Sie sind auf Seite 1von 58

Review: Fundamentals of Fluid Flow

AA200B Lecture 2 September 27, 2007

AA200B - Applied Aerodynamics II

AA200B - Applied Aerodynamics II

Lecture 2

Fundamentals of Fluid Flow


This chapter serves as a review of some of the fundamental concepts of uid dynamics, including the origin of pressure and shear forces, the basic equations of uid dynamics and their underlying approximations. It is not intended as a comprehensive introduction to uid dynamics, but as a supplement to conventional texts such as those cited in the references section. Applied aerodynamics is concerned with the measurement or prediction of aerodynamic properties. These may include any of the physical properties that are associated with the uid such as: pressure, temperature, density, velocity, other gas properties, and forces on bodies. Much of the work in applied aerodynamics is concerned with the prediction of these forces. This can be done in several ways, starting with
2

AA200B - Applied Aerodynamics II

Lecture 2

very simple considerations and moving on to the detailed eld equations. The chapter is divided into several sections: 1. Origin of Forces 2. Dimensionless Groups 3. Conservation Equations 4. Common Approximations 5. Fluid Flow Equations 6. Relating Pressure and Veloicity 7. References
3

AA200B - Applied Aerodynamics II

Lecture 2

Origin of Fluid Forces


We are often particularly interested in the forces and moments applied to bodies moving through the uid. These can be divided into just two types: pressures and shears. Pressures are created at the surface of a body due to (nearly) elastic collisions between molecules of the uid and the surface of the body. Shearing forces are produced by uid viscosity. This quantity is a measure of how well momentum is transferred between adjacent layers of the uid. Although both types of forces are important in applied aerodynamics, pressures are usually the dominant type of force.

AA200B - Applied Aerodynamics II

Lecture 2

Each of the shapes below, drawn to scale, have the same drag. The reason that the streamlined airfoil can be so much larger is that most of the force is due to shear stress, not pressure forces.

The cylinder, with its separated airow, has large pressure forces that give rise to high drag.

AA200B - Applied Aerodynamics II

Lecture 2

In fact, it is quite amazing how much force can be generated by dierences in pressure: Many airliners have wing loadings (weight / wing area) of over 100 lbs/ft2 (4.8 KPa). This means that it takes a section of wing only as large as a book to lift a large dog (for example).

This is possible because the normal atmospheric pressure is 2116 lb/ft2 (or 101 KPa) at sea level. So, in fact 100 psf (4.8 KPa) represents only a 5% change in the pressure on the upper side of the wing. At 68,000 ft (20.7 km) you would have to create a complete vacuum on the wing upper surface to lift that much weight.

AA200B - Applied Aerodynamics II

Lecture 2

Origin of Pressure Forces Pressure arises because each molecule that bounces o the surface transfers momentum to the body. If a particle of mass, m, hits the body straight-on and bounces o, it transfers momentum of the amount 2mc where c is the speed of the molecule. The pressure is then proportional to the number of molecules striking a unit area of the surface per unit time, (Number density*c), times the momentum transfer per particle, ( mc) or: p = k1c2.

Note that since temperature is dened as proportional to the mean kinetic energy of the molecules, T = k2c2. So we expect: p = kT , the perfect gas relation.
7

AA200B - Applied Aerodynamics II

Lecture 2

The component of molecular velocity normal to the surface is what is really needed in the above expression, and if the body is moving, we must add its velocity to the molecular velocity measured in the uid-xed reference frame. Typically, we do not consider these direct interactions, but rather model the molecules as a continuous uid. This works well for most ows of interest. However, for very rareed ows such as those associated with initial re-entry of space vehicles, it is sometimes possible to analyze aerodynamics with kinetic theory, keeping track of the molecular interactions. The gure below shows the results of one such calculation.

Image courtesy of Prof. D. Bagano, Stanford University


8

AA200B - Applied Aerodynamics II

Lecture 2

Dimensionless Groups
The forces on a body, moving through a uid, depend on the body velocity (V), the uid density (r), temperature, and viscosity (m), the size of the body (l), and its shape. Using the speed of sound, a, rather than temperature (they are directly related) we can then make the following table that shows the units associated with each of the parameters. Here the numbers indicate the power to which the mass, length, or time units are raised: Dimension Mass Length Time F = f( 1 1 -2 V 0 1 -1 1 -3 0 a 0 1 -1 1 -1 -1 l 0 1 0 shape) 0 0 0

AA200B - Applied Aerodynamics II

Lecture 2

The Buckingham pi theorem states that the number of dimensionless parameters is equal to the number of parameters minus the rank of the above matrix. In this case 7 - 3 = 4. So, there exists a functional relationship among the four dimensionless groups. We can express the force on a body, for instance, by a relationship between the following four dimensionless parameters: 1. Dimensionless force coecient: 2. Reynolds Number: 3. Mach Number: 4. Geometry
10

F V 2 l2

V l

V a

AA200B - Applied Aerodynamics II

Lecture 2

The relation is:

F V l V = f( , , shape) V 2l2 a

We will discuss each of these dimensionless groups in a moment, but lets rst look at the functional relationship between them. Much of applied aerodynamics involves nding the function f, but there is a great deal we can say, even without knowing it. For example, we can see that a wide variety of similar ows exist. The forces on a large, slow-moving body could be predicted from tests of a small higher-speed model as long as the speed of sound were suciently high. Also, the ow around a small insect could be represented by a large model in a very viscous uid. The idea behind model testing is to simulate the ow over one body by matching the dimensionless parameters of another.

11

AA200B - Applied Aerodynamics II

Lecture 2

This is not always easy or possible. The following gure, from J. McMasters of Boeing shows the Mach and Reynolds number range of several wind tunnels. Why cant wind tunnels be designed to more fully cover this range of parameters? What alternatives exist to wind tunnel tests?

12

AA200B - Applied Aerodynamics II

Lecture 2

Subsequent pages consider each of the dimensionless groups in a bit more detail. First note that we could have included other uid properties such as specic heats. This would lead to additional dimensionless parameters such as the Prandtl number which is important in the study of compressible boundary layers with heat conduction. We have also left out gravity which is often important in the ow of water around ships. This would lead to an additional dimensionless parameter called the Froude number. There are often several ways of combining the parameters to form dimensionless groups, but these are commonly used in aerodynamics.

13

AA200B - Applied Aerodynamics II

Lecture 2

Dimensionless Forces We use dimensionless force and moment coecients dened by: L = 0.5V 2SCL(Re, M, shape) M = 0.5V 2ScCm(Re, M, shape) CL is called the lift coecient, Cm the moment coecient. The length2 term in our rst dimensionless parameter has been replaced by the area, S . This area could be anything we choose (the contact area of the nose wheel, the wing planform area, the fuselage cross-sectional area). In a particular application people generally agree on a reference area. For car drag coecients the frontal area is often used. For aircraft the wing area is a common reference area. The c (for chord) in the moment coecient denition is similarly agreed upon. This agreement on reference area
14

AA200B - Applied Aerodynamics II

Lecture 2

is very important as can be seen in advertisements for cars. (Automobile drag coecients are usually based on frontal area and numbers like 0.4 are sometimes mentioned in car ads. But, the drag coecient means nothing by itself. If we chose the reference area to be the oor area of the Fremont GM plant, we would have very low drag coecients.) According to wikipedia: In 2003, Car and Driver adapted this metric the drag area, CD S and adopted it as a more intuitive way to compare the aerodynamic eciency of various automobiles. Average full-size passenger cars have a drag area of roughly 8.5f t2 (.79m2). Reported drag area ranges from the 1999 Honda Insight at 5.1f t2 (.47m2) to the 2003 Hummer H2 at 26.3f t2 (2.44m2).

15

AA200B - Applied Aerodynamics II

Lecture 2

Reynolds Number The quantity:


V l

is called the Reynolds number.

is the uid density, V is the speed, is the uid viscosity, and l is some characteristic length. This length is, like the areas in the denition of dimensionless force coecients, agreed on as a standard by whoever is using it. So, chord Reynolds numbers are based on wing chord lengths; other Reynolds numbers are based on the diameter of a sphere, or any other characteristic length that can be devised. The Reynolds number is one of the most important and strange dimensionless numbers. It varies over many orders of magnitude and expresses the importance of viscosity: high Reynolds numbers can be achieved by decreasing the viscosity or making the length or speed very large.

16

AA200B - Applied Aerodynamics II

Lecture 2

The Reynolds number, in a sense, represents a ratio of pressure to shear forces: V l/ = V 2/(V /l) V 2 is related to the pressure while V /l is related to dU/dy , the shear stress.

The range of Reynolds number from McMasters.


17

AA200B - Applied Aerodynamics II

Lecture 2

Viscosity, and hence Reynolds number, strongly aects the performance of wings and airfoils, making it an important parameter to match in wind tunnel tests. It is often not possible to match these dimensionless parameters precisely. The plot below shows the eect of Reynolds number on maximum lift to drag ratio for two dimensional airfoil sections. Note the plight of insects.

18

AA200B - Applied Aerodynamics II

Lecture 2

The plot here shows the eect of Reynolds number on the maximum section lift coecient of a few typical airfoil sections. Note that these are not necessarily the best sections for high lift, though. Recent studies have shown that substantial changes in CLmax are seen even at quite high Reynolds numbers, making it dicult to extrapolate data on small wind tunnel models.

19

AA200B - Applied Aerodynamics II

Lecture 2

Mach Number The Mach number is the ratio of ow speed, V, to the speed of sound, a: M = V a . It reects the importance of the compressibility of the uid. This ratio is important because pressure disturbances propagate in a uid at the local speed of sound and the compressibility of a uid permits a sound wave to travel. The speed of sound in a uid dp is related to the way in which density and pressure vary: a2 = d . Assuming isentropic ow and a perfect gas: a2 = RT with R the f t lb Nm gas constant R = 287.05 kg K (1718 sl R ) and 1.4 for air at standard conditions. The ow pattern and pressures can change dramatically with Mach number as the applicable dierential equation changes form. (See later sections.) At low subsonic speeds, the eect of compressibility is not large and the ow behaves almost as if pressure disturbances traveled with innite speed. As the ow speed is increased, but
20

AA200B - Applied Aerodynamics II

Lecture 2

remains subsonic the eects of compressibility start to appear slowly. (The connection between compressibility eects and speed is just related to the fact that as the speed is increased, the pressure changes are a suciently large fraction of the ambient pressure, that density changes are also signicant.) As the ow velocity approaches Mach 1 (transonic ow) more signicant compressibility eects appear quickly. Because of the increase in local velocity over parts of an airfoil, the local Mach number can be much higher than the freestream Mach number. In fact, compressibility eects can be important for high-lift sections at freestream Mach numbers as low as 0.3. As the ow velocity increases beyond Mach 1.0, it becomes supersonic and its characteristics change greatly. Very high velocity ows (usually above Mach 5 or 6) are called hypersonic. These types of ow are of great importance in the aerodynamics of rockets and re-entry vehicles, which achieve Mach numbers as high as 25.

21

AA200B - Applied Aerodynamics II

Lecture 2

Conservation Laws
To derive the equations of motion for uid particles we rely on various conservation principles. These principles are entirely intuitive. They are a statement of the fact that the rate of change of mass, momentum, or energy in a certain volume is equal to the rate at which it enters the borders of the volume plus the rate at which it is created inside. The rst two of these will be used extensively here.
S V , V , p n

The integral expressions for the basic conservation laws are written in terms of velocity, density, pressure, and force and internal energy per unit mass over an arbitrary volume, enclosed by a surface, S.
22

AA200B - Applied Aerodynamics II

Lecture 2

Continuity (mass conservation): t Momentum: t Energy: t V2 (e + ) dV = 2 V V2 (e + )(V n ) + pn V dS 2 S V dV =


V S

dV =
V S

V n dS

V (V n ) + pn dS +
V

f dV

23

AA200B - Applied Aerodynamics II

Lecture 2

These integral expressions are combined with the divergence theorem: F dV =


V S

F n dS

and the fact that they hold over arbitrary volumes to obtain the dierential form of the equations: Continuity: Momentum: + V = 0 t V + (V )V + p = f t This can also be written in component form. In the xi direction: p Vi + ( V )Vi + = fi t xi
24

AA200B - Applied Aerodynamics II

Lecture 2

Far-Field Forces We can use the momentum theorem by itself to obtain useful results. In this example, we apply the momentum theorem to relate the force on a body to the properties of the ow some distance from the body. This technique is useful in wind tunnel tests and is the basis of several fundamental theorems related to lift and induced drag of wings. We take the control volume shown in the following gure, bounded by the single surface, S which we divide into 3 parts: the outer surface (Souter ), the inner surface (Sinner ), and the pieces of the surface connecting the two (S ). We can write the integral form of the momentum equation for steady ow with no body forces as: V (V n ) + pn dS = 0
S
25

AA200B - Applied Aerodynamics II

Lecture 2

Souter

Sinner

Or, in terms of the three pieces of the surface: V (V n ) + pn dS +


Souter Sinner

V (V n ) + pn dS = 0

(Note that the contribution from the part of the surface connecting Sinner and Souter to the integrals is zero because as the two pieces of S are made close together, the unit normals point in opposite directions while p and V are equal. )
26

AA200B - Applied Aerodynamics II

Lecture 2

If we shrink the inner part of the surface down until it touches the body, then over this piece, V n = 0 (the ow if tangent to the body). Then, we can nd the resulting net pressure force, F , on the body by noting that: pn dS = F
Sinner

So: F =
Souter

V (V n ) + pn dS

That is, we can determine the force on the body, just from properties far away from the body.

27

AA200B - Applied Aerodynamics II

Lecture 2

Approximations
The equations of motion for a general uid are extremely complex and even if the problem could be formulated it would be impractical to solve. Thus, from the outset, certain simplifying approximations that are often very accurate, are made. These may include the following assumptions. Continuity and Homogeneity : We assume that the uid is composed of particles which are so small and plentiful that the statisticallyaveraged properties of interest are the same at any scale. This works well for gases and uids under most conditions. It does not work for studying the ow of sand. It does not work when the uid is so rareed that the mean free path is of the same order as the dimensions of interest in the problem. The mean free path varies with altitude as shown in the plot. We further assume that the medium can be treated as a single type of uid no suspensions of oil and water.
28

AA200B - Applied Aerodynamics II

Lecture 2

Inviscid : The eect of viscosity may sometimes be neglected or modeled indirectly. For many aerodynamic ows of interest, the region of high shear and vorticity is conned to a thin layer of uid. Outside this layer, the uid behaves as if it were inviscid. Thus the simpler equations of an inviscid uid are often solved outside of the shear layers. There are some uids which seem to be almost completely inviscid. Tests in superuid helium have given results similar to inviscid calculations. Incompressible (constant density): When the uid density does not change with changes in pressure, the uid is incompressible. Water density changes very little with changes in pressure and is generally treated as an incompressible uid. Air is compressible, but if pressure changes are small in comparison with some nominal value, the corresponding changes in density are small also and incompressible equations work quite well in describing the ow. The degree to which
29

AA200B - Applied Aerodynamics II

Lecture 2

the uid density changes with pressure is related to the speed of sound in the uid. Thus, assuming that the ow is incompressible is equivalent to assuming that the speed of sound is innite. When the local Mach number is less than 0.2 to 0.5 compressibility eects can often be ignored. The reason for this is discussed further in the chapter on compressibility, but one can see qualitatively that in order to make an appreciable change to the nominal 2116 lb/f t2 air pressure at sea level, substantial speeds are required. Irrotational : Circulation is dened as: = V ds It is a measure of the rotation of an area of uid. As the integration contour is shrunk down to a point, the ratio of circulation to the area enclosed by the curve is called the vorticity. = lim =V S 0 S
30

AA200B - Applied Aerodynamics II

Lecture 2

Fluid that starts out without rotational motion will not develop it unless there has been some shear stress acting on it. And if the shear is conned to a small region, the vorticity will be also. Thus, for many cases, especially in inviscid ow, much of the ow eld may be treated as irrotational: V = 0. When this is the case, the vector eld, V , may be written as the gradient of a scalar eld, : V = . where is the velocity potential. This simplies many of the equations discussed in subsequent sections. The velocity components are then: u = and v = x y . Some important exceptions to the idea that without viscosity irrotational ow remains irrotational: Vorticity can be created in a gravitational eld when density gradients exist or in a rotating system (such as the earth) due to Coriolus forces. These are important sources of vorticity in meteorology.

31

AA200B - Applied Aerodynamics II

Lecture 2

Steady :When the variables describing the uid properties at a given point do not change in time, the ow may be treated as steady and the time derivatives in the equations of motion are zero. This condition depends on the chosen coordinate system. If the system is at rest with respect to a body in uniform motion through a uid the equations in that system are steady, but expressed in a system xed with respect to the undisturbed uid, the ow is unsteady. It is often convenient to transform the coordinate system to one in which the ow is steady. This is, of course, not always possible. We will assume that the ow is steady in most of the discussions in this course but unsteady eects are often important in the study of bird ight, propellers, aircraft gust response, dynamics, and aeroelasticity as well as in the study of turbulence.

32

AA200B - Applied Aerodynamics II

Lecture 2

Equations of Fluid Flow


The conservation laws may be used to derive the equations of uid ow. These are supplemented with constitutive relations such as the perfect gas law: p = RT or the isentropic relation between pressure and density: p 2 2 = p 1 1 Some of the most commonly-solved equations are shown in the following table along with the corresponding assumptions.

33

AA200B - Applied Aerodynamics II Equation Navier-Stokes Reynolds-Averaged Navier-Stokes Euler Full Potential Transonic Small Disturbance Prandtl-Glauert Acoustic Laplace Inviscid X X X X X X Irrotational X X X X X Small Perturbations X X X Incompressible X Notes Homogeneous Modeled Turbulence

Lecture 2

M 1 .0
Linearized Linearized Unsteady

Navier-Stokes Equations The Navier-Stokes equations describe the ow of a continuous, Newtonian uid. They may be derived from the principal of conservation of momentum. (For more details see website).

34

AA200B - Applied Aerodynamics II

Lecture 2

where X, Y, Z are the body forces per unit mass in each direction and t is the stress tensor. X,Y, and Z are often associated with gravitational forces and are often neglected. The equations become more usable when the stress tensor is expressed in terms of viscosity and pressure. The pressure and shear forces may be expanded so that the NS equations are:

is the bulk viscosity , relating the normal stress to the rate of change of volume, V . If pressure is a function only of density and not of the rate of change of density, then: = 2/3
35

AA200B - Applied Aerodynamics II

Lecture 2

In the simplest case, with no body force, these equations become: DV = 2V p Dt

Solutions of the full Navier-Stokes equations show the onset of turbulence, the interaction of shear layers, and most interesting aerodynamic phenomena (with the exception of interacting or rareed gas ows). Unfortunately, the equations are very dicult to solve. As the Reynolds number is increased, the scale of the interesting dynamics gets smaller so that most solutions of the full NS equations are done at Reynolds numbers below 1000. Even in these cases the geometries that can be analyzed using the full NS equations are quite simple and it currently does not make sense to consider solving these equations for realistic aircraft congurations. Fortunately many of the approximate equations work quite well in such cases and are much more easily solved.
36

AA200B - Applied Aerodynamics II

Lecture 2

When the time averaged Navier-Stokes equations are not a sucient description of the problem, one may resort to large eddy simulations. This is a numerical solution of the timedependent Navier-Stokes equations, with only the smaller scales of turbulence modeled in an averaged way. While this is faster than solving the full equations, it is still very slow. The gure below shows results from a large eddy simulation of the ow over a 2D circular cylinder. Figure from NASA / Parviz Moin.

37

AA200B - Applied Aerodynamics II

Lecture 2

Reynolds Averaged Navier-Stokes Equations One of the most popular simplications made to the Navier-Stokes Equations is Reynolds Averaging. This simplication to the full Navier-Stokes equations involves taking time averages of the velocity terms in the equations. Writing: u =< u > +u , v =< v > +v , etc. (where <> represents a time average) with the uctuations having zero mean value: < u >= 0 we have: < u2 >=< u >2 + < u 2 >, < uv >=< u >< v > + < u v > This allows us to write the time-averaged NS equations as: (< u >< ux > + < v >< uy > + < w >< uz >) = < px > +2 < u > (< ux > + < u v >y + < u w >z ) and similarly for the y and z components.
38

AA200B - Applied Aerodynamics II

Lecture 2

This looks just like the more general Navier Stokes equations for incompressible ow, which hold for steady, laminar ow except that there are additional terms that act as additional stresses on the right hand side. These terms represent the eect of turbulence on the mean ow. They are called Reynolds stresses and are sometimes said to be caused by eddy viscosity. These terms are generally much larger than the normal viscous terms. The business of predicting these stresses and relating them to the computed mean ow properties is called turbulence modeling. This is usually accomplished empirically or by using the results of detailed time-dependent simulations. Reynolds averaged NS solvers are appropriate for the analysis of viscous, compressible ows and have been applied to rather general congurations, but one must be careful that the assumptions of the turbulence model are compatible with the characteristics of the ow of interest.
39

AA200B - Applied Aerodynamics II

Lecture 2

Euler Equations The momentum equation is sometimes called Eulers equation. (There are lots of equations called Euler equations!) But when people talk of solving the Euler equations these days, they are referring to the inviscid equations of motion given by: p Vi + V Vi + =0 t x With some work, the equation in the x direction becomes: u p ( + V u ) + =0 t x or in vector notation: DV + p = 0 Dt
40

AA200B - Applied Aerodynamics II

Lecture 2

(Recall

DF Dt ,

the substantial, or particle derivative of F is dened by: F DF = + V F Dt t

Also see the note in the derivation of the NS equations. It looks as though we have assumed constant density, but this is not the case.) These are combined with the equations of energy and continuity. The equations are often solved by nite dierences whereby the values of each velocity component, the density, and the internal energy are computed at each point in the ow. From these quantities constitutive relations such as the perfect gas law or the isentropic pressure relation are used to nd pressure. Since Euler equations permit rotational ow and enthalpy losses (through shock waves), they are very useful in solving transonic ow problems, propeller or rotor aerodynamics, and ows with vortical structures in the eld.
41

AA200B - Applied Aerodynamics II

Lecture 2

Full Potential Equation The full potential equation is derived from the assumption of irrotational ow and the equations of continuity and momentum. The pressure and density terms in the Euler equations can be combined when use is made of the perfect gas law and the isentropic relation between pressure and density. Ashley and Landahl show how we may derive the following vector form of the unsteady full potential equation: 1 2 a
2

V2 2 V 2 + + (V ) t2 t 2

=0

This may be simplied for the case of steady ow in 2-D to:


42

AA200B - Applied Aerodynamics II

Lecture 2

xx

2 x 1 2 a

+ yy

2 y 1 2 a

2 2 xy xy = 0 a

About the notation: When ow is irrotational: V = 0 and by denition of curl and gradient: () = 0, where is a scalar eld. Thus we can dene a nonphysical scalar potential, , that describes the velocity eld. is related to the velocities by the relation: V = . The equations can thus be written in terms of the unknown scalar rather than the 3 components of the velocity. This simplies their solution. In the above expressions: a is the local speed of sound, x is the streamwise coordinate, and V is the vector velocity. Subscripts denote partial derivatives with respect to the subscripted variables (e.g. Ux = u x ).
43

AA200B - Applied Aerodynamics II

Lecture 2

Transonic Small Disturbance Equation When the full potential equation is simplied by assuming that perturbation velocities are small and we relate the local speed of sound to the freestream value by making use of the isentropic relations we obtain the small disturbance equation: u 1 2 w 2 2 ux(1 + wz = 2ux M 1 + M + 2uz M U 2 U Letting the freestream Mach number approach 1.0 and ignoring the last term, we get the classic transonic small disturbance equation:
2 M )

u 2 ux(1 + wz = ux M ( + 1) U The TSD equation is used less frequently these days since nite dierence methods can be used to solve the full potential equation directly.
2 M )
44

AA200B - Applied Aerodynamics II

Lecture 2

Prandtl-Glauert Equation Prandtl-Glauert Equation The PrandtlGlauert equation is a linearized form of the full potential equation. Full potential: xx 2 x 1 2 a + yy 2 y 1 2 a 2 xy xy = 0 2 a

If the velocity perturbations are much smaller than the freestream velocity, this expression becomes:
2 1 M )xx + yy = 0

or in the unsteady case: 1


2 M )xx

+ yy

1 = 2 (2Uxt + tt) a
45

AA200B - Applied Aerodynamics II

Lecture 2

The 3-D version is easily constructed with the addition of z derivatives corresponding to the y derivatives shown here. Note that this linearized form of the equation does not hold near the nose of an airfoil where the velocity perturbation is of the same order as the freestream, unless the freestream Mach number is itself small. Also note that this expression holds for subsonic and supersonic ow (but not transonic ow). It forms the basis for many aerodynamic analysis methods.

Analysis of P51 Mustang from Analytical Methods, Inc. VSAERO, a code that solves the Prandtl-Glauert equations.

using

46

AA200B - Applied Aerodynamics II

Lecture 2

Laplaces Equation One of the simplest, but most useful expressions in applied aerodynamics is Laplaces equation: 2 = 0

Laplaces equation is the Prandtl-Glauert equation in the limit as the freestream Mach number goes to zero, although it does not require the strong assumption of small perturbations. It was actually rst derived by Euler. The derivation is very simple, requiring only the equation of continuity, and the assumptions of irrotational and constant density ow. The continuity equation for constant density ow in 2D can be written: ux + vy = 0.
47

AA200B - Applied Aerodynamics II

Lecture 2

Since the ow is irrotational we can write: V = . Substitution into the continuity equation yields: 2 = 0. It is interesting to note that Laplaces equation does not require the assumption of small perturbations, while the Prandtl-Glauert equation does. In fact, near the stagnation point of an airfoil where velocities become small, the full potential equation reduces to Laplaces equation, not the Prandtl-Glauert equation. Note also that all of the time dependent terms in the full potential 1 equation are multiplied by a 2 so that this form of the equation holds for unsteady phenomena as well.

48

AA200B - Applied Aerodynamics II

Lecture 2

Relating Velocity and Pressure: Bernoullis Equations


Some of the equations we have discussed are posed in terms of state variables that do not include pressures. In these cases (e.g. the potential ow equations) the dierential equations and boundary conditions allow one to compute the local velocities, but not the pressures. Once the velocities are known, however, the momentum equation can be used to nd the local pressure. Such equations are known as Bernoulli equations and they come in various forms, depending on the assumptions that can be made about the ow. The conservation of momentum principle is the source of the relation between pressure and velocity.

49

AA200B - Applied Aerodynamics II

Lecture 2

Incompressible Flow For steady incompressible ow along a streamline, the Bernoulli equation may be written: V2 + p = pT 2 When the ow is not steady, the Euler equations can be integrated to obtain a more general form of this result: Kelvins equation, the Bernoulli equation for irrotational ow. V 2 + + t 2 dp = F (t)

where F is an arbitrary function of time.


50

AA200B - Applied Aerodynamics II

Lecture 2

If we do not assume that the ow is irrotational, we cannot introduce the potential and the expression is not so nicely integrable. If, however, we assume that the ow is steady with no body forces, but not necessarily irrotational we can write the following expression that holds along a streamline: V2 + 2 dp =F

While the above equations hold for steady ows along a streamline, for irrotational ows they hold throughout the uid. Compressible Forms We can derive a more useful form of the Bernoulli equation by starting with the expression for steady ow without body forces shown just above.
51

AA200B - Applied Aerodynamics II

Lecture 2

If the ow is assumed to be isentropic ow (no entropy change or heat addition): p = k . Substitution yields the compressible Bernoulli equation: p V2 + = const 1 2 This actually works for adiabatic (no heat transfer) ows as well as isentropic ows. In summary, we often deal with one of two simple forms of the Bernoulli equation shown below. Incompressible: Compressible:
V2 2

+ p = pT + V2 =
2

p 1

pT 1 T
52

AA200B - Applied Aerodynamics II

Lecture 2

Static, Total, and Dynamic Pressures In both the incompressible and compressible forms of Bernoullis equation shown above there are 3 terms. The quantity pT is the total or stagnation pressure. It is the pressure that would be measured at points in the ow where V = 0. The other p in the above expressions is the static pressure. Note that in incompressible ow, the speed is directly related to the dierence in total and static pressure. This can be measured directly with a pitot-static probe shown below.

The dynamic pressure is dened as: q

V 2 2 .
53

AA200B - Applied Aerodynamics II

Lecture 2

Pressure Coecient The static pressure coecient is dened as: p p Cp = 1 2 2 U where p is the freestream static pressure. In incompressible ow, the expression for Cp is especially simple: V2 Cp = 1 2 U If the local velocity is expressed as a small perturbation in the freestream: V = U + U , then the incompressible Cp relation can U be written: Cp 2 U . Be careful with this expression! It is often not a good approximation and the correct expression is not very dicult.
54

AA200B - Applied Aerodynamics II

Lecture 2

Isentropic Pressure Rule The expression for Cp in compressible isentropic ow (sometimes called the isentropic pressure rule) is derived from the compressible Bernoulli equation along with the expression for the speed of sound in a perfect gas. In terms of the local Mach number the expression is:
1 2 2 1 + 2 M Cp = 1 2 2 M 1 + M 2

In air with gamma = 1.4 1 Cp = 2 0.7M


2 1 + 0.2M 1 + 0.2M 2 3 .5

55

AA200B - Applied Aerodynamics II

Lecture 2

Some interesting results follow from this expression... We can tell if the ow is supersonic, just by looking at the value of Cp. The critical value of Cp, denoted Cp is found by setting M = 1 in the above expression: 1 Cp = 2 0.7M
2 1 M + 1.2 6 3.5

Also, we see that there is a minimum value of Cp, corresponding to a complete vacuum. Setting the local Mach number to innity yields: 1 Cpvacuum = 2 0.7M Cp cannot be any more negative than this. Experiments show that airfoils can get to about 70% of vacuum Cp. This can limit the maximum lift of supersonic wings.
56

AA200B - Applied Aerodynamics II

Lecture 2

Cp* and Cp_vacuum vs. Mach


2 1 0
Cp
Cp* Cp_vacuum

-1 -2 -3 -4 -5 0 0.5 1
Mach

1.5

57

AA200B - Applied Aerodynamics II

Lecture 2

References
1. Anderson, Fundamentals of Aerodynamics, 2nd Edition, McGraw-Hill, 1991. 2. Shevell, R.S., Fundamentals of Flight, Prentice-Hall, 1983. 3. Kuethe and Chow, Foundations of Aerodynamics, 4th edition, Wiley, 1986. 4. Schlichting, H., Boundary Layer Theory, McGraw-Hill, 1955. 5. Moran, J., An Introduction to Theoretical and Computational Aerodynamics, Wiley, 1984. 6. Ashley, H., Landahl, M., Aerodynamics of Wings and Bodies, AddisonWesley, 1965, also Dover Edition, 1985. 7. Nixon, D. (ed.), Transonic Aerodynamics, Progress in Astronautics and Aeronautics, Vol. 81, AIAA, 1982. 8. McMasters, J., Commercial Aircraft Systems Engineering and Design, AIAA Short Course, Feb. 1993.
58

Das könnte Ihnen auch gefallen