You are on page 1of 14

Unit 3 Subsonic Compressible Flow over Airfoils: Linear Theory

Introduction - The velocity potential equation - The linearized velocity potential


equation - Prandtl-Glauert Compressibility correction - Critical Mach number -
Drag divergence mach number The sound barrier - Area Rule - Supercritical
Airfoil.

3.1 The Velocity Potential Equation


The inviscid, com iform
stream is irrotational; there is no mechanism in such a
elements. Thus, a velocity potential
Consider two-dimensional, steady, irrotational, is
potential, ned such that

or in terms of the cartesian velocity components,

Let us proceed to obtain an equation for which represents velocity potential


here and which is a combination of the continuity, momentum, and energy
equations.
The continuity equation for steady, two-
Equation (2.52) as

Substituting the value from Equation (1) and (2)


We are attempting to obtain an equation completely in terms of ; hence, we
need to eliminate from Equation (4). To do this, consider the momentum

This equation holds for a and


along a streamline. Solving Equation (5) further

Recall that we are also Thus, for that case


we have,

Substituting Equation (7) into (6)

Considering changes in the direction

Similarly, for changes in the direction

Substituting Equations (9) and (10) into (4), cancelling the which appears in
each term, and factoring out the second derivatives of , we obtain

This above Equation (11) is called the velocity potential equation.


3.2 The Linearized Velocity Potential Equation
Consider the two-
in Figure 3.1. The body is immersed in a uniform ow with velocity oriented
in the direction. At an arbitrary point P i with
the and components given by and , respectively. Let us now visualize the
velocity ements in
velocity. For example, the component of velocity in Figure 3.1 can be
visualized as plus an increment in velocity (positive or negative). Similarly,
the component of velocity can be visualized as a simple increment itself,
because the uniform ow has a zero component in the direction. These
increments are called perturbations, and

Figure 3.1

Where and are called the perturbation velocities. These perturbation


velocities are not necessarily small; indeed, they can be quite large in the
stagnation region in front of the blunt nose of the body shown in Figure 3.1. In
the same vein, because , we can de ne a perturbation velocity potential
such that

Where,
Hence,

Substituting these values in the Equation (11) that is velocity potential equation
we get,

This is said to be Equation (12) and called the perturbation velocity


potential equation. It is precisely the same equation as Equation (11) except that
it is expressed in terms of instead of . Now, let us recast Equation (12) in
terms of the perturbation velocities,

Also, from the energy equation

Substituting Equation (14) into (13) and rearranging them we have,

Keep in mind that the size of the perturbations can be large or small; Let us now
limit our considerations to small perturbations only; that is, assume that the
body is at small angle of attack. In such a case, and will be small in
comparison with . Therefore, we have
Keep in mind that products of and with their derivatives are also very small.
Thus, right side of the Equation (15) becomes zero

In terms of velocity potential

Thus, Equation (16) is known as The linearized perturbation velocity potential


equation

3.3 Prandtl-Glauert Compressibility correction


Over the past year we have collected data for low-speed aerodynamics, and
because there was no desire to totally discard such data, the natural approach
to high-speed subsonic aerodynamics was to search for methods that would

would approximately take into account the effects of compressibility. Such


methods are called compressibility corrections. Now, for thin airfoils at small
angles of attack the linearized perturbation velocity potential equation is given
by

And because of subsonic theory and begins to give inappropriate results at


values of = 0.7 and above.
. The
shape of the airfoil is given by . Assume that the airfoil is thin and that

by above Equation. And given by,


Thus, Equation (17) can be rewritten as
Let us transform the independent variables and into a new space, and ,
such that,
Moreover, in this transformed space, consider a new velocity potential such
that,
Equation (18) can be rewritten in terms of the transformed variables, by
recalling the chain rule of partial differentiation,

And

Now from Equation (19) we have

Hence, Equation (21) and (22) can be rewritten and, also considering Equation
(20)

Differentiating Equation (23) with respect to , we obtain

Differentiating Equation (24) with respect to , we obtain

Substituting value from Equation (25) and (26) into (18)


This Equation (27) is known as Laplace s Equation which is a governing relation
for incompressible flow.
Now, we know that for the linearized pressure we have,

of velocity potential, , where is a perturbation


. Thus, Equation (28) can be written as

the expression is simply the linearized pressur


.
Hence, Equation (29) gives

Equation (30) is called the Prandtl-Glauert rule; it states that, if we know the
incompressible pressure distribution over an airfoil, then the compressible
pressure distribution over the same airfoil can be obtained from Equation (30).
Therefore, Equation (30) is truly a compressibility correction to incompressible
data.

3.4 Critical Mach number


Consider an airfoil in a low- , as sketched in
Figure 3.2. In the expansion over the top surface of the airfoil, the l
Mach number increases. Let point A represent the location on the airfoil
surface where the pressure is a minimum, hence where is a maximum. In
Figure 3.2 (a), let us say this maximum is . Now assume that we
gradually increase the freestream Mach number. As increases, also
increases. Let us continue to increase until we achieve just the right value
such that the local Mach number at the minimum pressure point equals 1, that
is, such that , as shown in Figure 3.2 (c). When this happens, the
freestream Mach number is called the Critical Mach number.

Figure 3.2

at
which sonic ow is rst achieved on the airfoil surface. In Figure 3.2(c),

3.5 Drag divergence mach number The sound barrier


gle of attack in a wind tunnel,
and we wish to measure its . To begin
d,0, shown in
Figure 3.3. Now, as we gradually increase the freestream Mach number, we
observe that cd remains relatively constant all the way to the critical Mach
number, as illustrated in Figure 3.3. The ow elds associated with points , ,
and in Figure 3.3 are represented by Figure 3.2 a, b, and c, respectively. As we
continue to nudge higher, we encounter a point where the drag
suddenly starts to increase. This is given as point in Figure 3.3. The value of
at which this sudden increase in drag starts as the Drag-divergence
Mach number .
When approaches to 1, cd will tend to infinity by Prandtl-Glauert rule, as we
seen in Figure 3.3 at point g, and that point is known as Sound Barrier .

Figure 3.3

3.6 Area Rule


A plan view, cross section, and area distribution (cross-sectional area versus
distance along the axis of the airplane) for a typical airplane is sketched in Figure
3.4. Let A denote the total cross-sectional area at any given station. Note that
the cross-sectional area distribution experiences some abrupt changes along the
axis, with discontinuities in both A and dA/dx in the regions of the wing.
Figure 3.4

The variation of this cross-sectional area for an airplane should be smooth, with
no discontinuities. This meant that, in the region of the wings and tail, the
fuselage cross-sectional area should decrease to compensate for the addition of
the wing and tail cross sectional area. This led to a fuselage shape,
as shown in Figure 3.5. Here, the plan view and area distribution are shown for
an aircraft with a relatively smooth variation of A(x). This design philosophy is
called the Area Rule .

Figure 3.5
3.7 Supercritical Airfoil
The purpose of a supercritical airfoil is to increase the value of Mdrag-divergence,
although Mcr may change very little. Consider a 13-percent thick supercritical
airfoil is shown in Figure 3.6 (c) whereas Figure 3.6 (a) is just a normal airfoil. The
us encouraging a region of
comparatively lower local values of M. Also, the
terminating shock is weaker, thus creating less drag. Similar trends can be seen
by comparing the Cp distributions in Figure 3.6 (b) and (d). Clearly, the
supercritical airfoil s

numbers are lower, and the terminating shock wave is weaker. As a result, the
value of Mdrag-divergence will be higher for the supercritical airfoil.

Figure 3.6
Unit 4 Linearized Supersonic flow
Introduction to supersonic flow - Derivation of the Linearized supersonic
pressure coefficient formula - Application to supersonic airfoils - Supersonic
airfoil drag.

4.1 Introduction to supersonic flow


The linearized perturbation velocity potential equation (16) holds for both
subsonic and supersonic flow. We know that for subsonic flow Equation (16) is
an elliptical partial differential equation and for supersonic flow it is a hyperbolic
differential equation. Thus, we can say there is a fundamental difference in the
physical .

4.2 Derivation of the Linearized supersonic pressure coefficient


formula
For the case of the supersonic flow let us write Equation (16) as

Where,

A solution of this equation is the functional relation, i.e.

Now, to nt in supersonic
. Take the derivative of Equation (32)

Thus, we can say that


From the linearized boundary condition, we know that,

We can further reduce that for a small perturbations is small.


Hence,

Substituting Equation (35) into (33)

Recall

Substituting Equation (36) into (37)

Thus, Equation (38) is Linearized supersonic pressure coefficient formula for


subsonic flow.

4.3 Application to supersonic airfoils


In the above expressions, is always treated as a positive quantity, and the sign
of Cp is determined simply by looking at the body and noting whether the surface
is inclined into or away from the freestream.
For example, points A and B in Figure 4.1 are on surfaces inclined into the
freestream, and hence Cp,A and Cp,B are positive values given by

And when the surface is inclined away from the freestream direction, linearized
theory predicts a negative Cp. For example, points C and D in Figure 4.1 are on
surfaces inclined away from the freestream, and hence Cp,C and Cp,D are negative
values, given by
Figure 4.1

Also, we can get the equation for lift and drag coefficient

4.4 Supersonic Airfoil Drag


In addition to the skin-friction drag, a supersonic airfoil also experiences
supersonic wave. The source of wave drag is the pressure distribution exerted
over the airfoil surface and is a result of the shock-wave and expansion-wave
-friction drag is, of course,
the shear stress exerted over the airfoil surface and is the result of friction in the
ical mechanisms of wave drag and skin-friction drag clearly are
quite different.

You might also like