching the runway until forward speed is much slower. The purpose here is to avoid overstressing and damaging the nose gear when the nosewheel touches down on the runway. The landing is a transition from flying to taxiing. It demands more judgment and technique than any other maneuver. More accidents occur during the landing phase than any other phase of flying. Variables such as wind shear and up-and-down draft add to the problem of landing. Good pilots can be easily recognized. They land smoothly on the main wheels in the center of the runway and maintain positive directional control as the airplane slows to taxiing speed.

Section 5.5 - STALLS


Since stalls are the cause of much concern among student pilots and the nonflying public, we will discuss them here. We mentioned that an airplane must attain flying speed in order to take off. Sufficient airspeed must be maintained in flight to produce enough lift to support the airplane without requiring too large an angle of attack. At a specific angle of attack, called the critical angle of attack, air going over a wing will separate from the wing or "burble" (see figure 5-8 ), causing the wing to lose its lift (stall). The airspeed at which the wing will not support the airplane without exceeding this critical angle of attack is called the stalling speed. This speed will vary with changes in wing configuration (flap position). Excessive load factors caused by sudden maneuvers, steep banks, and wind gusts can also cause the aircraft to exceed the critical angle of attack and thus stall at any airspeed and any attitude. Speeds permitting smooth flow of air over the airfoil and control surfaces must be maintained to control the airplane. Flying an airplane, like other skills that are learned, requires practice to remain proficient. Professional pilots for the major airlines, military pilots, and flight instructors all return to the classroom periodically for updating their skills. Good judgment must be exercised by all pilots to ensure the safe and skillful operation of the airplanes they fly.

Chapter 6 - Aircraft Propulsion


The Reciprocating Engine At the end of this block of study, you should be able to: Define energy, potential energy, and kinetic energy. Explain the application of Boyle's and Charles' laws to the operation of a reciprocating engine. Identify the seven major parts of a reciprocating engine. Identify the actions in one complete cycle of a four-stroke reciprocating engine. Describe the difference between a fixed-pitch and a variable-pitch propeller. Explain how a propeller creates thrust. Discuss the effect "feathering" has on a propeller. Reciprocating engines power the conventional vehicles that we use for transportation, work, and pleasure. Reciprocating engines provide power for our automobiles, lawn mowers, tractors, motorcycles, boats, trains, airplanes, and a multitude of other devices used in today's modern lifestyle. All reciprocating engines are basically the same. They have the same major parts; most of them use liquid fuel; and all of them require an ignition system, a cooling system, and a lubrication system. The term reciprocating is the common denominator; it means that certain parts move back and forth in a straight-line motion. This straight-line motion has to be changed to rotary motion in order to turn the wheels of automobiles and trains and the propellers of boats and airplanes. You will understand how this is done as you read and think about what is written and shown in this chapter.

AIRCRAFT TURBINE AND RAMJET ENGINES At the end of this block of study, you should be able to: Describe the operation of a turbojet engine. List the advantages of the turbofan engine. Describe the operation of a turboprop engine. Explain the operation of the ramjet engine. All modern, powered airplanes that do not use reciprocating engines as their source of thrust use some type of turbine engine. The word turbine means whirl and refers to any type of wheel device that has vanes attached to it in a manner that will cause the wheel to turn as the vanes are struck by the force of a moving fluid. Remember, air is a fluid. The turbine principle is also used to generate electricity by flowing

water striking a turbine that is linked to a generator. Another method of generating electricity is to direct high-pressure steam against a turbine which is linked to a generator. Turbine engines found in aircraft use the force of hot, flowing gases striking a turbine. Some of these engines are geared to propellers which are similar to the types of propellers used with reciprocating engines. The turbine engine has also found widespread use as the source of power for military and civilian helicopters. In helicopters, the turbine is linked by gears to the helicopter's rotors in a manner that can be compared to the turbine-driven propellers for airplanes. Remembering that all turbine engines operate according to the same principles, let's briefly discuss the different types: the turbojet, the turbofan, and the turboprop. Sections in this Chapter: Section 6.1 - SCIENTIFIC PRINCIPLES AND TERMS Section 6.2 - RECIPROCATING-ENGINE OPERATING PRINCIPLES Section 6.3 - PROPELLERS Section 6.4 - TURBOJET ENGINES Section 6.5 - TURBOFAN ENGINES Section 6.6 - TURBOPROP ENGINES Section 6.7 - RAMJET ENGINES Section 6.8 - REVIEW EXERCISE

Section 6.1 SCIENTIFIC PRINCIPLES AND TERMS


Energy. Energy is the force behind the movement of all things. Animals (including humans) use food as their energy source for life and movement. Mechanical or inanimate objects use fuel energy to perform work. What is energy? Let's review what you probably have already learned in school. The word ENERGY as defined in physics, is "the capacity for doing work and overcoming resistance." Unless it is doing work, energy is known as potential energy (stored energy). The fuel (usually gasoline) used to power reciprocating engines is potential energy up to the moment it is mixed with air (oxygen) and burned. When potential energy is released from its source and causes movement of an object, it becomes kinetic energy (active energy). Thus. the movement of the parts of a reciprocating engine is an example of the potential energy of the fuel having been changed to kinetic energy. Potential and kinetic are broad classifications of energy. Energy is also given several other titles depending on the form it is in at a given moment. That is, energy can be changed from one form to another, so various titles are used to describe the forms. As examples: Heat energy can be changed into mechanical energy; mechanical energy can be changed into electrical energy; and electrical energy can be changed into heat, mechanical, or light energy. Boyle's and Charles' Laws. Boyle's law states that the volume of a gas varies inversely with the pressure on it (see figure 6-1 ). This means that any confined gas will double its pressure if its volume is decreased by one-half. If we have a cylinder in which ordinary air is present at 14.7 pounds per square inch

(psi) and we rammed an airtight piston into the cylinder one-half the length of the cylinder, the pressure of the gas, or air, would double to 29.4 psi. Then, if we were to ram the piston an additional one-half of the remaining distance in the cylinder, the pressure would increase to 58.8 psi.

What happens when the reverse is tried? Let's suppose that we begin with the piston in one-half the length of the cylinder. (The pressure within the cylinder is 14.7 psi.) If the piston were extracted quickly to the full length of the cylinder, what do you think would happen to the pressure within the cylinder? It would be reduced by one-half, to become 7.35 psi. We can say this another way. When the volume of a confined gas is doubled, its pressure is reduced by onehalf. As a general summary of Boyle's law, you should remember that a decrease in volume causes an increase in pressure. An increase in volume causes a decrease in pressure. When the piston in a cylinder moves inward and outward, increasing and decreasing the pressure of a confined gas, what is happening to the temperature of the gas? Charles' law states, the pressure and temperature of a confined gas are directly proportional. Thus, when the piston in a cylinder moves inward and compresses the gas, the temperature of the gas increases. How much the temperature increases depends on how far the piston t ravels. While an aircraft engine is operating, these two laws are being applied. It is through the understanding of these and related laws of physics that engineers have been able to improve the efficiency of engines.

Section 6.2 RECIPROCATING-ENGINE OPERATING PRINCIPLES


The reciprocating engine is also known as an internalcombustion engine. This name is used because the fuel mixture is burned within the engine. To understand how a reciprocating engine works, we must first study its parts and the functions they perform. The seven major parts are: (1) The cylinders (2) The pistons (3) The connecting rods (4) The crankshaft (5) The valves (6) The spark plugs (7) A valve operating mechanism (cam). Refer to the relative location of these parts in Figure 6-2 .

Engine Operation. The cylinder is closed on one end (the cylinder head), and the piston fits snugly in the cylinder. The piston wall is grooved to accommodate rings which fit tightly against the cylinder wall and help seal the cylinder's open end so that gases cannot escape from the combustion chamber. The combustion chamber is the area between the top of the piston and the head of the cylinder when the piston is at its uppermost point of travel. The up-and-down movement of the piston is converted to rotary motion to turn the propeller by the connecting rod and the crankshaft, just as in most automobiles. Note the crankshaft, connecting rod, and piston arrangement in Figure 6- and imagine how the movement of the piston is converted to the rotary motion of the crankshaft. Note particularly how the connecting rod is joined to the crankshaft in an offset manner. The valves at the top of the cylinder open and close to let in a mixture of fuel and air and to let out, or exhaust, burned gases from the combustion chamber. The opening and closing of a valve are done by a cam geared to the crankshaft. This gearing arrangement ensures that the two valves open and close at the proper times. Now let's consider the movement of the piston (four strokes) and the five events of a cycle (see figure 6-3 ).

1. The Intake Stroke The cycle begins with the piston at top center; as the crankshaft pulls the piston downward, a partial vacuum is created in the cylinder chamber. The cam arrangement has opened the intake valve, and the vacuum causes a mixture of fuel and air to be drawn into the cylinder. 2. & 3. Compression and Ignition Stroke As the crankshaft drives the piston upward in the cylinder, the fuel and air mixture is compressed. The intake valve has closed, of course, as this upward stroke begins. As the compression stroke is completed and just before the piston reaches its top position, the compressed mixture is ignited by the spark plug. 4. Power Stroke The very hot gases expand with tremendous force, driving the piston down and turning the crankshaft. The valves are closed during this stroke also.

5. Exhaust Stroke On the second upward (or outward, according to the direction the unit is pointed) stroke, the exhaust valve is opened and the burned gases are forced out by the piston. At the moment the piston completes the exhaust stroke, the cycle is started again by the intake stroke. Each piston within the engine must make four strokes to complete one cycle, and this complete cycle occurs hundreds of times per minute as the engine runs. The overall principles of reciprocating-engine operation are easy to understand if you remember what happens with each stroke that the piston makes. For this reason, you may find the chart in Table 6-3 helpful.

Table 6-3 Direction of Movement 1. 2. 3. 4. Inward (Down) Outward (Up) Inward (Down) Outward (Up) Event (what happens) Intake Compression and Ignition Power Exhaust

Reciprocating-Engine Horsepower. Most persons are acquainted with the term horsepower as applied to automobile and aircraft reciprocating engines. The term was coined by James Watt, the inventor of the steam engine, who wished to evaluate the power output of his steam engine. Watt hitched a horse to an apparatus and determined that the horse could lift 550 pounds one foot in one second. Thus, one horsepower became the power to lift 550 pounds one foot per second, or 33,000 foot-pounds per minute (550 x 60). If an aircraft reciprocating engine is rated at 150 horsepower, it means the engine is capable of producing this much power. However, the engine has to be running at a certain speed before that much power is produced. The same is true for all other types of reciprocating engines.

Section 6.3 - PROPELLERS


We can say that the propeller is the action end of an aircraft's reciprocating engine, because it converts the useful energy of the engine into thrust as it spins around and around. The propeller has the general shape of a wing, but the camber and chord (curvature and crosssectional length) of each section of the propeller are different, as shown in Figure 64 . The wing provides lift upward, while the propeller provides lift forward.

The wing has only one motion which is forward, while the propeller has forward and rotary motion. The path of these two motions is like a corkscrew as the propeller goes through the air (see figure 6-5 ). Like a wing, a propeller blade has a thick leading edge and a thin trailing edge. The blade back is the curved portion and is like the top of a wing. The blade face is comparatively flat and corresponds to the underside of a wing (see figure 6-6 for definitions of blade back and blade face). The blade shank is thick for strength and fits into a hub which is attached to the crankshaft directly or indirectly. The outer end of the blade is called the tip. Blade pitch is loosely defined as the angle made by the chord of the blade and its plane of rotation, as shown in Figure 6-6 . When the angle is great, the propeller is said to have high pitch. A high-pitch propeller will take a bigger bit of air and move the aircraft farther forward in one rotation than will a low-pitch propeller. Propellers may be classified as to whether the blade pitch is fixed or variable. The demands on the propeller differ according to circumstances. For example, in takeoffs and climbs more power is needed, and this can best be provided by low pitch. For speed at cruising altitude, high pitch will do the best job. A fixedpitch propeller is a compromise. There are two types of variable-pitch propellers adjustable and controllable. The adjustable propeller's pitch can be changed only by a mechanic to serve a particular purpose-speed or power. The controllable-pitch propeller permits pilots to change pitch to more ideally fit their requirements at the moment. In different aircraft, this is done by electrical or hydraulic means. In modern aircraft, it is done automatically, and the propellers are referred to as constant-speed propellers. As power requirements vary, the pitch automatically changes, keeping the engine and the propeller operating at a constant rpm. If the rpm rate increases, as in a dive, a governor on the hydraulic system changes the blade pitch to a higher angle. This acts as a brake on the crankshaft. If the rpm rate decreases, as in a climb, the blade pitch is lowered and the crankshaft rpm can increase. The constant-speed propeller thus ensures that the pitch is always set at the most efficient angle so that the engine can run at a desired constant rpm regardless of altitude or forward speed. Click here to see examples of early aviation propellers. The constant-speed propellers have a full-feathering capability. Feathering means to turn the blade approximately parallel with the line of flight, thus equalizing the pressure on the face and back of the blade and stopping the propeller. Feathering is necessary if for some reason the propeller is not being driven by the engine and is windmilling, a situation that can damage the engine and increase drag on the aircraft. Most controllable-pitch and constant-speed propellers also are capable of being reversed. This is done by rotating the blades to a negative or reverse pitch. Reversible propellers push air forward, reducing the required landing distance as well as reducing wear on tires and brakes.

Section 6.4 - TURBOJET ENGINES


The turbojet uses a series of fan-like compressor blades to bring air into the engine and compress it. An entire section of the turbojet engine performs this function, which can be compared to the compression stroke of the reciprocating engine. In this section, there is a series of rotor and stator blades. Rotor blades perform somewhat like propellers in that they gather and push air backward into the engine. The stator blades serve to straighten the flow of this air as it passes from one set of rotor blades to the next (see figure 6-7 ).

As the air continues to be forced further into the engine, it travels from the low-compression set of rotors and stators to the high-compression set. This last set puts what we might say is the final squeeze on the air. The combustion chamber receives the high-pressure air, mixes fuel with it, and burns the mixture. The hot, very high-velocity gases produced strike the blades of the turbine and cause it to spin rapidly. The turbine is mounted on a shaft which is connected to the compressor. Thus, the spinning is what causes the compressor sections to function. After passing the turbine blades, the hot, highly accelerated gases go into the engine's exhaust section. The exhaust section of the jet engine is designed to give additional acceleration to the gases and thereby increase thrust. The exhaust section also serves to straighten the flow of the gases as they come from the turbine. Basically, the exhaust section is a cone mounted in the exhaust duct. This duct is also referred to as the tailpipe. The shape of the tailpipe varies, depending on the design operating temperatures and the speed-performance range of the engine. With all the heat produced in the turbojet engine, you probably wonder how it is kept from overheating. Like most aircraft reciprocating engines, the jet is also air-cooled. Of all the air coming into the compressor section, only about 25 percent is used to produce thrust; the remaining 75 percent passes around the combustion chamber and turbine area to serve as a coolant.

Section 6.5 - TURBOFAN ENGINES


The turbofan engine has gained popularity for a variety of reasons. As shown in Figure 6-8, one or more rows of compressor blades extend beyond the normal compressor blades. The result is that four times as much air is pulled into the turbofan engine as in the simple turbojet. However, most of this excess air is ducted through bypasses around the power section and out the rear with the exhaust gases. Also, a fan burner permits the burning of additional fuel in the fan airstream. With the burner off, this engine can operate economically and efficiently at low altitudes and low speeds. With the burner on, the thrust is doubled by the burning fuel, and it can operate on high speeds and high altitudes fairly efficiently. The turbofan has greater thrust for takeoff, climbing, and cruising on the same amount of fuel than the conventional turbojet engine.

With better all-around performance at a lower ate of fuel consumption, plus less noise resulting from its operation, it is easy to understand why most new jet-powered airplanes are fitted with turbofan engines. This includes military and civilian types. Jet Engine Thrust. The force produced by a jet engine is expressed in terms of pounds of thrust. This is a measure of the mass or weight of air moved by an engine times the acceleration of the air as it goes through the engine. Technically, if the aircraft were to stand still and the pressure at the exit plane of the jet engine was the same as the atmospheric pressure, the formula for the jet engine thrust would be: weight of air in pounds per second X velocity Thrust = -------------------------------------------------------------32.2 (normal acceleration due to gravity, in feet per second2) Imagine an aircraft standing still, capable of handling 215 pounds of air per second. Assume the velocity of the exhaust gases to be 1,500 feet per second. The thrust would then be: 215 lbs of air per second Thrust = --------------------------X 1,500 feet per second= 6.68 X 1,500 = 32.2 feet per second2 Thrust = 10,020 lbs If the pressure at the exit plane is not the same as the atmospheric pressure and the aircraft were not standing still, the formula would be someone different (See Level 3). It is not very practical to try to compare jet engine output in terms of horsepower. As a rule of thumb, however, you might remember that at 375 miles per hour (mph), one pound of thrust equals one horsepower; at 750 mph, one pound of thrust equals two horsepower.

Section 6.6 - TURBOPROP ENGINES


The turboprop engine is an effort to combine the best features of turbojet and propeller aircraft. The first is more efficient at high speeds and high altitudes; the latter is more efficient at speeds under 400 mph and below 30,000 feet. The turboprop uses a gas turbine to turn a propeller. Its turbine uses almost all the engine's energy to turn its compressor and propeller, and it depends on the propeller for thrust, rather than on the high-velocity gases going out of the exhaust. Strictly speaking, it is not a jet. Study Figure 6-9 and note how the turbine turns the compressors and the propeller. The gas turbine can turn a propeller with twice the power of a reciprocating engine. Reduction gears slow the propeller below the turbine's rpm, and this must be done because of the limitations of propellers. That is, no propeller is capable of withstanding the forces generated when it is turned at the same rate as that of the gas turbine. Even so, the turboprop engine receives fairly extensive use in military and civilian aviation circles. In summary, aircraft turbine engines may be classified as turbojet, turbofan, or turboprop. As a group, the turbine engines have many advantages over reciprocating engines, the most obvious being the capability of higher-altitude and higher-speed performance. Vibration stress is relieved as a result of rotating rather than reciprocating parts. Control is simpler because one lever controls both speed and power. With the large airflow, cooling is less complicated. Spark plugs are used only for starting, and the continuous ignition system of reciprocating engines is not needed. A carburetor and mixture control are not needed. The major disadvantages have been the high fuel consumption and poor performance at low power setting, low speeds, and low altitudes. Turboprop and turbofan developments have greatly improved aircraft turbine engines in these areas.

Section 6.7 - RAMJET ENGINES


The ramjet engine is the simplest type of the all-jet engines because it has no moving parts. Figure 6-10 shows a typical arrangement of the parts of a ramjet engine. Note that it may have an internal body that serves to compress the air as it enters the intake.

The spray bar injects a mist of fuel into the airstream and the mixture is ignited by a spark. The grill-type flame holder provides a type of barrier to the burning mixture while allowing hot, expanding gases to escape through the exhaust nozzle. The high-pressure air coming into the combustion chamber keeps the burning mixture from effectively reacting toward the intake end of the engine.

Ramjets will not function until enough air is coming through the intake to create a highpressure flow. Otherwise, the expanding gases of the burning fuel-air mixture would be expelled from both ends of the engine. As you can see, this would amount to a single explosive reaction. Therefore, the ramjet has to be traveling through the air very fast before it is started. This means that it has to be boosted to the proper speed by some other type of engine. In theory, the ramjet engine has no maximum speed; it can keep accelerating indefinitely as long as it stays within the atmosphere. In practice, the ramjet is limited, at this time, to low hypersonic speeds (five times the speed of sound) because atmospheric friction will melt it. The biggest drawback of the ramjet is its high rate of fuel consumption.

You might also like