You are on page 1of 12

Lift (force)

From Wikipedia, the free encyclopedia

Jump to: navigation, search

Boeing 747-8F landing A fluid flowing past the surface of a body exerts a surface force on it. Lift is defined to be the component of this force that is perpendicular to the oncoming flow direction.[1] It contrasts with the drag force, which is defined to be the component of the surface force parallel to the flow direction.

Contents

1 Overview 2 Description of lift on an airfoil o 2.1 Newton's laws: lift and the deflection of the flow 2.1.1 Deflection 2.1.2 Flow turning 2.1.3 Pressure differences 2.1.4 Criticisms of deflection/turning o 2.2 A more rigorous physical description 2.2.1 Lift in an established flow 2.2.2 Flowfield formation o 2.3 Other alternative explanations for the generation of lift 2.3.1 "Popular" explanation based on equal transit-time 2.3.2 Coand effect 2.3.3 In terms of a difference in areas 3 Methods to determine lift on an airfoil o 3.1 Lift coefficient o 3.2 KuttaJoukowski theorem o 3.3 Pressure integration 4 Lift forces on bluff bodies 5 See also 6 References and notes

7 Further reading 8 External links

Overview

Forces on an airfoil. If the fluid is air, the force is called an aerodynamic force. An airfoil is a streamlined shape that is capable of generating significantly more lift than drag.[2] Aerodynamic lift is commonly associated with the wing of a fixed-wing aircraft, although lift is also generated by propellers; kites; helicopter rotors; rudders, sails and keels on sailboats; hydrofoils; wings on auto racing cars; wind turbines and other streamlined objects. While the common meaning of the word "lift" assumes that lift opposes gravity, lift in its specialized technical sense obviously can be in any direction with respect to gravity, since it is defined with respect to the direction of flow, rather than to the direction of gravity. When an aircraft is flying straight and level (cruise) most[3] of the lift opposes gravity. However, when an aircraft is climbing, descending, or banking in a turn, for example, the lift is tilted with respect to the vertical.[4] Lift may also be entirely downwards in some aerobatic manoeuvres, or on the wing on a racing car. In this last case, the term downforce is often used. Lift may also be horizontal, for instance on a sail on a sailboat. Non-streamlined objects such as bluff bodies and plates (not parallel to the flow) may also generate lift when moving relative to the fluid. This lift may be steady, or it may oscillate due to vortex shedding. Interaction of the object's flexibility with the vortex shedding may enhance the effects of fluctuating lift and cause vortex-induced vibrations.[5]

Description of lift on an airfoil


There are several ways to explain how an airfoil generates lift. Some are more complicated or more mathematically rigorous than others; some have been shown to be incorrect.[6] This article will start with the simplest explanation; more complicated and alternative explanations will follow.

Newton's laws: lift and the deflection of the flow


Deflection

Airstreams around an airfoil in a wind tunnel. Note the curved streamlines and the overall downward deflection of the air. One way to understand the generation of lift is to observe that the air is deflected as it passes the airfoil. Since the airfoil must exert a force on the air to change its direction, the air must exert a force of equal magnitude but opposite direction on the airfoil. In the case of an airplane wing, the wing exerts a downward force on the air and the air exerts an upward force on the wing.[8][10][11][12] This explanation relies on the second and third of Newton's laws of motion: The net force on an object is equal to its rate of momentum change, and: To every action there is an equal and opposite reaction.[13] Flow turning Another way to describe deflection is to say that the air "turns" as it passes the airfoil and follows a path that is curved. When airflow changes direction, a force is generated.[14] Pressure differences It may also be described in terms of air pressure: pressure is just force per unit area perpendicular to that area. So, wherever there is force there is also a pressure difference.[15] Since deflection/flow turning tells us that there is a force, it also tells us there is a pressure difference.[16][17] This pressure difference implies higher pressure on the underside of the wing and lower pressure on the upper side. Criticisms of deflection/turning

While the theory correctly reasons that deflection implies that there must be a force on the wing, it does not explain why the air is deflected. Intuitively, one can say that the air follows the curve of the foil,[18] but this is not very rigorous or precise. The theory, while correct in as far as it goes, is not sufficient to allow one to do engineering. Fluid stresses including pressure need to be related to the fluid motion (e.g. through constitutive equations). Thus, textbooks on aerodynamics use more complex models to provide a full description of lift.

A more rigorous physical description

Flow around an airfoil: the dots move with the flow. Note that the velocities are much higher at the upper surface than at the lower surface. The black dots are on timelines, which split into two an upper and lower part at the leading edge. The part of a timeline below the airfoil does not catch up with the one above. Colors of the dots indicate streamlines. The airfoil is a Krmn Trefftz airfoil, with parameters x = 0.08, y = +0.08 and n = 1.94. The angle of attack is 8, and the flow is a potential flow. Lift is generated in accordance with the fundamental principles of physics. The most relevant physics reduce to three principles:

Newton's laws of motion, especially Newton's second law which relates the net force on an element of air to its rate of momentum change, conservation of mass, including the common assumption that the airfoil's surface is impermeable for the air flowing around, and an expression relating the fluid stresses (consisting of pressure and shear stress components) to the properties of the flow.[19]

In the last principle, the pressure depends on the other flow properties, such as its mass density, through the (thermodynamic) equation of state, while the shear stresses are related to the flow through the air's viscosity.[19] Application of the viscous shear stresses to Newton's second law for an airflow results in the NavierStokes equations. But in many instances approximations suffice for a good description of lifting airfoils: in large parts of the flow viscosity may be neglected. Such an inviscid flow can be described mathematically through the Euler equations, resulting from the Navier-Stokes equations when the viscosity is neglected. The Euler equations for a steady and inviscid flow can be integrated along a streamline, resulting in Bernoulli's equation. The particular form of Bernoulli's equation found depends on the equation of state used.[20] At low Mach numbers, compressibility effects may be neglected, resulting in an incompressible flow approximation. In incompressible and inviscid flow the Bernoulli equation is just an integration of Newton's second lawin the form of the description of momentum evolution by the Euler equationsalong a streamline.[20] Explaining lift while considering all of the principles involved is a complex task and is not easily simplified.[7][21] As a result, there are numerous different explanations of lift with different levels of rigour and complexity. For example, there is an explanation based directly on Newtons laws of motion and an explanation based on Bernoullis principle. Neither of these explanations is incorrect, but each appeals to a different audience.[22]

In order to explain lift as it applies to an airplane wing, consider the incompressible flow around a 2-D, symmetric airfoil at positive angle of attack in a uniform freestream. Instead of considering the case where an airfoil moves through a fluid as seen by a stationary observer, it is equivalent and simpler to consider the picture when the observer follows the airfoil and the fluid moves past it. Lift in an established flow

Streamlines around a NACA 0012 airfoil at moderate angle of attack. If one takes the experimentally observed flow around an airfoil as a starting point, then lift can be explained in terms of pressures using Bernoulli's principle (which can be derived from Newton's second law) and conservation of mass.[7] The image to the right shows the streamlines over a NACA 0012 airfoil computed using potential flow theory, a simplified model of the real flow. The flow approaching an airfoil can be divided into two streamtubes, which are defined based on the area between two streamlines. By definition, fluid never crosses a streamline in a steady flow; hence mass is conserved within each streamtube. One streamtube travels over the upper surface, while the other travels over the lower surface; dividing these two tubes is a dividing line (the stagnation streamline) that intersects the airfoil on the lower surface, typically near to the leading edge. The stagnation streamline leaves the airfoil at the sharp trailing edge, a feature of the flow known as the Kutta condition. In calculating the flow shown, the Kutta condition was imposed as an initial assumption; the justification for this assumption is explained below. The upper stream tube constricts as it flows up and around the airfoil, a part of the so-called upwash. From the conservation of mass, the flow speed must increase as the stream tube area decreases.[7] The area of the lower stream tube increases, causing the flow inside the tube to slow down. It is typically the case that the air parcels traveling over the upper surface will reach the trailing edge before those traveling over the bottom. From Bernoulli's principle, the pressure on the upper surface where the flow is moving faster is lower than the pressure on the lower surface. The pressure difference thus creates a net aerodynamic force, pointing upward and downstream to the flow direction. The component of the force normal to the freestream is considered to be lift; the component parallel to the freestream is drag. In conjunction with this force by the air on the airfoil, by Newton's third law, the airfoil imparts an equal-and-opposite force on the surrounding air that creates the downwash.

Measuring the momentum transferred to the downwash is another way to determine the amount of lift on the airfoil.[23] Flowfield formation In attempting to explain why the flow follows the upper surface of the airfoil, the situation gets considerably more complex. It is here that many simplifications are made in presenting lift to various audiences, some of which are explained after this section. Consider the case of an airfoil accelerating from rest in a viscous flow. Lift depends entirely on the nature of viscous flow past certain bodies:[24] in inviscid flow (i.e. assuming that viscous forces are negligible in comparison to inertial forces), there is no lift without imposing a net circulation, the proper amount of which can be determined by applying the Kutta condition. In a viscous flow like in the physical world, however, the lift and other properties arise naturally as described here. When there is no flow, there is no lift and the forces acting on the airfoil are zero. At the instant when the flow is turned on, the flow is undeflected downstream of the airfoil and there are two stagnation points on the airfoil (where the flow velocity is zero): one near the leading edge on the bottom surface, and another on the upper surface near the trailing edge. The dividing line between the upper and lower streamtubes mentioned above intersects the body at the stagnation points. Since the flow speed is zero at these points, by Bernoulli's principle the static pressure at these points is at a maximum. As long as the second stagnation point is at its initial location on the upper surface of the wing, the circulation around the airfoil is zero and, in accordance with the KuttaJoukowski theorem, there is no lift. The net pressure difference between the upper and lower surfaces is zero. The effects of viscosity are contained within a thin layer of fluid called the boundary layer, close to the body. As flow over the airfoil commences, the flow along the lower surface turns at the sharp trailing edge and flows along the upper surface towards the upper stagnation point. The flow in the vicinity of the sharp trailing edge is very fast and the resulting viscous forces cause the boundary layer to accumulate into a vortex on the upper side of the airfoil between the trailing edge and the upper stagnation point.[25] This is called the starting vortex. The starting vortex and the bound vortex around the surface of the wing are two halves of a closed loop. As the starting vortex increases in strength the bound vortex also strengthens, causing the flow over the upper surface of the airfoil to accelerate and drive the upper stagnation point towards the sharp trailing edge. As this happens, the starting vortex is shed into the wake,[26] and is a necessary condition to produce lift on an airfoil. If the flow were stopped, there would be a corresponding "stopping vortex".[27] Despite being an idealization of the real world, the vortex system set up around a wing is both real and observable; the trailing vortex sheet most noticeably rolls up into wing-tip vortices. The upper stagnation point continues moving downstream until it is coincident with the sharp trailing edge (as stated by the Kutta condition). The flow downstream of the airfoil is deflected downward from the free-stream direction and, from the reasoning above in the basic explanation,

there is now a net pressure difference between the upper and lower surfaces and an aerodynamic force is generated.

Other alternative explanations for the generation of lift


It is amazing that today, almost 100 years after the first flight of the Wright Flyer, groups of engineers, scientists, pilots, and others can gather together and have a spirited debate on how an airplane wing generates lift. Various explanations are put forth, and the debate centers on which explanation is the most fundamental. John D. Anderson, Curator of Aerodynamics at the National Air and Space Museum[7] Many other alternative explanations for the generation of lift by an airfoil have been put forward, of which a few are presented here. Most of them are intended to explain the phenomenon of lift to a general audience. Although the explanations may share features in common with the explanation above, additional assumptions and simplifications may be introduced. This can reduce the validity of an alternative explanation to a limited sub-class of lift generating conditions, or might not allow a quantitative analysis. Several theories introduce assumptions which proved to be wrong, like the equal transit-time theory. "Popular" explanation based on equal transit-time

An illustration of the (incorrect) equal transit-time theory. An explanation of lift frequently encountered in basic or popular sources is the equal transit-time theory. Equal transit-time states that because of the longer path of the upper surface of an airfoil, the air going over the top must go faster in order to catch up with the air flowing around the bottom,[28] i.e. the parcels of air that are divided at the leading edge and travel above and below an airfoil must rejoin when they reach the trailing edge. Bernoulli's Principle is then cited to conclude that since the air moves faster on the top of the wing the air pressure must be lower. This pressure difference pushes the wing up. However, equal transit time is not accurate[29] and the fact that this is not generally the case can be readily observed.[30] Although it is true that the air moving over the top of a wing generating lift does move faster, there is no requirement for equal transit time. In fact the air moving over

the top of an airfoil generating lift is always moving much faster than the equal transit theory would imply.[6] The assertion that the air must arrive simultaneously at the trailing edge is sometimes referred to as the "Equal Transit-Time Fallacy".[31][32][33][34] Note that while this theory depends on Bernoulli's principle, the fact that this theory has been discredited does not imply that Bernoulli's principle is incorrect. Coand effect Main article: Coand effect In a limited sense, the Coand effect refers to the tendency of a fluid jet to stay attached to an adjacent surface that curves away from the flow, and the resultant entrainment of ambient air into the flow. The effect is named for Henri Coand, the Romanian aerodynamicist who exploited it in many of his patents. One of the first known uses was in his patent for a high-lift device[35] that used a fan of gas exiting at high speed from an internal compressor. This circular spray was directed radially over the top of a curved surface shaped like a lens to decrease the pressure on that surface. The total lift for the device was caused by the difference between this pressure and that on the bottom of the craft. Two aircraft, the Antonov An-72 and An-74 "Coaler", use the exhaust from topmounted jet engines flowing over the wing to enhance lift,[36] as did the Boeing YC-14 and the McDonnell Douglas YC-15.[37] [38] The effect is also used in high-lift devices such as a blown flap.[39] More broadly, some consider the effect to include the tendency of any fluid boundary layer to adhere to a curved surface, not just the boundary layer accompanying a fluid jet. It is in this broader sense that the Coand effect is used by some to explain lift.[40] Jef Raskin,[41] for example, describes a simple demonstration, using a straw to blow over the upper surface of a wing. The wing deflects upwards, thus supposedly demonstrating that the Coand effect creates lift. This demonstration correctly demonstrates the Coand effect as a fluid jet (the exhaust from a straw) adhering to a curved surface (the wing). However, the upper surface in this flow is a complicated, vortex-laden mixing layer, while on the lower surface the flow is quiescent. The physics of this demonstration are very different from that of the general flow over the wing.[42] The usage in this sense is encountered in some popular references on aerodynamics.[40][41] In the aerodynamics field, the Coand effect is commonly defined in the more limited sense above[42][43][44] and viscosity is used to explain why the boundary layer attaches to the surface of a wing.[27] In terms of a difference in areas When a fluid flows relative to a solid body, the body obstructs the flow, causing some of the fluid to change its speed and direction in order to flow around the body. The obstructive nature of the solid body causes the streamlines to move closer together in some places, and further apart

in others.[7][45][46] When fluid flows past a 2-D cambered airfoil at zero angle of attack, the upper surface has a greater area (that is, the interior area of the airfoil above the chordline) than the lower surface and hence presents a greater obstruction to the fluid than the lower surface.[45] This asymmetry causes the streamlines in the fluid flowing over the upper surface to move closer together than the streamlines over the lower surface. As a consequence of mass conservation, the reduced area between the streamlines over the upper surface results in a higher velocity than that over the lower surface. The upper streamtube is squashed the most in the nose region ahead of the maximum thickness of the airfoil, causing the maximum velocity to occur ahead of the maximum thickness.[7] In accordance with Bernoulli's principle, where the fluid is moving faster the pressure is lower, and where the fluid is moving slower the pressure is greater. The fluid is moving faster over the upper surface, particularly near the leading edge, than over the lower surface so the pressure on the upper surface is lower than the pressure on the lower surface. The difference in pressure between the upper and lower surfaces results in lift.[7]

Methods to determine lift on an airfoil


Lift coefficient
Main article: Lift coefficient If the lift coefficient for a wing at a specified angle of attack is known (or estimated using a method such as thin-airfoil theory), then the lift produced for specific flow conditions can be determined using the following equation:[47]

where L is lift force, is air density v is true airspeed, A is planform area, and CL is the lift coefficient at the desired angle of attack, Mach number, and Reynolds number[48]

This equation is basically the same as the drag equation, only the lift/drag coefficient is different.

KuttaJoukowski theorem
Main article: KuttaJoukowski theorem

Lift can be calculated using potential flow theory by imposing a circulation. It is often used by practising aerodynamicists as a convenient quantity in calculations, for example thinairfoil theory and lifting-line theory. The circulation is the line integral of the velocity of the air, in a closed loop around the boundary of an airfoil. It can be understood as the total amount of "spinning" (or vorticity) of air around the airfoil. The section lift/span L' can be calculated using the KuttaJoukowski theorem:[23]

where is the air density, V is the free-stream airspeed. Kelvin's circulation theorem states that circulation is conserved.[49] There is conservation of the air's angular momentum. When an aircraft is at rest, there is no circulation. The challenge when using the KuttaJoukowski theorem to determine lift is to determine the appropriate circulation for a particular airfoil. In practice, this is done by applying the Kutta condition, which uniquely prescribes the circulation for a given geometry and freestream velocity. A physical understanding of the theorem can be observed in the Magnus effect, which is a lift force generated by a spinning cylinder in a freestream. Here the necessary circulation is induced by the mechanical rotation acting on the boundary layer, causing it to induce a faster flow around one side of the cylinder and a slower flow around the other. The asymmetric distribution of airspeed around the cylinder then produces a circulation in the outer inviscid flow.[50]

Pressure integration
The force on the wing can be examined in terms of the pressure differences above and below the wing, which can be related to velocity changes by Bernoulli's principle. The total lift force is the integral of vertical pressure forces over the entire wetted surface area of the wing:[51]

where:

L is the lift, A is the wing surface area p is the value of the pressure, n is the normal unit vector pointing into the wing, and k is the vertical unit vector, normal to the freestream direction.

The above lift equation neglects the skin friction forces, which typically have a negligible contribution to the lift compared to the pressure forces. By using the streamwise vector i parallel to the freestream in place of k in the integral, we obtain an expression for the pressure drag Dp (which includes induced drag in a 3D wing). If we use the spanwise vector j, we obtain the side force Y.

One method for calculating the pressure is Bernoulli's equation, which is the mathematical expression of Bernoulli's principle. This method ignores the effects of viscosity, which can be important in the boundary layer and to predict friction drag, which is the other component of the total drag in addition to Dp. The Bernoulli principle states that the sum total of energy within a parcel of fluid remains constant as long as no energy is added or removed. It is a statement of the principle of the conservation of energy applied to flowing fluids. A substantial simplification of this proposes that as other forms of energy changes are inconsequential during the flow of air around a wing and that energy transfer in/out of the air is not significant, then the sum of pressure energy and speed energy for any particular parcel of air must be constant. Consequently, an increase in speed must be accompanied by a decrease in pressure and vice-versa. It should be noted that this is not a causational relationship. Rather, it is a coincidental relationship, whatever causes one must also cause the other as energy can neither be created nor destroyed. It is named for the Dutch-Swiss mathematician and scientist Daniel Bernoulli, though it was previously understood by Leonhard Euler and others. Bernoulli's principle provides an explanation of pressure difference in the absence of air density and temperature variation (a common approximation for low-speed aircraft). If the air density and temperature are the same above and below a wing, a naive application of the ideal gas law requires that the pressure also be the same. Bernoulli's principle, by including air velocity, explains this pressure difference. The principle does not, however, specify the air velocity. This must come from another source, e.g., experimental data. In order to solve for the velocity of inviscid flow around a wing, the Kutta condition must be applied to simulate the effects of viscosity. The Kutta condition allows for the correct choice among an infinite number of flow solutions that otherwise obey the laws of conservation of mass and conservation of momentum.

Lift forces on bluff bodies


Further information: Vortex shedding and Vortex-induced vibration The flow around bluff bodies may also generate lift, besides a strong drag force. For instance, the flow around a circular cylinder generates a Krmn vortex street: vortices being shed in an alternating fashion from each side of the cylinder. The oscillatory nature of the flow is reflected in the fluctuating lift force on the cylinder, whereas the mean lift force is negligible. The lift force frequency is characterised by the dimensionless Strouhal number, which depends (among others) on the Reynolds number of the flow.[52][53] For a flexible structure, this oscillatory lift force may induce vortex-induced vibrations. Under certain conditions for instance resonance or strong spanwise correlation of the lift force the resulting motion of the structure due to the lift fluctuations may be strongly enhanced. Such vibrations may pose problems, even collapse, in man-made tall structures like for instance industrial chimneys, if not properly taken care of in the design.[5]

You might also like