Applied Computational Aerodynamics
Applied Computational Aerodynamics
Applied Computational Aerodynamics
COMPUTATIONAL
AERODYNAMICS
Preface
Objectives
These notes are intended to fill a significant gap in the literature available to students. There is a huge
disparity between the aerodynamics covered in typical aerodynamics courses and the application of
aerodynamic theory to design and analysis problems using computational methods. As an elective
course for seniors, Applied Computational Aerodynamics provides an opportunity for students to gain
insight into the methods and means by which aerodynamics is currently practiced. The specific
threefold objective is: i) physical insight into aerodynamics that can arise only with the actual
calculation and subsequent analysis of flowfields, ii) development of engineering judgment to answer
the question how do you know the answer is right? and iii) establishment of a foundation for future
study in computational aerodynamics; exposure to a variety of methods, terminology, and jargon.
Two features are unique. First, when derivations are given, all the steps in the analysis are included.
Second, virtually all the examples used to illustrate applied aerodynamics ideas were computed by the
author, and were made using the codes available to the students. The exercises are an extremely
important component of the course, where parts of the course are possibly best presented as a
workshop, rather than as a series of formal lectures. To meet the objectives, many old fashioned
methods are included. Using these methods a student can learn much more about aerodynamic design
than by performing a few large modern calculations. For example (articulated to the author by Prof.
Ilan Kroo), the vortex lattice method allows the student to develop an excellent mental picture of the
flowfield. Thus these methods provide a context within which to understand Euler or Navier-Stokes
calculations.
Audience
We presume that the reader has had standard undergraduate courses in fluid mechanics and
aerodynamics. In some cases the material is repeated to illustrate issues important to computational
aerodynamics. Access to a computer and the ability to program is assumed for the exercises.
Warnings
Computational aerodynamics is still in an evolutionary phase. Although most of the material in the
early chapters is essentially well established, the viewpoint adopted in the latter chapters is necessarily
a snapshot of the field at this time. Students that enter the field can expect to use this material as a
starting point in understanding the continuing evolution of computational aerodynamics.
These notes are not independent of other texts. At this point several of the codes used in the instruction
are based on source codes copyrighted in other sources. Use of these codes without owning the text
may be a violation of the copyright law.
The traditional printed page is inadequate and obsolete for the presentation of computational
aerodynamics information. The reader should be alert to advances in information presentation, and
take every opportunity to make use of advanced color displays, interactive flowfield visualization and
virtual environment technology.
The codes available on disk provide a significant capability for skilled users. However, as discussed in
the text, few computational aerodynamics codes are ever developed and tested to the level that they
are bug free. They are for educational use only, and are only aides for education, not commercial
programs, although they are entirely representative of codes in current use.
Acknowledgements
Many friends and colleagues have influenced the contents of these notes. Specifically, they reflect
many years developing and applying computational aerodynamics at Grumman, which had more than
its share of top flight aerodynamicists. Initially at Grumman and now at VPI, Bernard Grossman
provided access to his as yet unpublished CFD course notes. At NASA, many friends have contributed
help, insight and computer programs. Nathan Kirschbaum read the notes and made numerous
contributions to the content and clarity. Several classes of students have provided valuable feedback,
found typographical and actual errors. They have also insisted that the notes and codes be completed.
I would like to acknowledge these contributions.
W.H. Mason
Return to the main table of contents
4.1 An Introduction
The incompressible potential flow model provides reliable flowfield predictions over a wide range
of conditions. For the potential flow assumption to be valid for aerodynamics calculations the
primary requirement is that viscous effects are small in the flowfield, and that the flowfield must be
subsonic everywhere. Locally supersonic velocities can occur at surprisingly low freestream Mach
numbers. For high-lift airfoils the peak velocities around the leading edge can become supersonic
at freestream Mach numbers of 0.20 ~ 0.25. If the local flow is at such a low speed everywhere
that it can be assumed incompressible (M .4, say), Laplaces Equation is essentially an exact
representation of the inviscid flow. For higher subsonic Mach numbers with small disturbances to
the freestream flow, the Prandtl-Glauert (P-G) Equation can be used. The P-G Equation can be
converted to Laplaces Equation by a simple transformation.1 This provides the basis for estimating
the initial effects of compressibility on the flowfield, i.e., linearized subsonic flow. In both
cases, the flowfield can be found by the solution of a single linear partial differential equation. Not
only is the mathematical problem much simpler than any of the other equations that can be used to
model the flowfield, but since the problem is linear, a large body of mathematical theory is
available.
The Prandtl-Glauert Equation can also be used to describe supersonic flows. In that case the
mathematical type of the equation is hyperbolic, and will be mentioned briefly in Chapter 12.
Recall the important distinction between the two cases:
subsonic flow:
supersonic flow:
2/24/98
4-1
(4.1)
There are several ways to view the solution of this equation. The one most familiar to
aerodynamicists is the notion of singularities. These are algebraic functions which satisfy
Laplaces equation, and can be combined to construct flowfields. Since the equation is linear,
superposition of solutions can be used. The most familiar singularities are the point source, doublet
and vortex. In classical examples the singularities are located inside the body. Unfortunately, an
arbitrary body shape cannot be created using singularities placed inside the body. A more
sophisticated approach has to be used to determine the potential flow over arbitrary shapes.
Mathematicians have developed this theory. We will draw on a few selected results to help
understand the development of panel methods. Initially, we are interested in the specification of the
boundary conditions. Consider the situation illustrated Fig. 4-1.
*
The singularities are distributed across the panel. They are not specified at a point. However, the boundary
conditions usually are satisfied at a specific location.
**
These will be mentioned in more detail in Chapter 9.
2/24/98
(4-2)
(4-3)
Potential flow theory states that you cannot specify both arbitrarily, but can have a mixed
boundary condition, a + b /n on + . The Neumann Problem is identified as analysis
above because it naturally corresponds to the problem where the flow through the surface is
specified (usually zero). The Dirichlet Problem is identified as design because it tends to
correspond to the aerodynamic case where a surface pressure distribution is specified and the body
shape corresponding to the pressure distribution is sought. Because of the wide range of problem
formulations available in linear theory, some analysis procedures appear to be Dirichlet problems,
but Eq. (4-3) must still be used.
Some other key properties of potential flow theory:
If either or /n is zero everywhere on + then = 0 at all interior points.
cannot have a maximum or minimum at any interior point. Its maximum value can
only occur on the surface boundary, and therefore the minimum pressure (and
maximum velocity) occurs on the surface.
2/24/98
divAdV = A n dS
R
(4-4)
we follow the classical derivation and consider the interior problem as shown in Fig. 4-2.
S0
n
z
y
R0
x
Figure 4-2. Nomenclature for integral equation derivation.
To start the derivation introduce the vector function of two scalars:
A = grad grad .
(4-5)
Substitute this function into the Gauss Divergence Theorem, Eq. (4-4), to obtain:
(4-6)
Now use the vector identity: F= F + F to simplify the left hand side of Eq. (4-6).
Recalling that A = divA, write the integrand of the LHS of Eq. (4-6) as:
div(grad grad ) = ( ) ( )
= +
= 2 2
2/24/98
(4-7)
(4-8)
( )dV = n n dS .
2
(4-9)
1 2
r r ndS = r
S0
R0
2 1
dV .
r
(4-10)
R 0 is the region enclosed by the surface S0 . Recognize that on the right hand side the first term,
2 , is equal to zero by definition so that Eq. (4-10) becomes
1
2 1
r r ndS =
S0
R0
dV .
(4-11)
1
If a point P is external to S0 , then 2 = 0 everywhere since 1/r is a source, and thus satisfies
r
Laplaces Equation. This leaves the RHS of Eq. (4-11) equal to zero, with the following result:
1
r r n dS = 0.
(4-12)
S0
However, we have included the origin in our region S0 as defined above. If P is inside S0 , then
1
2 at r = 0. Therefore, we exclude this point by defining a new region which excludes
r
the origin by drawing a sphere of radius around r = 0, and applying Eq. (4-12) to the region
between and S0 :
2/24/98
r r ndS r r + r2 dS = 0
(4-13)
S0
4 42443
1
444
4244 44
3 1
sphere
arbitrary region
or:
r r + r2 dS = r r ndS .
(4-14)
S0
Consider the first integral on the left hand side of Eq. (4-14). Let 0, where (as 0)
we take constant ( / r == 0 ), assuming that is well-behaved and using the mean value
theorem. Then we need to evaluate
dS
r2
over the surface of the sphere where = r. Recall that for a sphere* the elemental area is
dS = r 2 sin dd
(4-15)
where we define the angles in Fig. 4-3. Do not confuse the classical notation for the spherical
coordinate angles with the potential function. The spherical coordinate will disappear as soon as
we evaluate the integral.
z
y
x
sin dd .
See Hildebrand, F.B., Advanced Calculus for Applications, 2nd Ed., Prentice-Hall, Englewood Cliffs, 1976 for an
excellent review of spherical coordinates and vector analysis.
2/24/98
=0 =0 sin dd = 4 .
(4-16)
The final result for the first integral in Eq. (4-14) is:
1
r r + r2 dS = 4 .
(4-17)
Replacing this integral by its value from Eq. (4-17) in Eq. (4-14), we can write the expression
for the potential at any point P as (where the origin can be placed anywhere inside S0 ):
( p) =
1
4
r r n dS
(4-18)
s0
and the value of at any point P is now known as a function of and /n on the boundary.
We used the interior region to allow the origin to be written at point P. This equation can be
extended to the solution for for the region exterior to R 0 . Apply the results to the region between
the surface SB of the body and an arbitrary surface enclosing SB and then let go to infinity. The
integrals over go to as goes to infinity. Thus potential flow theory is used to obtain the
important result that the potential at any point P' in the flowfield outside the body can be expressed
as:
( p ) =
1
4
r r n dS .
(4-19)
SB
Here the unit normal n is now considered to be pointing outward and the area can include not only
solid surfaces but also wakes. Equation 4-19 can also be written using the dot product of the
normal and the gradient as:
( p ) =
1
4
r n n r dS .
(4-20)
SB
The 1/r in Eq. (4-19) can be interpreted as a source of strength / n , and the (1/r) term in
Eq. (4-19) as a doublet of strength . Both of these functions play the role of Greens functions in
the mathematical theory. Therefore, we can find the potential as a function of a distribution of
sources and doublets over the surface. The integral in Eq. (4-20) is normally broken up into
body and wake pieces. The wake is generally considered to be infinitely thin. Therefore, only
doublets are used to represent the wakes.
2/24/98
1
4
r n r dS .
(4-21)
SB
The problem is to find the values of the unknown source and doublet strengths and for a
specific geometry and given freestream, .
What just happened? We replaced the requirement to find the solution over the entire flowfield
(a 3D problem) with the problem of finding the solution for the singularity distribution over a
surface (a 2D problem). In addition, we now have an integral equation to solve for the unknown
surface singularity distributions instead of a partial differential equation. The problem is linear,
allowing us to use superposition to construct solutions. We also have the freedom to pick whether
to represent the solution as a distribution of sources or doublets distributed over the surface. In
practice its been found best to use a combination of sources and doublets. The theory can be
extended to include other singularities.
At one time the change from a 3D to a 2D problem was considered significant. However, the
total information content is the same computationally. This shows up as a dense 2D matrix vs. a
sparse 3D matrix. As methods for sparse matrix solutions evolved, computationally the problems
became nearly equivalent. The advantage in using the panel methods arises because there is no
need to define a grid throughout the flowfield.
This is the theory that justifies panel methods, i.e., that we can represent the surface by panels
with distributions of singularities placed on them. Special precautions must be taken when
applying the theory described here. Care should be used to ensure that the region SB is in fact
completely closed. In addition, care must be taken to ensure that the outward normal is properly
defined.
Furthermore, in general, the interior problem cannot be ignored. Surface distributions of
sources and doublets affect the interior region as well as exterior. In some methods the interior
problem is implicitly satisfied. In other methods the interior problem requires explicit attention. The
need to consider this subtlety arose when advanced panel methods were developed. The problem is
not well posed unless the interior problem is considered, and numerical solutions failed when this
aspect of the problem was not addressed. References 4 and 5 provide further discussion.
2/24/98
2/24/98
q(s)
(s)
=
{
+
lnr
ds
224
2
4
3
1
23
uniform onset flow
1
(4-22)
and = tan (y/x). Although the equation above shows contributions from various components of
-1
the flowfield, the relation is still exact. No small disturbance assumption has been made.
4.4 The Classic Hess and Smith Method
A.M.O. Smith at Douglas Aircraft directed an incredibly productive aerodynamics development
group in the late 50s through the early 70s. In this section we describe the implementation of the
theory given above that originated in his group.* Our derivation follows Morans description6 of
the Hess and Smith method quite closely. The approach is to i) break up the surface into straight
line segments, ii) assume the source strength is constant over each line segment (panel) but has a
different value for each panel, and iii) the vortex strength is constant and equal over each panel.
Roughly, think of the constant vortices as adding up to the circulation to satisfy the Kutta
condition. The sources are required to satisfy flow tangency on the surface (thickness).
Figure 4-5 illustrates the representation of a smooth surface by a series of line segments. The
numbering system starts at the lower surface trailing edge and proceeds forward, around the
leading edge and aft to the upper surface trailing edge. N+1 points define N panels.
node
N -1
N
N+1
panel
Figure 4-5. Representation of a smooth airfoil with straight line segments.
The potential relation given above in Eq. (4-22) can then be evaluated by breaking the integral
up into segments along each panel:
N
= V ( xcos + ysin ) +
j=1panel j
q(s)
2 ln r 2 dS
(4-23)
In the recent AIAA book, Applied Computational Aerodynamics, A.M.O. Smith contributed the first chapter, an
account of the initial development of panel methods.
2/24/98
i+1
li
^t
i
^n
i
i+1
i
x
a) basic nomenclature
x x
cos i = i+1 i
li
(4-24)
(4-25)
We will find the unknowns by satisfying the flow tangency condition on each panel at one
specific control point (also known as a collocation point) and requiring the solution to satisfy the
Kutta condition. The control point will be picked to be at the mid-point of each panel, as shown in
Fig. 4-7.
Y
smooth shape
control point
panel
X
2/24/98
y +y
yi = i i +1
2
(4-26)
and the velocity components at the control point xi , yi are ui = u(xi , yi ), vi = v(xi , yi ).
The flow tangency boundary condition is given by V n = 0, and is written using the relations
given here as:
ui sini + vi cos i = 0,
(4-27)
The remaining relation is found from the Kutta condition. This condition states that the flow
must leave the trailing edge smoothly. Many different numerical approaches have been adopted to
satisfy this condition. In practice this implies that at the trailing edge the pressures on the upper and
lower surface are equal. Here we satisfy the Kutta condition approximately by equating velocity
components tangential to the panels adjacent to the trailing edge on the upper and lower surface.
Because of the importance of the Kutta condition in determining the flow, the solution is extremely
sensitive to the flow details at the trailing edge. When we make the assumption that the velocities
are equal on the top and bottom panels at the trailing edge we need to understand that we must
make sure that the last panels on the top and bottom are small and of equal length. Otherwise we
have an inconsistent approximation. Accuracy will deteriorate rapidly if the panels are not the same
length. We will develop the numerical formula using the nomenclature for the trailing edge shown
in Fig. 4-8.
^t
N
N+1
^t
1
1
(4-28)
and taking the difference in direction of the tangential unit vectors into account this is written as
V t 1 = V t N .
2/24/98
(4-29)
(4-30)
The expression for the potential in terms of the singularities on each panel and the boundary
conditions derived above for the flow tangency and Kutta condition are used to construct a system
of linear algebraic equations for the strengths of the sources and the vortex. The steps required are
summarized below. Then we will carry out the details of the algebra required in each step.
Steps to determine the solution:
1. Write down the velocities, ui, v i, in terms of contributions from all the singularities. This
includes qi, from each panel and the influence coefficients which are a function of the
geometry only.
2. Find the algebraic equations defining the influence coefficients.
To generate the system of algebraic equations:
3. Write down flow tangency conditions in terms of the velocities (N eqns., N+1
unknowns).
4. Write down the Kutta condition equation to get the N+1 equation.
5. Solve the resulting linear algebraic system of equations for the qi, .
6. Given qi, , write down the equations for uti, the tangential velocity at each panel control
point.
7. Determine the pressure distribution from Bernoullis equation using the tangential
velocity on each panel.
We now carry out each step in detail. The algebra gets tedious, but theres no problem in
carrying it out. As we carry out the analysis for two dimensions, consider the additional algebra
required for the general three dimensional case.
2/24/98
Step 1. Velocities
The velocity components at any point i are given by contributions from the velocities induced
by the source and vortex distributions over each panel. The mathematical statement is:
N
j=1
j =1
j=1
j=1
(4-31)
q jvs ij + vvij
where qi and are the singularity strengths, and the usij , v sij , uvij, and v vij are the influence
coefficients. As an example, the influence coefficient usij is the x-component of velocity at x i due to
a unit source distribution over the j th panel.
Step 2. Influence coefficients
To find usij , v sij , uvij, and v vij we need to work in a local panel coordinate system x*, y* which
leads to a straightforward means of integrating source and vortex distributions along a straight line
segment. This system will be locally aligned with each panel j, and is connected to the global
coordinate system as illustrated in Fig. 4-9.
Y
Y*
lj
j+1
X*
j
j
2/24/98
(4-32)
Q
x
,
2
2 x + y2
v(x,y) =
Q
y
.
2
2 x + y2
(4-34)
In general, if we locate the sources along the x-axis at a point x = t, and integrate over a length l,
the velocities induced by the source distributions are obtained from:
q(t)
x t
dt
t=0 2 (x t)2 + y 2
us =
t=l
q(t)
y
vs =
dt
t=0 2 (x t)2 + y 2
t=l
(4-35)
To obtain the influence coefficients, write down this equation in the ( )* coordinate system,
with q(t) = 1 (unit source strength):
u*sij =
1 lj
xi* t
dt
2 0 (x * t) 2 + y*2
i
v*sij
1 lj
y*i
=
dt
2 0 (x * t) 2 + y*2
i
i
(4-36)
u*sij =
1 *
ln x t
2 i
1 t=l j
2
2
+ y*i 2
t=0
(4-37)
t=l
v*sij
j
*
1
1 yi
=
tan *
2
xi t t=0
To interpret these expressions examine Fig. 4-10. The notation adopted and illustrated in the
sketch makes it easy to translate the results back to global coordinates.
2/24/98
y*
x*, y*
i i
rij
ri,j+1
ij
0
j
j+1
lj
x*
1 ri, j+1
ln
2 rij
ij
(4-38)
0
v*sij = l
=
2
2
Here rij is the distance from the j th node to the point i, which is taken to be the control point
location of the i th panel. The angle ij is the angle subtended at the middle of the i th panel by the j th
panel.
The case of determining the influence coefficient for a panels influence on itself requires
some special consideration. Consider the influence of the panel source distribution on itself. The
source induces normal velocities, and no tangential velocities, Thus, u*sii = 0 and vs*ii depends on
the side from which you approach the panel control point. Approaching the panel control point
from the outside leads to ii = , while approaching from inside leads to ii = -. Since we are
working on the exterior problem,
ii = ,
(4-39)
and to keep the correct sign on ij, j i, use the FORTRAN subroutine ATAN2, which takes into
account the correct quadrant of the angle.*
*
2/24/98
e .
2r
(4-40)
Compared to the source flow, the u, v components simply trade places (with consideration of the
direction of the flow to define the proper signs). In Cartesian coordinates the velocity due to a point
vortex is:
x
u(x,y) = +
,
v(x, y) =
.
(4-41)
2
2
2
2 x + y
2 x + y2
where the origin (the location of the vortex) is x = y = 0.
Using the same analysis used for source singularities for vortex singularities the equivalent
vortex distribution results can be obtained. Summing over the panel with a vortex strength of unity
we get the formulas for the influence coefficients due to the vortex distribution:
uv*ij = +
1 lj
y*i
dt = ij
2
2 0 (x * t)2 + y*
2
i
i
1 lj
xi* t
1 ri, j+1
v*vij =
dt
=
ln
2 0 (x * t)2 + y*2
2 rij
i
i
(4-42)
where the definitions and special circumstances described for the source singularities are the same
in the current case of distributed vortices.* In this case the vortex distribution induces an axial
velocity on itself at the sheet, and no normal velocity.
Step 3. Flow tangency conditions to get N equations.
Our goal is to obtain a system of equations of the form:
N
Aij q j + Ai,N +1 = bi
i = 1,...N
(4-43)
j=1
which are solved for the unknown source and vortex strengths.
Recall the flow tangency condition was found to be:
ui sini + vi cos i = 0,
*
for each i,
i = 1,...N
Note that Morans Equation (4-88) has a sign error typo. The correct sign is used in Eq. (4-42) above.
2/24/98
(4-44)
j=1
j =1
j=1
j=1
(4-45)
q jvs ij + vvij
Substituting into Eq. (4-45), the flow tangency equations, Eq. (4-44), above:
N
N
N
N
j =1
j=1
j=1
j =1
(4-46)
which is rewritten into:
N
+ cos i q jvsij
j=1
j =1
j=1
j=1
Aij
N
N
j=1
j=1
14444424 4444
3
(4-47)
Ai, N +1
Now get the formulas for A ij and A i,N+1 by replacing the formulas for usij , v sij ,uvij,v vij with the ( )*
values, where:
u = u * cos j v*sin j
v = u * sin j + v * cos j
and we substitute into Eq. (4-47) for the values in A ij and A i,N+1 above.
Start with:
2/24/98
(4-48)
= cos i u*sij sin j + v*sij cos j sini u*sij cos j v*sij sin j
(4-49)
= cosi sin j sini cos j u*sij + cos i cos j sini sin j vs*ij
and we use trigonometric identities to combine terms into a more compact form. Operating on the
first term in parenthesis:
1
1
cosi sin j = sin i + j + sin i j
2
2
(4-50)
1
1
= sin i + j sin i j
2
2
and
1
1
sini sin j = sin i + j + sin i j
(4-51)
2
2
results in:
({
})
(4-52)
1
1
cos i + j + cos i j
2
2
1
1
sin i sin j = cos i j cos i + j
2
2
cosi cos j =
(4-53)
and
1
1
1
1
cosi cos j +sin i sin j = cos i + j + cos i j + cos i j cos i + j
2
2
2
2
= cos i j
(4-54)
so that the expression for A ij can be written as:
(4-55)
ri,j+1 1
1
sin( i j )ln
+
cos i j ij .
2
ri, j 2
(4-56)
2/24/98
(4-57)
(4-58)
(4-59)
bi = V sin(i ) .
(4-60)
and
so that we get:
N
N
= cos i uv*ij sin j + vv*ij cos j sini u*vij cos j vv*ij sin j
j=1
(
N
j=1
j=1
(4-61)
*
*
= cosi cos j + sini sin j vv ij + cos i sin j sini cos j uvij
444 43
144 44244443
j =1 14 4442
a
b
N
a = cos i j
b = sin i j
(4-62)
(4-63)
j=1
2/24/98
ri,j+1
1 N
cos(
)ln
sin(
i
j
i
j ij .
2 j=1
ri, j
(4-64)
ri,j+1 1
1
sin(i j )ln
+
cos i j ij
2
ri, j 2
ri,j+1
1 N
cos(
)ln
sin(
i
j
i
j ij
2 j=1
ri, j
bi = V sin( i )
Step 4. Kutta Condition to get equation N+1
To complete the system of N+1 equations, we use the Kutta condition, which we previously
defined as:
u1 cos1 + v1 sin1 = uN cos N v N sin N
(4-66)
and substitute into this expression the formulas for the velocities due to the freestream and
singularities given in equation (4-31). In this case they are written as:
N
j =1
j=1
vv1 j
j=1
j=1
j=1
uN = V cos + q j us Nj + uvNj
vN = V sin + q j vsNj +
j=1
2/24/98
vvNj
j=1
(4-67)
N
N
V cos + q j us + uv cos1
1j
1j
j=1
j=1
N
N
+ V sin + q j vs1j + vv1 j sin1
j=1
j=1
(4-68)
N
N
+ V cos + q jus Nj + uv Nj cos N
j =1
j=1
N
N
and our goal will be to manipulate this expression into the form:
N
AN+1,jq j + AN+1,N+1
= bN +1
(4-69)
j=1
which is the N + 1st equation which completes the system for the N + 1 unknowns.
Start by regrouping terms in the above equation to write it in the form:
cos1 + vs1 j sin1 + usNj cos N + vs Nj sin N )q j
(1us4
1j
44 4444 4424 444 4444 43
N
j=1
AN +1, j
(4-70)
AN +1, N +1
(4-71)
and using the trigonometric identities to obtain the expression for bN+1 :
bN+1 = V cos (1 ) V cos( N )
2/24/98
(4-72)
(4-73)
(4-74)
(4-75)
sin j cos N
)
.
+ ( cos j sin1 sin j cos1 )v*s1 j
+ ( cos j sin N sin j cos N )v*sNj
(4-76)
(
)
cos j cos N + sin j sin N = cos ( j N )
cos j sin1 sin j cos1 = sin ( j 1)
cos j sinN sin j cos N = sin ( j N )
cos j cos1 + sin j sin1 = cos j 1
2/24/98
(4-77)
(4-78)
1 r1,j+1
ln
2 r1,j
us*Nj =
1,j
v*s1 j =
v*sNj =
1 rN, j+1
ln
2 rN, j
(4-79)
N, j
2
ln
2
2
r1,j
rN, j
sin j 1
2
1, j
sin j N
2
(4-80)
N,j
Finally, use symmetry and odd/even relations to write down the final form:
sin(1 j )1,j + sin( N j ) N,j
1
r1, j+1
rN ,j+1 .
AN+1, j =
2 cos(1 j )ln
cos( N j )ln
r1,j
rN, j
(4-81)
(4-82)
j=1
where we substitute in for the ( )* coordinate system, Eq. (4-32), and obtain:
*
*
*
*
uv1 j cos j vv1 j sin j cos1 + uv1 j sin j + vv1j cos j sin1
AN+1,N+1 =
(4-83)
*
*
*
*
or:
2/24/98
j
1
j
1 v1j
j
1
j
1 v1 j
144
44 244 443
144442444 43
N
cos( j 1 )
sin( j 1 )
AN+1,N+1 =
+
cos
cos
+
sin
sin
u
j=1
j
N
j
N v
j
N
j
N v Nj
42444 44
3 Nj 14 44442444 44
3
144 44
cos( j N )
sin(j N )
(4-84)
which is:
*
*
AN+1,N+1 =
,
* sin( )v*
+cos(
)u
j=1
j
N
v
j
N
v
Nj
Nj
N
and using odd/even trig relations we get the form given by Moran6 :
*
*
)u
j=1
1
j
v
N
j
v
1j
Nj
N
(4-86)
We now substitute the formulas derived above for the influence coefficients given in Eq. (442). The final equation is:
r1, j+1
rN ,j+1
1 N sin 1 j ln r + sin N j ln r
AN+1,N+1 =
i,j
N, j .
2 j =1
(4-86)
After substituting in the values of the velocities in terms of the singularity strengths, and
performing some algebraic manipulation, a form of the coefficients suitable for computations is
obtained.
The final equations associated with the Kutta condition are:
sin(1 j )1,j + sin( N j ) N,j
1
r1, j+1
rN ,j+1
AN+1, j =
2 cos(1 j )ln
cos( N j )ln
1,j
rN, j
(4-81)
r1, j+1
rN ,j+1
1 N sin 1 j ln r + sin N j ln r
AN+1,N+1 =
i,j
N, j
2 j =1
(4-86)
(4-72)
2/24/98
Aijq j + Ai,N+1 = bi
i = 1,...N
j=1
AN+1,jq j + AN+1,N+1
(4-87)
= bN +1
j=1
N
N
j=1
j=1
(4-88)
N
N
j=1
j=1
N
N
*
*
*
*
uti = V cos + usij cos j vsij sin j q j + uv ij cos j vvij sin j cosi
j=1
j=1
N
N
+ V sin + u*sij sin j + v*sij cos j q j + uv*ij sin j + vv*ij cos j
j =1
j=1
sin i
(4-89)
or:
2/24/98
{us*ij cos j cos i vs*ij sin j cos i + us*ij sin j sini + vs*ij cos j sini }q j .
N
j=1
{uv*ij cos j cosi v*vij sin j cos i + u*vij sin j sini + v*sij cos j sini }
N
j=1
(4-90)
Collecting terms:
uti = (cos cosi + sin sin i ) V
144 44244443
cos ( i )
+ (cos j cosi +sin j sin i )us*ij + (cos j sin i sin j cosi )vs*ij q j
14 4442444 43
14 4442444 43
j=1
cos( j i )
sin( j i )
(4-91)
+ (cos j cos i + sin j sini )u*vij + (cos j sin i sin j cos i )v*vij
144 4424 4443
14 4442444 43
j=1
cos( j i )
sin( j i )
which becomes:
N
(4-92)
j=1
Using the definitions of the ( )* influence coefficients, and some trigonometric identities, we
obtain the final result:
N
r
q
uti = cos(i )V + i sin( i j )ij cos(i j )ln i, j+1
2
ri, j
j=1
ri,j+1
sin(
)ln
+
cos(
i
j
i
j ij
2 j=1
ri, j
2/24/98
(4-93)
(4-94)
c 2
( N 1)
i = 1,..., N .
(4-95)
These locations are then altered when camber is added (see Eqns. (A-1) and (A-2) in App. A).
This approach is used to provide a smoothly varying distribution of panel node points which
concentrate points around the leading and trailing edges.
An example of the accuracy of program PANEL is given in Fig. 4-11, where the results
from PANEL for the NACA 4412 airfoil are compared with results obtained from an exact
conformal mapping of the airfoil (comments on the mapping methods are given in Chapter 9 on
Geometry and Grids. Conformal transformations can also be used to generate meshes of points for
use in field methods). The agreement is nearly perfect.
Numerical studies need to be conducted to determine how many panels are required to obtain
accurate results. Both forces and moments and pressure distributions should be examined.
2/24/98
-2.50
PANEL
-2.00
-1.50
-1.00
Cp
-0.50
0.00
0.50
1.00
-0.2
0.0
0.2
0.4
0.6
0.8
1.0
1.2
x/c
Figure 4-11. Comparison of results from program PANEL with an essentially exact
mapping solution for the NACA 4412 airfoil at 6 angle-of-attack.
You can select the number of panels used to represent the surface. How many should you
use? Most computational programs provide the user with freedom to decide how detailed
(expensive - in dollars or time) the calculations should be. One of the first things the user should
do is evaluate how detailed the calculation should be to obtain the level of accuracy desired. In the
PANEL code your control is through the number of panels used.
We check the sensitivity of the solution to the number of panels by comparing force and
moment results and pressure distributions with increasing numbers of panels. This is done using
two different methods. Figures 4-12 and 4-13 present the change of drag and lift, respectively,
using the first method. For PANEL, which uses an inviscid incompressible flowfield model, the
drag should be exactly zero. The drag coefficient found by integrating the pressures over the airfoil
is an indication of the error in the numerical scheme. The drag obtained using a surface (or
nearfield) pressure integration is a numerically sensitive calculation, and is a strict test of the
method. The figures show the drag going to zero, and the lift becoming constant as the number of
2/24/98
0.010
CD
0.008
0.006
0.004
0.002
0.000
20
40
60
80
No. of Panels
100
120
0.975
CL
0.970
0.965
0.960
0.955
0.950
20
40
60
80
No. of Panels
100
2/24/98
120
0.012
0.010
0.008
CD 0.006
0.004
0.002
0.000 0
0.01
0.02
0.03
0.04
0.05
0.06
1/n
Figure 4-14. Change of drag with the inverse of the number of panels.
0.980
0.975
CL
0.970
0.965
0.960
0.955
0.950
0.01
0.02
0.03
0.04
0.05
0.06
1/n
Figure 4-15. Change of lift with the inverse of the number of panels.
-0.240
-0.242
-0.244
Cm
-0.246
-0.248
-0.250
0.03
0.04
0.05
0.06
1/n
Figure 4-16. Change of pitching moment with the inverse of the number of panels.
2/24/98
0.01
0.02
0.00
1.00
0.0
0.2
0.4
0.6
0.8
1.0
x/c
Figure 4-17. Pressure distribution from progrm PANEL, 20 panels.
2/24/98
-5.00
NACA 0012 airfoil, = 8
-4.00
20 panels
60 panels
-3.00
-2.00
CP
-1.00
0.00
1.00
0.0
0.2
0.4
x/c
0.6
0.8
1.0
-5.00
NACA 0012 airfoil, = 8
-4.00
60 panels
100 panels
-3.00
-2.00
CP
-1.00
0.00
1.00
0.0
0.2
0.4
x/c
0.6
0.8
2/24/98
1.0
Having examined the convergence of the mathematical solution, we investigate the agreement
with experimental data. Figure 4-20 compares the lift coefficients from the inviscid solutions
obtained from PANEL with experimental data from Abbott and von Doenhof.12 Agreement is
good at low angles of attack, where the flow is fully attached. The agreement deteriorates as the
angle of attack increases, and viscous effects start to show up as a reduction in lift with increasing
angle of attack, until, finally, the airfoil stalls. The inviscid solutions from PANEL cannot capture
this part of the physics. The difference in the airfoil behavior at stall between the cambered and
uncambered airfoil will be discussed further in Chapter 10. Essentially, the differences arise due to
different flow separation locations on the different airfoils. The cambered airfoil separates at the
trailing edge first. Stall occurs gradually as the separation point moves forward on the airfoil with
increasing incidence. The uncambered airfoil stalls due to a sudden separation at the leading edge.
An examination of the difference in pressure distributions to be discussed next can be studied to
see why this might be the case.
2.50
2.00
1.50
1.00
0.50
CL, NACA 0012 - PANEL
CL, NACA 0012 - exp. data
0.00
-0.50
-5.0
5.0
10.0
15.0
20.0
25.0
Figure 4-20. Comparison of PANEL lift predictions with experimental data, (Ref. 12).
2/24/98
Cm
-0.05
c/4
-0.10
-0.15
Cm, NACA 0012 - PANEL
Cm, NACA 4412 - PANEL
Cm, NACA 0012 - exp. data
Cm, NACA 4412 - exp. data
-0.20
-0.25
-0.30
-5.0
0.0
5.0
10.0
15.0
20.0
25.0
Figure 4-21. Comparison of PANEL moment predictions with experimental data, (Ref. 12).
We do not compare the drag prediction from PANEL with experimental data. In twodimensional incompressible inviscid flow the drag is zero. In the actual case, drag arises from skin
friction effects, further additional form drag due to the small change of pressure on the body due to
the boundary layer (which primarily prevents full pressure recovery at the trailing edge), and drag
due to increasing viscous effects with increasing angle of attack. A well designed airfoil will have a
drag value very nearly equal to the skin friction and nearly invariant with incidence until the
maximum lift coefficient is approached.
In addition to the force and moment comparisons, we need to compare the pressure
distributions predicted with PANEL to experimental data. Figure 4-22 provides one example. The
NACA 4412 experimental pressure distribution is compared with PANEL predictions. In general
2/24/98
-0.4
Cp
-0.0
Predictions from PANEL
0.4
= 1.875
M = .191
Re = 720,000
transition free
0.8
1.2
0.0
0.2
0.4
x/c
0.6
0.8
1.0
1.2
2/24/98
-0.60
-0.60
PANEL
-0.40
FLO36
-0.20
Cp
0.00
FLO36
-0.40
PANEL
-0.20
Cp
0.00
0.20
0.20
NACA 651 -012
= 8.8
0.40
0.60
0.6
0.7
0.8
0.9
1.0
X/C
a. 6-series, cusped TE
0.40
1.1
0.60
0.6
0.8
0.9
1.0
1.1
X/C
b. 6A-series, finite TE angle
NACA 65(1)-012
NACA 65A012
-0.05
0.70
0.80
x/c
0.90
1.00
Figure 4-24. Comparison at the trailing edge of 6- and 6A-series airfoil geometries.
This case demonstrates a situation where this particular panel method is not accurate. Is this a
practical consideration? Yes and no. The 6-series airfoils were theoretically derived by specifying a
pressure distribution and determining the required shape. The small trailing edge angles (less than
half those of the 4-digit series), cusped shape, and the unobtainable zero thickness specified at the
trailing edge resulted in objections from the aircraft industry. These airfoils were very difficult to
use on operational aircraft. Subsequently, the 6A-series airfoils were introduced to remedy the
problem. These airfoils had larger trailing edge angles (approximately the same as the 4-digit
series), and were made up of nearly straight (or flat) surfaces over the last 20% of the airfoil. Most
applications of 6-series airfoils today actually use the modified 6A-series thickness distribution.
This is an area where the user should check the performance of a particular panel method.
2/24/98
-2.00
Expansion/recovery around leading edge
(minimum pressure or max velocity,
first appearance of sonic flow)
-1.50
-1.00
CP
-0.50
lower surface
0.00
0.50
0.1
0.3
0.5
x/c
0.7
0.9
1.1
Figure 4-25. Key areas of interest when examining airfoil pressure distributions.
Remember that we are making an incompressible, inviscid analysis when we are using
program PANEL. Thus, in this section we examine the basic characteristics of airfoils from that
point of view. We will examine viscous and compressibility effects in subsequent chapters, when
we have the tools to conduct numerical experiments. However, the best way to understand airfoil
characteristics from an engineering standpoint is to examine the inviscid properties, and then
consider changes in properties due to the effects of viscosity. Controlling the pressure distribution
through selection of the geometry, the aerodynamicist controls, or suppresses, adverse viscous
effects. The mental concept of the flow best starts as a flowfield driven by the pressure distribution
that would exist if there were no viscous effects. The airfoil characteristics then change by the
2/24/98
2/24/98
-5.00
NACA 0012 airfoil
Inviscid calculation from PANEL
-4.00
= 0
-3.00
=4
CP
=8
-2.00
-1.00
0.00
1.00
-0.1
0.1
0.3
0.5
x/c
0.7
0.9
1.1
2/24/98
-0.30
-0.20
-0.10
y/c
0.00
0.10
0.20
0.30
-0.1
0.1
0.3
0.5
x/c
0.7
0.9
Figure 4-27. Comparison of NACA 4-digit airfoils of 6, 12, and 18% thicknesses.
-1.00
Inviscid calculation from PANEL
-0.50
C
0.00
0.50
1.00
-0.1
2/24/98
NACA 0006, = 0
NACA 0012, = 0
NACA 0018, = 0
0.1
0.3
0.5
0.7
0.9
1.1
x/c
Figure 4-28. Effect of airfoil thickness on the pressure distribution at zero lift.
1.1
-3.00
Inviscid calculation from PANEL
-2.50
NACA 0006, = 4
NACA 0012, = 4
-2.00
NACA 0018, = 4
-1.50
C
-1.00
-0.50
0.00
0.50
1.00
-0.1
0.1
0.3
0.5
x/c
0.7
0.9
1.1
The next effect to examine is camber. Figure 4-30 compares the shapes of the NACA 0012
and 4412 airfoils. The pressure distributions on the cambered airfoil for two different angles of
attack are shown in Figure 4-31. Note the role of camber in obtaining lift without producing a
leading edge expansion followed by a rapid recompression immediately behind the expansion. This
reduces the possibility of leading edge separation.
2/24/98
0.30
0.20
0.10
y/c 0.00
-0.10
NACA 0012 (max t/c = 12%)
NACA 4412 foil (max t/c = 12%)
-0.20
-0.30
-0.1
0.1
0.3
0.5
x/c
0.7
0.9
1.1
-2.00
Inviscid calculation from PANEL
-1.50
NACA 4412, = 0
NACA 4412, = 4
-1.00
CP
-0.50
0.00
0.50
Note: For a comparison of cambered and uncambered
presuure distributions at the same lift, see Fig. 4-32.
1.00-0.1
0.5
0.7
0.9
1.1
x/c
Figure 4-31. Effect of angle of attack on cambered airfoil pressure distributions at low lift.
2/24/98
0.1
0.3
-2.00
Inviscid calculation from PANEL
-1.50
NACA 0012, = 4
NACA 4412, = 0
-1.00
-0.50
CP
0.00
0.50
1.00
-0.1
2/24/98
0.1
0.3
0.5
0.7
0.9
x/c
Figure 4-32. Camber effects on airfoil pressure distributions at CL = 0.48.
1.1
-4.00
Inviscid calculations from PANEL
NACA 0012, = 8
NACA 4412, = 4
-3.00
-2.00
C
-1.00
0.00
1.00
-0.1
0.1
0.3
0.5
0.7
0.9
x/c
Figure 4-33. Camber effects airfoil pressure distributions at CL = 0.96.
1.1
-6.00
Inviscid calculations from PANEL
-5.00
NACA 0012, = 12
NACA 4412, = 8
-4.00
-3.00
C
-2.00
-1.00
0.00
1.00
-0.1
2/24/98
0.1
0.3
0.5
0.7
0.9
x/c
Figure 4-34. Camber effects airfoil pressure distributions at CL = 1.43.
1.1
Finally, we examine the effect of extreme aft camber, which was part of the design strategy of
Whitcomb when the so-called NASA supercritical airfoils were developed. This effect can be
simulated using the NACA 6712 airfoil, as shown in Figure 4-35. The resulting pressure
distribution is given in Figure 4-36. Note that the aft camber opens up the pressure distribution
near the trailing edge. Two adverse properties of this type of pressure distribution are the large zero
lift pitching moment and the delayed and then rapid pressure recovery on the upper surface. This
type of pressure recovery is a very poor way to try to achieve a significant pressure recovery
because the boundary layer will separate early. Whitcombs design work primarily improved the
pressure recovery curve.
0.15
y/c 0.05
-0.05
-0.1
0.1
0.3
0.5
x/c
0.7
0.9
1.1
-1.00
-0.50
CP
0.00
0.50
NACA 6712
1.00
-0.1
2/24/98
0.1
0.3
0.5
0.7
0.9
x/c
Figure 4-36. Example of the use of aft camber to "open up"
the pressure distribution near the trailing edge.
1.1
0.2
0.4
x/c
0.6
0.8
1.0
-0.50
Cp
0.00
0.50
GA(W)-1
= 0
1.00
0.0
0.2
0.4
0.6
0.8
1.0
X/C
2/24/98
2/24/98
2/24/98
Impermeable Surface
Tail Wake
Nor Shown
Body Wake
Carry-Over
Wakes
Wing Wake
a) wing-body-tail configuration panel scheme with wakes
Impermeable
Surfaces
Body Wake
Tail-Body
Carry-Over
Wake
Tail Wake
Wing Wake
Wing-Body
Carry-Over
Wake
2/24/98
2/24/98
Figure 4-41. The Boeing 737-300 relative to the model 737-200 (Ref.24).
Figure 4-42. The panel representation of the 737-300 with 15 flap deflection (Ref. 4).
2/24/98
An understanding of the wing flowfield for two different takeoff flap settings was desired.
The cases are flaps 15, the normal takeoff setting, and flaps 1, the high altitude, hot day
setting. The work was conducted in concert with the flight test program to provide insight into the
flight test results by providing complete flowfield details not available from the flight test. The
computational models used 1750 panels for flaps 1 and 2900 panels for flaps 15. The modeling
used to simulate this flowfield illustrates typical idealizations employed when applying panels
methods to actual aircraft. Although typical, it is one of the most geometrically complicated
examples ever published.
Figure 4-43 shows the wing leading edge and nacelle. The inboard Krueger flap was actually
modeled as a doublet of zero thickness. The position was adjusted slightly to allow the doublet
sheet to provide a simple matching of the trailing edge of the Krueger and the leading edge of the
wing. These types of slight adjustments to keep panel schemes relatively simple are commonly
used. The outboard leading and trailing edge flap geometries were also modified for use in this
inviscid simulation. Figure 4-44 a) shows the actual and computational flaps 1 geometry. In this
case the airfoil was modeled as a single element airfoil. The flaps 15 trailing edge comparison
between the actual and computational geometry is shown in Fig. 4-44 b). The triple slotted flap
was modeled as a single element flap. At this setting the gap between the forward vane and main
flap is closed, and the gap between the main and aft flap is very small.
Figure 4-43. Inboard wing leading edge and nacelle details (Ref. 24).
2/24/98
a) Comparison of actual and computational wing geometry for the flaps 1 case (Ref. 24).
Actual Geometry
Computational Geometry
b) Actual and computational trailing edge geometry for the flaps 15 case (Ref. 4).
Figure 4-44. Examples of computational modeling for a real application.
Several three-dimensional modeling considerations also required attention. In the flaps 1 case
shown in Fig. 4-45, spanwise discontinuities included the end of the outboard leading edge slat
and trailing edge discontinuities at the back of the nacelle installation (called the thrust gate)
between the inboard and outboard flaps. At the outboard leading edge the edges of the slat and
wing were paneled to prevent leakage. A 0.1 inch gap was left between these surfaces. At the
trailing edge discontinuity a wake was included to model a continuous trailing edge from which a
trailing vortex sheet could be shed.
2/24/98
Figure 4-45. Spanwise discontinuity details requiring modeling for flaps 1 case (Ref. 24).
Similar considerations are required for the flaps 15. Here, special care was taken to make sure
that the configuration was closed, and contained no holes in the surface at the ends of the flap
segments.
Another consideration is the nacelle model. This requires the specification of the inlet flow at
the engine face, a model of the strut wake, and both the outer bypass air plume and the primary
wake from the inner hot gas jet. Figure 4-46 provides the details.
2/24/98
a) flaps 1 case
b) flaps 15 case
Figure 4-47. Spanwise distribution of lift coefficient on the Boeing 737-300 (Ref.24).
2/24/98
a) flaps 1 case
b) flaps 15 case
Figure 4-48. Comparison of section lift coefficient change with angle of attack(Ref.24)
2/24/98
Figure 4-49 completes this example by presenting the comparison of pressure distributions for
the two cases at four spanwise stations. The flaps 1 case agreement is generally good. Calculations
are presented for both the actual angle of attack, and the angle of attack which matches the lift
coefficient. Matching lift coefficient instead of angle of attack is a common practice in
computational aerodynamics. Considering the simplifications made to the geometry and the
absence of the simulation of viscous effects the agreement is very good. The flaps 15 case starts to
show the problems that arise from these simplifications. This is a good example of the use of a
panel method. It illustrates almost all of the considerations that must be addressed in actual
applications.
a) flaps 1 case
b) flaps 15 case
Figure 4-49. Comparison of pressure distributions between flight and computations for the 737300, solid line is PAN AIR at flight lift, dashed line is PAN AIR at flight angle of attack (Ref. 24).
2/24/98
C pcrit
1 2 1
2 1 + 2 M
=
2 1 +1
M
(4-96)
2/24/98
2/24/98
2/24/98
Table 4 - 1
Comparison of Some Major Panel Method Programs: Early Codes
Originator and
Method Name
Year
1964
Panel
Geometry
Source
Type
Doublet
Type
Boundary
Conditions
Restrictions
flat
constant
none
specification
of normal
flow
non-lifting
wings and
bodies only
flat
none
constant
normal flow
planar wings
only
Woodward
(Woodward I)
1967
flat
constant
linear
normal flow
wings must
be planar
Rubbert and
Saaris4
(Boeing A-230)
1968
flat
constant
constant
normal flow
nearly
constant
panel density
Hess I 5
1972
flat
constant
linear
normal flow
wings and
bodies only
USSAERO 6
(Woodward II)
1973
flat
W12SC3 7
(Grumman)
1983
flat
mixed design
and
analysis
Comments
mainly
supersonic,
includes
design &
optimization
subsonic and
supersonic,
analysis only
combines
Woodward
1&2
features
Hess, J.L., and Smith, A.M.O., "Calculation of Nonlifting Potential Flow About Arbitrary
Three-Dimensional Bodies," Douglas Report ES40622, Douglas Aircraft Company, 1962.
Rubbert, P.E., "Theoretical Characteristics of Arbitrary Wings by a Nonplanar Vortex Lattice Method,"
Boeing Report D6-9244, The Boeing Company, 1964.
Woodward, F.A., Tinoco, E.N., and Larsen, J.W., "Analysis and Design of Supersonic Wing-Body
Combinations, Including Flow Properties in the Near Field," Part I - Theory and Application, NASA
CR-73106, 1967.
Rubbert, P.E., and Saaris, G.R., "A General Three-Dimensional Potential Flow Method Applied to
V/STOL Aerodynamics," SAE Paper No. 680304, 1968.
Hess, J.L., "Calculation of Potential Flow About Arbitrary 3-D Lifting Bodies," Douglas Report
MDC-J5679-01, October 1972.
Woodward, F.A., "An Improved Method for the Aerodynamic Analysis of Wing-Body-Tail Configurations
in Subsonic and Supersonic Flow," NASA CR-2228, Parts I and II, 1973.
Mason, W.H., and Rosen, B.S., "The COREL and W12SC3 Computer Programs for Supersonic Wing
Design and Analysis," NASA CR 3676, 1983 (contributions by A. Cenko and J. Malone acknowledged).
from Magnus and Epton, NASA CR 3251, April 1980 (with extensions)
2/24/98
Table 4 - 2
Comparison of Some Major Panel Method Programs: Advanced Methods
Originator and
Method Name
Roberts and
Rundle1
Mercer, Weber
and Lesford 2
Year
1973 paraboloidal
1973
Panel
Geometry
Source
Type
quadratic
Doublet
Type
quadratic
Boundary
Conditions
normal flow
normal flow
in least
planar wings
squares sense
subsonic/
supersonic,
cubic
spanwise,
quadratic
chordwise
none
smooth,
cubic/
quadratic
continuous,
hyperboloidal
constant
constant
potential
linear
quadratic
normal flow
normal flow
arbitrary in
,
Ehlers and
Rubbert 5
(Mach line
paneling)
1976
flat
linear
continuous
quadratic
Ehlers et al. 6
(PAN AIR
"pilot code")
1977
continuous
piecewise
flat
linear
continuous
quadratic
Comments
numerical
integration,
very
expensive
flat
1975 paraboloidal
Restrictions
no thin
configurations
unsteady
planar
wings,
special
paneling
supersonic
flow
subsonic and
supersonic
Roberts, A., and Rundle, K., "Computation of First Order Compressible Flow About Wing-Body
Configurations," AERO MA No. 20, British Aircraft Corporation, February, 1973.
Mercer, J.E., Weber, J.A., and Lesford, E.P., "Aerodynamic Influence Coefficient Method Using
Singularity Splines," NASA CR-2423, May 1974.
Morino, L., and Kuo, C-C, "Subsonic Potential Aerodynamics for Complex Configurations: A General
Theory," AIAA Journal, Vol. 12, No. 2, pp 191-197, February, 1974.
Johnson, F.T., and Rubbert, P.E., "Advanced Panel-Type Influence Coefficient Methods Applied to
Subsonic Flow," AIAA Paper No. 75-50, January 1975.
Ehlers, F.E., and Rubbert, P.E., "A Mach Line Panel Method for Computing the Linearized Supersonic
Flow," NASA CR-152126, 1979.
Ehlers, F.E., Epton, M.A., Johnson, F.T., Magnus, A.E., and Rubbert, P.E., "A Higher Order Panel
Method for Linearized Flow," NASA CR-3062, 1979.
from Magnus and Epton, NASA CR 3251, April 1980 (with extensions)
2/24/98
Table 4-3
Comparison of Some Major Panel Method Programs: Production Codes
Originator and
Method Name
Year
Panel
Geometry
Source
Type
Doublet
Type
MCAIR 1
(McDonnell)
1980
flat
constant
quadratic
PAN AIR 2
(Boeing)
1980
continuous
piecewise
flat
Hess II 3
(Douglas)
1981
parabolic
linear
quadratic
normal flow
VSAERO4
(AMI)
1981
flat
constant
constant
exterior and
interior
normal flow
QUADPAN 5
(Lockheed)
1981
flat
constant
constant
PMARC 6
(NASA Ames)
1988
flat
constant
constant
continuous continuous
linear
quadratic
Boundary
Conditions
Restrictions
Comments
design
option
arbitrary in
,
subsonic and
supersonic
subsonic
unsteady,
wake rollup
Bristow, D.R., "Development of Panel Methods for Subsonic Analysis and Design," NASA CR 3234,
1980.
Magnus, A.E., and Epton, M.A., "PAN AIR - A Computer Program for Predicting Subsonic or
Supersonic Linear Potential Flows About Arbitrary Configurations Using a Higher Order Panel Method,"
Volume I - Theory Document (Version 1.0), NASA CR 3251, 1980.
Hess, J.L., and Friedman, D.M., "An Improved Higher Order Panel Method for Three Dimensional
Lifting Flow," Douglas Aircraft Co. Report No. NADC-79277-60, 1981.
Maskew, B., "Prediction of Subsonic Aerodynamic Characteristics: A Case for Lower Order Panel
Methods," AIAA Paper No. 81-0252, 1981.
Coopersmith, R.M., Youngren, H.H., and Bouchard, E.E., "Quadrilateral Element Panel Method
(QUADPAN)," Lockheed-California LR 29671, 1981.
Ashby, D.L., Dudley, M.R., and Iguchi, S.K., "Development and Validation of an Advanced Low-Order
Panel Method," NASA TM 101024, 1988 (also TM 102851, 1990).
from Magnus and Epton, NASA CR 3251, April 1980 (with extensions)
2/24/98
4.12 Exercises
1. Program PANEL.
a) Obtain a copy of program PANEL and the sample case.
b) Convert PANEL to run on your PC.
c) Run the sample case: NACA 4412, 20 pts. upper, 20 pts. lower, = 4 , and verify
against sample case.
d) Document
i) compile time required on your PC
(cite computer and compiler used)
ii) the execution time for the sample case
iii) the accuracy relative to the sample case.
iv) the exact modifications required to make the code
work on your computer
2. Start work on program PANEL
a) Save a reference copy of the working code!
b) Check convergence with panels (NLOWER+NUPPER must be less than 100
currently). How many panels do you need to get results independent of the number of
panels? What happens to the computer time as the number of panels increases?
c) Check the coordinates generated by the airfoil routine vs. exact (consider using the
NACA 0012, see App. A for geometry definition), including examination of the
coordinates at the trailing edge. This is best done by making a table of exact and
computed values at selected values of x/c. What did you find out?
d) Locate the source strengths, and sum the source strengths x panel lengths to get the
total source strength. Does it sum to zero? Should it?
e) Where is the moment reference center in this code?
Submit an assessment of your findings.
3. Modify program PANEL:
You need a version of PANEL that will allow you to compute the pressure
distribution on arbitrary airfoils. This exercise will give you this capability. Modify
the code to interpolate input airfoil points to the program defined surface points, x/c.
The resulting code should:
a) accept arbitrary airfoil input data
b) echo all the input data on the output
c) generate an output file for post processing
(both for plotting and as the input to a boundary layer code)
d) output Cm about the airfoil quarter chord point.
Hint: Dont alter the panel distribution. The paneling scheme should be independent
of the input distribution of airfoil coordinates. This produces a much more general
and accurate program. This problem is usually solved by finding both the x/c and y/c
2/24/98
(1 x /c )
x /c
2/24/98
4.13 References
1
Anderson, John P., Jr., Modern Compressible Flow, 2nd Ed., McGraw-Hill, New York,
1990, pp. 258-269.
2
Hess, J. L., Panel Methods in Computational Fluid Dynamics, Annual Review of Fluid
Mechanics, Vol. 22, 1990, pp. 255-274.
3
Hess, J.L., Linear Potential Schemes, Applied Computational Aerodynamics, P.A. Henne,
ed., AIAA, Washington, 1990. pp.21-36.
4
Erickson, L.L., Panel MethodsAn Introduction, NASA TP-2995, Dec. 1990.
5
Katz, J., and Plotkin, A., Low-Speed Aerodynamics From Wing Theory to Panel Methods,
McGraw-Hill, Inc., New York, 1991.
6
Karamcheti, K., Principles of Ideal-Fluid Aerodynamics, John Wiley & Sons, New York,
1966.
8
Ashley, H, and Landahl, M., Aerodynamics of Wings and Bodies, Addison-Wesley, Reading,
1965 (republished in paperback by Dover Publishing).
9
Curle, N., and Davis, H.J., Modern Fluid Dynamics, Volume 1: Incompressible Flow, Van
Nostrand, London, 1968.
10
Houghton, E.L., and Carpenter, P.W., Aerodynamics for Engineering Students, 4th Ed.,
Halsted Press, New York, 1993, pp. 257-265, 203-211.
11
Kuethe, A.M., and Chow, C-Y., Foundations of Aerodynamics, 4th Ed., John Wiley, New
York, 1986, pp. 128-137.
12
Abbott, I.H., and von Doenhoff, A.E., Theory of Wing Sections, Dover, New York, 1959.
13
14
2/24/98
21
Hess, J. L. "The Problem of Three-Dimensional Lifting Potential Flow and Its Solution by
Means of Surface Singularity Distributions", Computer Methods in Applied Mechanics and
Engineering 4 (1974) pp. 283-319.
22
Carmichael, R.L., and Erickson, L.L., "PAN AIR - A Higher Order Panel Method for
Predicting Subsonic or Supersonic Linear Potential Flows About Arbitrary Configurations," AIAA
Paper No. 81-1255, June 1981.
23
PAN AIR Users Class Short Course Presentation Material, 1981.
24
Tinoco, E.N., Ball, D.N., and Rice, F.A. II, PAN AIR Analysis of a Transport High-Lift
Configuration, Journal of Aircraft, Vol. 24, No. 3, March 1987, pp. 181-188.
25
Magnus, A.E., and Epton, M.A., "PAN AIR - A Computer Program for Predicting Subsonic or
Supersonic Linear Potential Flows About Arbitrary Configurations Using a Higher Order Panel
Method," Volume I - Theory Document (Version 1.0), NASA CR 3251, April 1980.
26
Margason, R.J., Kjelgaard, S.O., Sellers, W.L., Moriis, C.E.K., Jr., Walkley, K.B., and
Shields, E.W., Subsonic Panel Methods - A Comparison of Several Production Codes, AIAA
Paper 85-0280, Jan. 1985.
27
Ashby, D.L., Dudley, M.R., Iguchi, S.K., Browne, L., and Katz, J., Potential Flow Theory
and Operation Guide for the Panel Code PMARC, NASA TM 102851, Jan. 1991.
2/24/98
3/11/98
NASA TN D-6142
NASA TN D-7921
NASA TM 83303
1971
1975
1982
6-1
3/11/98
Vsin
V cos
Figure 6-1. Basic coordinate system for boundary condition analysis.
(6-2)
q(x,y)
123
a disturbance velocity
where q is a disturbance velocity with components u and v. If we assume irrotational flow, then
these components are described in terms of a scalar potential function, u = x and v = y. The
total velocity V then becomes in terms of velocity components:
uTOT = V cos + u(x,y)
vTOT = V sin + v(x, y)
(6-3)
(6-4)
3/11/98
(6-5)
df (x)
{ y f (x)} =
x
dx
.
{ y f (x) } = 1
y
(6-7)
(V cos + u )
df
+ (V sin + v ) = 0
dx
(6-8)
df
V
dx
v
u df
= 1+
V V dx
(6-11)
(6-12)
the linearized boundary condition is obtained by neglecting u/V compared with unity
(consistent with the previous approximations). With this assumption, the linearized boundary
condition becomes:
v
df
(6-13)
=
on y = f (x) .
V dx
*
Observe that even when the flowfield model is defined by a linear partial differential equation, an
assumption which we have not yet made, the boundary condition can make the problem nonlinear.
3/11/98
v
+... .
y y=0
(6-14)
For the thin surfaces under consideration, f(x) is small, and because the disturbances are assumed
small, v/y is also small. For example, assume that v and v/y are the same size, equal to 0.1,
and df/dy is also about 0.1. The relation between v on the airfoil surface and the axis is:
v{x, y = f(x)} = (.1) +(.1)(.1) = .1 + .01
{ .
(6-15)
neglect
(6-16)
We now apply both the upper and lower surface boundary conditions on the axis y = 0, and
distinguish between the upper and lower surface shapes by using:
f = fu
f = fl
(6-17)
Using Eq. (6-17), we write the upper and lower surface boundary conditions as:
v x,0 +
V
df
= u ,
dx
up
v x,0
V
dfl
.
dx
(6-18)
low
The simplification introduced by applying boundary conditions on constant coordinate surfaces justifies
the use of rather elaborate tranformations, which will be dicussed in more detail in Chap. 9, Geometry and
Aerodynamics.
3/11/98
(6-19)
and the general problem is then divided into the sum of three parts as shown in Fig. 6-2.
General
Thickness
Camber
airfoil
at = 0
at = 0
at
Figure 6-2. Decomposition of a general shape at incidence.
Flat plate
at
The decomposition of the problem is somewhat arbitrary. Camber could also be considered
to include angle of attack effects using the boundary condition relations given above, the sign is
the same for both the upper and lower surface. The aerodynamicist must keep track of details for
a particular problem. To proceed further, we make use of the basic vortex lattice method
assumption: the flowfield is governed by a linear partial differential equation (Laplaces
equation). Superposition allows us to solve the problem in pieces and add up the contributions
from the various parts of the problem. This results in the final form of the thin airfoil theory
boundary conditions:
v x,0+
V
v
x,0
V
df c dft
+
dx dx
up
.
=
(6-20)
df c dft
dx dx
low
The problem can be solved for the various contributions and the contributions are added together
to obtain the complete solution. If thickness is neglected the boundary conditions are the same
for the upper and lower surface.
3/11/98
(6-21)
and we expand the velocity considering disturbances to the freestream velocity using the
approximations discussed above:
V 2 = (V cos + u) + ( V sin + v)
2
V
Expanding:
(6-22)
V2
u
u2
v v2
2
=1
+
2
+
+
+
2
+
.
V V2
V V2
V2
(6-23)
u
u2
v v2
2
CP = 1 1 + 2
+
+ + 2
+
V V2
V V2
= 1 1 2
u2
(6-24)
v2
u
v
2 2 2
V V
V V2
and if , u/V and v/V are << 1, then the last four terms can be neglected in comparison with
the third term. The final result is:
C p = 2
u
.
V
(6-25)
This is the linearized or thin airfoil theory pressure formula. From experience gained
comparing various computational results, Ive found that this formula is a slightly more severe
restriction on the accuracy of the solution than the linearized boundary condition. Equation (625) shows that under the small disturbance approximation, the pressure is a linear function of u,
and we can add the Cp contribution from thickness, camber, and angle of attack by
superposition. A similar derivation can be used to show that Eq. (6-25) is also valid for
compressible flow up to moderate supersonic speeds.
3/11/98
UPPER
Using superposition, the pressures can be obtained as the contributions from wing thickness,
camber, and angle of attack effects:
C pLOWER = C pTHICKNESS + C pCAMBER + C pANGLE OF ATTACK
C pUPPER = C pTHICKNESS C pCAMBER C pANGLE OF ATTACK
so that:
(6-27)
(6-28)
Equation (6-28) demonstrates that for cases where the linearized pressure coefficient relation is
valid, thickness does not contribute to lift to 1st order in the velocity disturbance!
The importance of this analysis is that we have shown:
1. how the lifting effects can be obtained without considering thickness, and
2. that the cambered surface boundary conditions can be applied on a flat coordinate
surface, resulting in an easy to apply boundary condition.
The principles demonstrated here for transfer and linearization of boundary conditions can be
applied in a variety of situations other than the application in vortex lattice methods. Often this
idea can be used to handle complicated geometries that cant easily be treated exactly.
The analysis here produced an entirely consistent problem formulation. This includes the
linearization of the boundary condition, the transfer of the boundary condition, and the
approximation between velocity and pressure. All approximations are consistent with each other.
Improving one of these approximations without improving them all in a consistent manner may
actually lead to worse results. Sometimes you can make agreement with data better, sometimes it
may get worse. You have to be careful when trying to improve theory on an ad hoc basis.
3/11/98
(6-29)
V=
e
2 r
where V is the irrotational vortex flow illustrated in Fig. 6-3.
V
1/r
V
Vortex normal
to page
a) streamlines
b) velocity distribution
Figure 6-3. The point vortex.
What is the extension of the point vortex idea to the case of a general three-dimensional
vortex filament? Consider the flowfield induced by the vortex filament shown in Fig. 6-4, which
defines the nomenclature.
3/11/98
vortex
filament
dl
q
rpq
p
.
4 r 3
pq
To obtain the velocity induced by the entire length of the filament, integrate over the length
of the vortex filament (or line) recalling that is constant. We obtain:
r r
r
dl rpq
(6-31)
Vp =
r 3 .
4
rpq
To illustrate the evaluation of this integral we give the details for several important
examples. The vector cross product definition is reviewed in the sidebar below. Reviewing the
cross product properties we see that the velocity direction dV p induced by the segment of the
vortex filament dl is perpendicular to the plane defined by dl and rpq, and its magnitude is
computed from Eq. (6-31).
Case #1: the infinitely long straight vortex.
In the first illustration of the computation of the induced velocity using the Biot-Savart Law, we
consider the case of an infinitely long straight vortex filament. The notation is given in Fig. 6-5.
3/11/98
A review: the meaning of the cross product. What does a x b mean? Consider the following
sketch:
c
b
a
Here, the vectors a and b form a plane, and the result of the cross product operation is a vector c,
where c is perpendicular to the plane defined by a and b. The value is given by:
c = a b = a b sin e
and e is perpendicular to the plane of a and b. One consequence of this is that if a and b are
parallel, then a x b = 0.
a b = area of the parallelogram
Also:
and
i
a b = Ax
Bx
j
Ay
By
k
Az = Ay Bz Az By i ( Ax Bz Az Bx ) j + Ax By Ay Bx k
Bz
dl
h
rpq
-
.
Figure 6-5. The infinitely long straight vortex.
3/11/98
dl rpq sin V
3
4
V
rpq
V sin dl
2
4 V rpq
(6-33)
Next, simplify the above expressions so they can be readily evaluated. Use the nomenclature
shown in Fig. 6-6.
p
r pq
h
l1
q
l
(6-34)
(6-35)
(6-36)
h
d(l1 l) = d
= d(hcot)
tan
dl = h d(cot )
which is
(6-37)
dl
d
=h
(cot) = hcosec2
d
d
(6-38)
V sin dl
2
4 V rpq
V sin hcosec2 d
.
2
4 V
h
sin
sin
=
d
(direction understood)
4 h
=
Thus we integrate:
Vp =
sin
d =
4
h
4 h
(6-39)
sin d
(6-40)
=0
where here the limits of integration would change if you were to consider a finite straight length.
Carrying out the integration:
Vp =
[cos ] =
=0
4 h
=
[(1) (1)]
4 h
and:
(6-41)
Vp =
2 h
which agrees with the two dimensional result.
Although a vortex cannot end in a fluid, we can construct expressions for infinitely long
vortex lines made up of a series of connected straight line segments by combining expressions
developed using the method illustrated here. To do this we simply change the limits of
integration. Two cases are extremely useful for construction of vortex systems, and the formulas
are given here without derivation.
Case #2: the semi-infinite vortex.
This expression is useful for modeling the vortex extending from the wing to downstream
infinity.
+
dl
0
h
p
Vp =
3/11/98
(1 + cos 0 )e
4h
(6-42)
dl
h
p
Vp =
(cos1 cos2 )e
4h
(6-43)
3/11/98
dl
r2
1
r1
rp
C
ro = AB
r1 = AC
r2 = BC
(6-46)
Vp =
(cos1 cos 2 )e
4 rp
so that, substituting using the definition in Eq. (6-44), we get
Vp =
3/11/98
r0 r0 r1 r0 r2 r1 r2
4 r1 r2 r0 r1 r0 r2 r1 r2
(6-47)
rp
r
2
A
r
1
2AABC = Ap
ro rp = r1 r2
which can be rewritten to obtain the expression given by Bertin and Smith:14
r r
rp = 1 2
r0
Collecting terms and making use of the vector identity provided in the sidebar, we obtain the
Bertin and Smith statement of the Biot-Savart Law for a finite length vortex segment:
Vp =
r1 r2 r1 r2
r
.
4 r1 r2 2 0 r1 r2
(6-48)
For a single infinite length horseshoe vortex we will use three segments, each using the
formula given above. The nomenclature is given in the sketch below. The primary points are the
3/11/98
(6-49)
where the total velocity is the sum of the contributions from the three separate straight line
vortex segments making up the horseshoe vortex, as shown in Fig. 6-11.
z
y
B
n
n
A
(6-50)
We now write the expression for the velocity field at a general point in space (x,y,z) due to
the horseshoe vortex system. At C(x,y,z) find the induced velocity due to each vortex segment.
Start with AB, and use Fig. 6-12.
Now we define the vectors as:
3/11/98
(6-51)
r
0
r
2
A
r
1
C(x,y,z)
Figure 6-12. Velocity induced at Point C due to the vortex between A and B.
and we simply substitute into:
VC =
r1 r2
r1
r2
r
.
0
0
4 r1 r2 2
r1
r2
42444
3
1
424
3 144
(6-52)
(6-53)
where and are lengthy expressions. By following the vector definitions, Eqn. (6-53) can be
written in Cartesian coordinates. The vector is:
=
r1 r2
r1 r2
+
x
x
y
x
y
y
k
(
)
(
)
(
)
(
)
[
]
1n
2n
2n
1n
.
=
( y y )( z z ) ( y y )( z z ) 2
1n
2n
2n
1n ]
[
2
+ [( x x1n )(z z2n ) ( x x2n )(z z1n )]
3/11/98
(6-54)
= ro
r1
r
ro 2
r1
r2
[( x2n x1n )( x x1n ) + ( y2n y1n )( y y1n ) + (z2n z1n )(z z1n )]
( x x1n )
+ ( y y1n ) + ( z z1n )
2
(6-55)
We find the contributions of the trailing vortex legs using the same formula, but redefining
the points 1 and 2. Then, keeping the 1 and 2 notation, define a downstream point, 3 and let x3
go to infinity. Thus the trailing vortex legs are given by:
1.0 +
(6-56)
and
4 ( z z2n )2 + (y2n y )2
1.0 +
.(6-57)
( x x 2n )2 + ( y y2n )2 + (z z2n )2
x x2n
Note that n is contained linearly in each expression, so that the expression given above can
be arranged much more compactly by using (6-49) with (6-52), (6-56), and (6-57) as:
Vm = Cmnn
(6-58)
and Cmn is an influence coefficient for the nth horseshoe vortex effect at the location m,
including all three segments.
Now that we can compute the induced velocity field of a horseshoe vortex, we need to
decide where to place the horseshoe vortices to represent a lifting surface.
6.6 Selection of Control Point/Vortex Location
Since we are interested in using the horseshoe vortex defined above to represent a lifting
surface, we need to examine exactly how this might be done. In particular: where do you locate
the vortex, and where do you locate a control point to satisfy the surface boundary condition?
Tradition has been to determine their locations by comparison with known results. In particular,
we use two dimensional test cases, and then apply them directly to the three dimensional case.
3/11/98
X
b
a
r
c
Figure 6-13. The notation for control point and vortex location analysis.
The velocity at the control point, cp, due to the point vortex is:
vcp =
.
2r
(6-59)
V
dx
(6-60)
v BC
=
V
(6-61)
v BC = V .
(6-62)
= V
2r
(6-63)
or:
Now, we equate vBC and vcp:
3/11/98
.
2rV
(6-64)
(6-65)
L = 12 V2 c{ 2
.
123 S {
C
(6-66)
ref
Equate the expressions for lift, Eqns. (6-65) and (6-66) and substitute for using the
expression in Eqn. (6-64) given above:
V = 12 V2c2
= 12 V2c2
1=
2rV
(6-67)
1c
2r
and finally:
c
(6-68)
.
2
This defines the relation between the vortex placement and the control point in order for the
single vortex model to reproduce the theoretical lift of an airfoil predicted by thin airfoil theory.
Since the flat plate has constant (zero) camber everywhere, this case doesnt pin down placement
(distance of the vortex from the leading edge) completely. Intuitively, the vortex should be
located at the quarter-chord point since that is the location of the aerodynamic center of a thin
flat plate airfoil. The next example is used to determine the placement of the vortex.
r=
2 ( a b )
(6-69)
v BC = V c .
dx
3/11/98
(6-70)
df
= c .
2 ( a b)V dx
(6-71)
(6-72)
dfc (x)
x
= 4 1 2
dx
c
(6-73)
so that:
x
= 4 1 2 .
2 ( a b)V
c
(6-74)
(6-75)
Substitute for from (6-75) into (6-74), and satisfy the boundary condition at x = b:
Vc ( + 2 )
b
= 4 1 2
c
2 ( b a )V
or:
1 c
b
( + 2 ) = 4 1 2 .
c
2 b a
(6-76)
2 b a
b
c
= 4 1 2
b a
(6-77)
and we solve for a and b. The first relation can be solved for (b-a):
c
(6-78)
(b a) =
2
and we obtain the same results obtained above, validating our previous analysis ( r = c/2).
3/11/98
b
c = 41 2 (b a)
c
or:
b c
c = 41 2
c 2
(6-79)
b
1 = 21 2
c
1
b
=1 2
c
2
b
1 3
=1 + =
c
2 2
b 13 3
=
=
c 22 4,
(6-80)
and use this to solve for a/c starting with Eqn. (6-78) :
(b a) =
c
2
or:
b a 1
=
c c 2
(6-81)
and:
a b 1 3 1
= = .
c c 2 4 2
Finally:
a 1
= .
c 4
(6-82)
Thus the vortex is located at the 1/4 chord point, and the control point is located at the 3/4
chord point. Naturally, this is known as the 1/4 - 3/4 rule. Its not a theoretical law, simply a
placement that works well and has become a rule of thumb. It was discovered by Italian Pistolesi.
Mathematical derivations of more precise vortex/control point locations are available (see Lan6 ),
but the 1/4 - 3/4 rule is widely used, and has proven to be sufficiently accurate in practice.
3/11/98
0.90
0.80
0.70
Cp
0.60
0.50
0.40
5% Circular arc camber line
=0
40 panels
0.30
0.20
0.10
0.0
0.2
0.4
x/c
0.6
0.8
1.0
Figure 6-14. Comparison in 2D of the 1/4-3/4 rule for vortex-control point locations with
linear theory, and including a comparison between placing the vortex and
control point on the camber line or on the axis.
*
A relatively large camber for a practical airfoil, the NACA 4412 example we used in Chapter 4 was an extreme
case, and it has 4% camber.
3/11/98
X
X
X
X
X
x
Trailing vortices extend to infinity
Figure 6-15. The horseshoe vortex layout for the classical vortex lattice method.
Note that the lift is on the bound vortices. To understand why, consider the vector statement
of the Kutta-Joukowski Theorem, F = V . Assuming the freestream velocity is the primary
contributor to the velocity, the trailing vortices are parallel to the velocity vector and hence the
force on the trailing vortices are zero. More accurate methods find the wake deformation
required to eliminate the force in the presence of the complete induced flowfield.
3/11/98
(6-60)
This is the velocity induced at the point m due to the nth horseshoe vortex, where Cm,n is a
vector, and the components are given by Equations 6-54, 6-58 and 6-59.
The total induced velocity at m due to the 2N vortices (N on each side of the planform) is:
Vmind = umind i + vmind j+ wmind k =
2N
Cm,nn .
(6-83)
n=1
The solution requires the satisfaction of the boundary conditions for the total velocity, which
is the sum of the induced and freestream velocity. The freestream velocity is (introducing the
possibility of considering vehicles at combined angle of attack and sideslip):
V = V cos cosi V sin j + V sin cos k
(6-84)
(6-1)
F(x, y,z) = 0.
(6-86)
F
= V F = 0 .
F
(6-87)
This equation provides freedom to express the surface in a number of forms. The most general
form is obtained by substituting Eqn. (6-85) into Eqn. (6-87) using Eqn. (6-83). This can be
written as:
ind
or:
3/11/98
2N
2N
2N
n=1
n=1
n=1
F
F
F
x i + y j + z k = 0
(6-89)
Carrying out the dot product operation and collecting terms:
2N
2N
F
F
(V cos cos + Cm,ni n )+
V sin + Cm,nj n +
x
y
n=1
n=1
2N
F
(V sin cos + Cm,nk n ) = 0 .
z
n=1
(6-90)
Recall that Eqn. (6-90) is applied to the boundary condition at point m. Next, we collect terms to
clearly identify the expression for the circulation. The resulting expression defines a system of
equations for all the panels, and is the system of linear algebraic equations that is used to solve
for the unknown values of the circulation distribution. The result is:
2N
x Cm,ni
n=1
F
F
F
F
F
Cm,nj +
Cm,nk n = V cos cos
sin
+ sin cos
y
z
x
y
z
m = 1,...2N (6-91)
This is the general equation used to solve for the values of the circulation. It is arbitrary,
containing effects of both angle of attack and sideslip (if the vehicle is at sideslip the trailing
vortex system should by yawed to align it with the freestream).
If the surface is primarily in the x-y plane and the sideslip is zero, we can write a simpler
form. In this case the natural description of the surface is
z = f(x, y)
(6-92)
(6-93)
F
f F
f F
= ,
= ,
= 1.
x
x y
y z
(6-94)
and
The gradient of F becomes
Substituting into the statement of the boundary condition, Eqn. (6-91), we obtain:
2N
n=1
This equation provides the solution for the vortex lattice problem.
3/11/98
m = 1,...,2N .
(6-95)
Note that if an essentially vertical surface is of interest, the form of F is more naturally
F = y - g(x,z),
and this should be used to work out the boundary condition in a similar fashion.
To illustrate the usual method, consider the simple planar surface case, where there is no
dihedral. Furthermore, recalling the example in the last section and the analysis at the beginning
of the chapter, the thin airfoil theory boundary conditions can be applied on the mean surface,
and not the actual camber surface. We also use the small angle approximations. Under these
circumstances, Eq. (6-95) becomes:
wm =
2N
Cm,nk n = V dxc
df
n=1
(6-96)
.
m
Thus we have the following equation which satisfies the boundary conditions and can be used to
relate the circulation distribution and the wing camber and angle of attack:
2N
Cm,nk V n = dxc m
n=1
df
m = 1,...,2N .
(6-97)
or
1. The Analysis Problem. Given camber slopes and , solve for the circulation strengths,
(/V ) [ a system of 2N simultaneous linear equations].
2. The Design Problem. Given (/V ), which corresponds to a specified surface
loading, we want to find the camber and required to generate this loading (only
requires simple algebra, no system of equations must be solved).
Notice that the way dfc/dx and are combined illustrates that the division between camber,
angle of attack and wing twist is arbitrary (twist can be considered a separate part of the camber
distribution, and is useful for wing design). However, care must be taken in keeping the
bookkeeping straight.
One reduction in the size of the problem is available in many cases. If the geometry is
symmetrical and the camber and twist are also symmetrical, then n is the same on each side of
the planform (but not the influence coefficient). Therefore, we only need to solve for N s, not
2N (this is true also if ground effects are desired, see Katz and Plotkin11). The system of
equations for this case becomes:
N
n=1
3/11/98
df
m = 1,..., N .
(6-98)
3/11/98
-0.50
Warren 12 planform
0.00
0.50
projection of bound
vortices on this row
intersect control point
on other side
of planform
horseshoe vortex
1.00
control point
1.50
2.00
2.50
trailing legs extend to infinity
3.00
-2.0
-1.5
-1.0
-0.5
0.0
0.5
1.0
1.5
2.0
Figure 6-16. Example of case requiring special treatment, the intersection of the projection of
a vortex with a control point.
A model problem illustrating this can be constructed for a simple finite length vortex
segment. The velocity induced by this vortex is shown in Fig. 6-17. When the vortex is
approached directly, x/l = 0.5, the velocity is singular for h = 0. However, as soon as you
approach the axis (h = 0) off the end of the segment (x/l > 1.0) the induced velocity is zero. This
illustrates why you can set the induced velocity to zero when this happens.
This second case that needs to be discussed arises when two or more planforms are used with
this method. This is one of the most powerful applications of the vortex lattice method. However,
care must be taken to make sure that the trailing vortices from the first surface do not intersect
the control points on the second surface. In this case the induced velocity is in fact infinite, and
the method breaks down. Usually this problem is solved by using the same spanwise distribution
of horseshoe vortices on each surface. This aligns the vortex the legs, and the control points are
well removed from the trailing vortices of the forward surfaces.
3/11/98
1.00
l
0.80
x
h
x=0
p
x/l
0.5
0.60
Vp
0.40
1.5
0.20
2.0
3.0
0.00
0.0
0.5
1.0
h/l
1.5
3/11/98
2.0
Model Tips
start simple
crude representation
of the planform is all right
keep centroids of areas
the same: actual planform to
vlm model
Make edges
streamwise
second planform
x
Figure 6-18. Example of a vlm model of an aircraft configuration. Note that one side of a
symmetrical planform is shown.
3/11/98
Three-Surface F-15
F/A-18
Figure 6-19.
3/11/98
Vortex Lattice
233 panels
Woodward
208 panels
Pan Air
162 panels
Figure 6-20. Panel models used for the three-surface F-15 (Ref.15).
Figure 6-21. Canard and horizontal tail height representation (Ref. 15).
3/11/98
Cm
(1/deg)
CL
(1/deg)
Wind Tunnel
15.70
.00623
.0670
Vortex Lattice
15.42
.00638
.0666
Woodward
14.18
.00722
.0667
Pan Air
15.50
.00627
.0660
Wind Tunnel
17.70
.00584
.0800
Vortex Lattice
16.76
.00618
.0750
Pan Air
15.30
.00684
.0705
Wind Tunnel
40.80
-.01040
.0660
Woodward
48.39
-.01636
.0700
Data Source
M = 0.2
M = 0.8
M = 1.6
3/11/98
16
Neutral
Point 12
percent c
Vortex lattice
Wind tunnel data
8
0.07
CL
per deg
0.06
-20
-10
10
20
3/11/98
-0.02
Cm
per deg.
-0.01
0.00
0.02
CL
H
per deg.
Wind tunnel
Vortex lattice
Woodward
0.01
0.00
0.0
0.4
0.8
1.2
Mach Number
1.6
2.0
0.0012
0.0008
Cl
a
per deg.
0.0004
0.0000
0.0
(a = 20)
Wing
Vortex lattice: Sym chordwise
panels
6
8
10
Wind tunnel
0.2
0.4
0.6
Mach Number
0.8
3/11/98
1.0
Figure 6-25. F/A-18 panel scheme with wing-tip missile and launcher (Ref. 15).
3/11/98
Table 6-2
F/A-18 Increments Due To Adding Wing-Tip Missiles and
Launchers,
and Vertical Tails
Mach
Number
Wing Tip
Missiles
and
Launchers
Vertical
Tails
Neutral
Point
(% mac)
CM
(1/deg)
CL
(1/deg)
Wind
Tunnel
0.2
1.10
-0.00077
0.0020
"
0.8
1.50
-0.00141
0.0030
"
1.6
-1.60
0.00148
0.0030
Vortex
Lattice
0.2
1.48
-0.00121
0.0056
"
0.8
2.11
-0.00198
0.0082
Woodward
0.2
1.52
-0.00132
0.0053
"
0.8
1.77
-0.00180
0.0079
"
1.6
-0.17
0.00074
0.0022
Wind
Tunnel
0.2
1.50
-0.00110
0.0050
"
0.8
2.00
-0.00202
0.0080
Vortex
Lattice
0.2
1.11
-0.00080
0.0022
"
0.8
1.32
-0.00108
0.0026
Finally, the effects of the number of panels and the way they are distributed is presented in
Figure 6-26. In this case the vlm method is seen to take between 130 to 220 panels to produce
converged results. For the vortex lattice method it appears important to use a large number of
spanwise rows, and a relatively small number of chordwise panels (5 or 6 appear to be enough).
3/11/98
Neutral
point, 2
percent c
(9)
(5)
Vortex Lattice Method
0
0
40
80
120
160
Total number of panels
200
240
(4)
4
Neutral
point, 2
percent c
(8)
Woodward Program
0
0
40
80
120
160
Total number of panels
200
240
(3)
Pan Air
0
80
120
160
200
Total number of panels
c) Pan Air program sensitivity to number of panels
Figure 6-26. F/A-18 panel convergence study (Ref. 15).
240
Although this study has been presented last in this section, a study like this should be
conducted before making a large number of configuration parametric studies. Depending on the
relative span to length ratio the paneling requirements may vary. The study showed that from
about 120 to 240 panels are required to obtain converged results. The vortex lattice methods
obtains the best results when many spanwise stations are used, together with a relatively small
number of chordwise panels. In that case about 140 panels provided converged results.
3/11/98
Figure 6-27. Highly swept lifting body type hypersonic concept (Ref. 16).
3/11/98
Figure 6-28. Comparison of CL , Cm, and CD predictions with data (Ref. 16).
Non-planar results from Kalman, Rodden and Giesing,17
All of the examples presented above considered essentially planar lifting surface cases. The
vortex lattice method can also be used for highly non-planar analysis, and the example cases
used at Douglas Aircraft Company in a classic paper 17 have been selected to illustrate the
capability. To avoid copyright problems, several of the cases were re-computed using the
Virginia Tech code JKayVLM, and provide an interesting comparison with the original results
from Douglas. Figure 6-29 presents an example of the prediction capability for the pressure
loading on a wing. In this case the geometry is complicated by the presence of a wing fence. The
pressures are compared with data on the inboard and outboard sides of the fence. The agreement
is very good on the inboard side. The comparison is not so good on the outboard side of the
fence. This quality of agreement is representative of the agreement that should be expected using
vortex lattice methods at low Mach numbers in cases where the flow would be expected to be
attached.
Figure 6-30 provides an example of the results obtained for an extreme non-planar case: the
ring, or annular, wing. In this case the estimates are compared with other theories, and seen to be
very good. The figure also includes the estimate of Cmq. Although not included in the present
discussion, Cmq and Clp can be computed using vlm methods, and this capability is included in
the vortex lattice method provided here, VLMpc.
3/11/98
3/11/98
8.0
7.0
c/4
6.0
CL
5.0
4.0
AR
4
3.0
2.0
1.0 0.0
1
0.2
0.4
0.6
0.8
1.0
1.2
1.4
h/c
0.20
0.10
CM
AR
1
2
4
0.00
-0.10
-0.20
-0.30
h/c
1.4
Figure 6-31. Example of ground effects for a simple rectangular wing (a case from Ref. 17).
The wings also experience a significant change in the pitching moment slope
(aerodynamic center shift), and this is also shown. Note that the predictions start to differ as the
ground is approached. JKayVLM actually rotates the entire surface to obtain another solution to
use in estimating the lift curve slope. The standard procedure used by most methods is to simply
change the slope condition at the mean line, as discussed previously in this chapter. Because of
3/11/98
10
CL
h/c = 0.5
6
h/c = 1
4
h/c =
2
Note: Most of the decrease in lift curve slope for the wing out of ground
effect is due to the effect of reduced span as the wing rotates from
the zero dihedral condition
0
-40
-30
-20
-10
0
10
Dihedral angle,
20
30
Figure 6-32. Example of ground effects for a wing with dihedral (a case from Ref. 17).
3/11/98
40
=2 2
LE = 53.54
Swing = 2 2
AR
CL
1.50
1.91421
= 2.743 / rad
CM = 3.10 / rad
0.50
1.41421
Warren-12 Planform
Figure 6-33. Definition and reference results for the Warren-12 wing.
For the results cited above, the reference chord used in the moment calculation is the average
chord (slightly nonstandard, normally the reference chord used is the mean aerodynamic chord)
and the moment reference point is located at the wing apex (which is also nonstandard).
3/11/98
For a trapezoidal wing with an elliptic spanload the maximum value of the local lift coefficient occurs at = 1 - .
3/11/98
drag. For a specified span, the performance of the wing is evaluated by finding the value of the
span efficiency factor, e, as described in Chapter 5. Finally, the wing root bending moment
provides an indication of the structural loading requirements that the wing structure must be
designed to accommodate. When considering the total system, the basic aerodynamic efficiency
may be compromised to reduce structural wing weight. The shape of the spanload can be
controlled through a combination of planform selection and wing twist. Typical twist
distributions required to produce good wing characteristics are presented below.
The simplest example of planform shaping is the selection of wing aspect ratio, AR, wing
taper, , and wing sweep, . While the aerodynamicist would like to see high values the aspect
ratio, several considerations limit aspect ratio. Perhaps the most important limitation is the
increase of wing structural weight with increasing aspect ratio. In addition, the lift coefficient
required to maximize the benefits of high aspect ratio wings increases with the square root of the
aspect ratio. Hence, airfoil performance limits can restrict the usefulness of high aspects ratios,
especially for highly swept wings based on airfoil concepts. In recent years advances in both
aerodynamics and structures have allowed aircraft to be designed with higher aspect ratios and
reduced sweep. Table 6-3 provides some key characteristics of transport wings designed to
emphasize efficient cruise while meeting takeoff and landing requirements.
Taper: Several considerations are used in selecting the wing taper. For a straight untwisted,
unswept wing, the minimum induced drag corresponds to a taper ratio of about 0.4. However, a
tapered wing is more difficult and hence expensive to build than an untapered wing. Many
general aviation aircraft wings are built with no taper (all ribs are the same, reducing fabrication
cost, and the maximum section Cl occurs at the root, well away from the control surface). To
reduce structural weight the wing should be highly tapered, with < .4. However, although
highly tapered wings are desirable structurally, the section lift coefficient near the tip may
become high. This consideration limits the amount of taper employed (current jet transports use
taper ratios in the range of 0.2 to 0.3, as well as progressively increasing twist upward from the
tip). As an example, the Aero Commander 500 had an aspect ratio of 9.5 and a taper ratio of 0.5
(it also had -6.5 of twist and the quarter chord of the wing was swept forward 4).
Sweep: Sweep is used primarily to delay the effects of compressibility and increase the drag
divergence Mach number. The Mach number controlling these effects is approximately equal to
the Mach number normal to the leading edge of the wing, Meff = M cos. The treatise on swept
planforms by Kchemann is very helpful in understanding swept wing aerodynamics.18
Aerodynamic performance is based on the wingspan, b. For a fixed span, the structural span
increases with sweep, bs = b/cos, resulting in a higher wing weight. Wing sweep also leads to
aeroelastic problems. For aft swept wings flutter becomes an important consideration. If the wing
3/11/98
is swept forward, divergence is a problem. Small changes in sweep can be used to control the
aerodynamic center when it is not practical to adjust the wing position on the fuselage (the DC-3
is the most famous example of this approach).
Table 6-3
Typical Planform Characteristics of Major Transport Aircraft
1st Flight
Aircraft
W/S
AR
(c/4 )
1957
B707-120
105.6
7.04
35
0.293
1958
DC-8-10
111.9
7.32
30
0.230
1963
B707-320C
110.0
7.06
35
0.250
1970
B747-200B
149.1
6.96
37.5
0.240
1970
L-1011
124.4
8.16
35
0.200
1972
DC-10-30
153.7
7.57
35
0.230
1972
A300 B2
107.9
7.78
28
0.230
1982
A310-100
132.8
8.80
28
0.260
1986
B767-300
115.1
7.99
31.5
0.182
1988
B747-400
149.9
7.61
37.5
0.240
1990
MD-11
166.9
7.57
35
0.230
1992
A330
119.0
9.3
29.74
0.192
To understand the effects of sweep, the Warren 12 wing is compared with wings of the same
span and aspect ratio, but unswept and swept forward. The planforms are shown in Figure 6-34.
The wing leading edge sweep of the aft swept wing becomes the trailing edge sweep of the
forward swept wing. Figure 6-35 provides the spanload and section lift coefficient distributions
from VLMpc. The spanload, ccl / ca , is given in Fig. 6-35a, where, c is the local chord, cl is the
local lift coefficient, based on the local chord, and ca is the average chord, S/b. Using this
nomenclature, the area under the curve is the total wing lift coefficient. Note that sweeping the
wing aft increases the spanload outboard, while sweeping the wing forward reduces the spanload
outboard. This follows directly from a consideration of the vortex lattice model of the wing. In
both cases, the portion of the wing aft on the planform is operating in the induced upwash
flowfield of the wing ahead of it, resulting in an increased spanload. Figure 6-35b shows the
corresponding value of the local lift coefficient. Here the effect of sweep is more apparent. The
forward swept wing naturally results in a spanload with a nearly constant lift coefficient. This
means that a comparatively higher wing lift coefficient can be achieved before the wing stall
begins. The program LIDRAG can be used to compare the span es associated with these
spanloads (an exercise for the reader).
3/11/98
V
Forward
Swept
Wing
Unswept
Wing
Aft
Swept
Wing
Figure 6-34. Comparison of forward, unswept, and aft swept wing planforms, AR = 2.8.
3/11/98
1.60
1.40
1.20
Spanload,
1.00
c cl / ca
0.80
0.60
unswept wing
0.40
0.20
0.00
0.2
0.4
y/(b/2)
0.6
0.8
1.0
0.8
1.0
a) comparison of spanloads
1.40
aft swept wing
1.20
1.00
Section CL
0.80
unswept wing
forward swept wing
0.60
0.40
0.20
Warren 12 planform, sweep changed
0.00
0.0
0.2
0.4
0.6
y/(b/2)
3/11/98
V
Forward
Swept
Wing
Unswept
Wing
Aft
Swept
Wing
Figure 6-36. Comparison of forward, unswept, and aft swept wing planforms, AR = 8.
3/11/98
1.6
Aspect ratio 8 wings
1.4
1.2
Spanload,
cc / c 1.0
l
0.8
aft swept wing
0.6
unswept wing
0.4
0.2
0.0
0.2
0.4
0.6
0.8
1.0
0.8
1.0
y/(b/2)
a) comparison of spanloads
1.50
aft swept wing
1.00
unswept wing
Section CL
0.50
0.2
0.4
0.6
y/(b/2)
3/11/98
0.08
data from NACA RM A50K27
0.06
Cm
0.04
x ref = c/4
0.02
0.00
-0.02
-0.04
VLMpc calculation
-0.06
-0.4
-0.2
0.0
0.2
0.4
CL
0.6
0.8
1.0
1.2
Figure 6-38. Example of isolated wing pitchup: NACA data19 compared with VLMpc.
To control the spanload, the wing can be twisted. Figure 6-39 shows typical twist
distributions for aft and forward swept wings, obtained from John Lamars program LamDes.20
(see Chapter 5 for a description). In each case the twist is used to reduce the highly loaded areas,
and increase the loading on the lightly loaded portions of the wing. For an aft swept wing this
means the incidence is increased at the wing root, known as washin, and reduced, known as
washout, at the wing tip. Just the reverse is true for the forward swept wing. The sudden drop in
required twist at the tip for the forward dwept wing case is frequently seen in typical design
methods and attribute to a weakness in the method and faired out when the aerodynamicist
gives his design to the lofting group.
Although geometric sweep is used to reduce the effective Mach number of the airfoil, the
geometric sweep is not completely effective. The flowfield resists the sweep. In particular, the
wing root and tip regions tend to effectively unsweep the wing. Aerodynamicists study lines of
constant pressure on the wing planform known as isobars to investigate this phenomenon. Figure
6-40 presents the computed isobars for a typical swept wing,21 using a transonic small
disturbance method.22 The effect is dramatic. The effective sweep may actually correspond to
the isobar line from the wing root trailing edge to the leading edge at the wing tip. To increase
the isobar sweep, in addition to geometric sweep and twist, the camber surface and thickness are
typically adjusted to move the isobars forward at the wing root and aft at the wing tip. This is a
key part of the aerodynamic wing design job, regardless of the computational, methodology used
to obtain the predicted isobar pattern.
3/11/98
0.2
0.4
0.6
y/(b/2)
0.8
7.0
6.0
5.0
, 4.0
deg. 3.0
2.0
1.0
0.0
-1.0
0
0.2
0.4
0.6
y/(b/2)
0.8
AR = 4
= 0.6
45
3/11/98
= 0.925
= 0.475
0.05
0.04
0.03
(z-zle)/c
0.02
= 0.075
0.01
0.00
-0.01
-0.2
0.2
0.4
x/c
0.6
0.8
1.2
Figure 6-41. Comparison of camber lines required to develop the same chord load shape at
the root, mid-span and tip region of an aft swept wing. (from LamDes 20)
Many other refinements are available to the aerodynamic designer. Insight into both the
human and technical aspects of wing design prior to the introduction of computational
aerodynamics is available in two recent books describing the evolution of the Boeing series of jet
transports.23,24 One interesting refinement of swept wings has been the addition of trailing edge
area at the wing root. Generally known as a Yehudi flap, this additional area arises for at least
two reasons. The reason cited most frequently is the need to provide structure to attach the
landing gear at the proper location. However, the additional chord lowers the section lift
coefficient at the root, where wing-fuselage interference can be a problem, and the lower
required section lift makes the design job easier. Douglas introduced this planform modification
for swept wings on the DC-8, while Boeing did not incorporate it until the -320 model of the
707. However, the retired Boeing engineer William Cook, in his book, 23 on page 83, says it was
first introduced on the B-29 to solve an interference problem between the inboard nacelle and the
fuselage. The aerodynamic benefit to the B-29 can be found in the paper by Snyder. 25 Cook says,
in a letter to me, that the device got its name because each wind tunnel part needed a name and
there was a popular radio show at the time that featured the continuing punch line Whos
Yehudi? (the Bob Hope Radio show featuring Jerry Colonna, who had the line). Thus, a Boeing
engineer decided to call it a Yehudi flap. This slight extra chord is readily apparent when
examining the B-29, but is very difficult to photograph.
3/11/98
Cn = Cs cos
Mn = Mcos
t/c)n = t/c)s/cos
Cn
Cs
(699)
and
CLs = CLn cos2
3/11/98
3/11/98
3/11/98
-0.2
h
-0.4
Nose down
0
+25
-25
-0.6
0
20
40
, deg
60
80
Figure 6-44. Pitching moment characteristics of an F-16 type wind tunnel model.28
Upwash outboard
of canard tips
Canard wake
streams over
wing
+
w 0
A
y
Downwash from canard across wing
at Section A-A
3/11/98
Cp
Cp
0
x/c 1
x/c
0.0350
M = 0.3, CL = 0.5
Canard lift must be
negative to trim
Canard height
above wing ,z/b
CDi
0.1
0.2
0.0300
Static Margin
0.3
Stable Unstable
0.0250
0.0200
-0.4 -0.3
Forward
Advantage of relaxed
static stability evident
-0.2
-0.1
0.0
0.1
x/c
Advantage of vertical
separation clearly evident
0.2
0.3
0.4
Aft
b) Minimum trimmed drag variation with trim position and canard-wing separation.
Figure 6-46. Example of relation of minimum trimmed drag to balance (stability).20
3/11/98
8.0
in presence
of canard
6.0
canard tip
vortex effect
in presence
of canard
4.0
2.0
without
canard
0.0
canard tip
vortex effect
-2.0
0.0 0.2 0.4 0.6 0.8 1.0
0.4
0.6
0.8
1.0
y/(b/2)
y/(b/2)
a) aft swept wing
b) forward swept wing
Figure 6-47. Effects of canard on twist requirements. Twist required for minimum drag
using LamDes20 (Note: results depend on configuration details, balance).
1
(x ) dx .
w(x) =
2 x x
(6-100)
For thin wing theory the vertical velocity can be related to the slope as shown above in Section
6.2. The vortex strength can be related to the streamwise velocity by = u+ - u-. This in turn can
be used to relate the vorticity to the change in pressure, Cp through:
Cp = Cpl C pu = 2u 2u+ = 2 u+ u
(6-101)
= (x)
2
resulting in the expression for camber line slope in terms of design chord load:
(6-102)
dz
1 Cp
=
dx .
dx
4 x x
0
3/11/98
(6-103)
x k (1 xk )
xi x k
(6-104)
where:
xk =
1
(2k 1)
1 cos
2N
2
k = 1,2,..., N
(6-105)
and:
xi =
1
i
1 cos
N
2
i = 0,1,2,..., N .
(6-106)
Here N + 1 is the number of stations on the camber line at which the slopes are obtained.
Given dz/dx, the camber line is then computed by integration using the trapezoidal rule
(marching forward starting at the trailing edge):
dz
x x dz
zi +1 = z i i+1 i
+
.
dx i dx i+1
2
(6-107)
(6-108)
The camber line can then be redefined in standard nomenclature, i.e., z(x=0) = z(x=1) = 0.0:
zi = zi (1 x i )tan DES
(6-109)
How well does this work? Program DesCam implements the method described here, and the
users manual is provided in App. D.7. Here we compare the results from DesCam with the
analytic formula given in Appendix A.1 for the NACA 6 Series mean line with a = .4. The
results are shown in Figure 6-48 below. Notice that the camber scale is greatly enlarged to
3/11/98
(Z-Z0)/C - DesCam
Z/C - from Abbott & vonDoenhoff
0.12
2.00
Design Chord
Loading
0.10
1.50
CP
0.08
Z/C
1.00
0.06
0.50
0.04
0.00
0.02
0.00
0.0
0.2
0.4
0.6
0.8
-0.50
1.0
X/C
Figure 6-48. Example and verification of camber design using DesCam.
3/11/98
6.12 Vortex Flow Effects and the Leading Edge Suction Analogy
For highly swept wings at even moderate angles of attack, the classical attached flow/trailing
edge Kutta condition flow model weve adopted is wrong. Instead of the flow remaining attached
on the wing and leaving the trailing edge smoothly, the flow separates at the leading edge,
forming a well defined vortex. This vortex plays an important role in the design of highly swept,
or slender wing aircraft. The most notable example of this type of configuration is the
Concorde. Sharp leading edges promote this flow phenomena. The basic idea is illustrated in the
sketch from Payne and Nelson29 given here in Fig. 6-49.
Vortical flow
abova a delta wing
Primary vortex
Secondary vortex
Figure 6-49. Vortex flow development over a delta wing with sharp edges.29
An important consequence of this phenomena is the change in the characteristics of the
lift generation as the wing angle of attack increases. The vortex that forms above the wing
provides an additional low pressure force due to the strongly spiraling vortex flow. The low
pressure associated with the centrifugal force due to the vortex leads to the lower pressure on the
wing surface. As the wing increases its angle of attack the vortex gets stronger, further reducing
the pressure on the wing. The resulting increase in lift due to the vortex can be large, as shown in
Fig. 6-50, from Polhamus.30
This is an important flow feature. Slender wings have very low attached flow lift curve
slopes, and without the additional vortex lift it would be impractical to build configurations with
low aspect ratio wings. The low attached flow alone lift curve slope would prevent them from
being able to land at acceptable speeds or angle of attack. Vortex lift made the Concorde
possible. Another feature of the flow is the high angle of attack at which maximum lift occurs,
3/11/98
Subsonic lift
L.E. = 75
1.2
Vortex
flow
CL
Vortex lift
0.6
Attached flow
0.0
20
, deg
40
Figure 6-50. Vortex lift changes the characteristic lift development on wings. 30
Although the vortex lattice method formulation presented above does not include this effect,
vortex lattice methods are often used as the basis for extensions that do include the leading edge
vortex effects. A remarkable, reasonably accurate, flow model for leading edge vortex flows was
introduced by Polhamus31,32 at NASA Langley in 1966 after examining lots of data. This flow
model is known as the Leading Edge Suction Analogy. The concept is quite simple and was
invented for sharp edged wings. The leading edge suction that should exist according to attached
flow theory (see section 5.8) is assumed to rotate 90 and generate a vortex induced force instead
of a suction when leading edge vortex flow exists. Thus the vortex flow force is assumed to be
equal to the leading edge suction force. However, the force now acts in the direction normal to
the wing surface in the direction of lift rather than in the plane of the wing leading edge. The
concept is shown in the Fig. 6-51 from the original Polhamus NASA report.31 Further details on
the effects of vortex flow effects are also available in the reports by Kulfan.33,34
3/11/98
CS
CS/2
CS
Separated flow
(sharp edge)
ttachment
point
CSS
Spiral
vortex
Section A-A
Figure 6-51. The Polhamus leading edge suction analogy.
Polhamus developed charts to compute the suction force for simple wing shapes. For a
delta wing with a sharp leading edge, the method is shown compared with the data of Bartlett
and Vidal35 in Figure 6-52. The agreement is quite good (my reconstruction doesnt show
agreement as good as that presented by Polhamus,31 but it is still impressive).
The figure also shows the large size of the vortex lift, and the nonlinear shape of the lift
curve when large angles are considered. This characteristic was exploited in the design of the
Concorde.
To find the vortex lift using the leading edge suction analogy, an estimate of the leading edge
suction distribution is required. However the suction analogy does not result in an actual
flowfield analysis including leading edge vortices. The Lamar vortex lattice code (VLMpc)
optionally includes a fully developed suction analogy based on Polhamus ideas, with extensions
to treat side edge suction by John Lamar3 also included.
Other approaches have been developed to compute leading edge vortex flows in more detail.
Many of these methods allow vortex filaments, simulating the leading edge vortices, to leave the
leading edge. The location of these vortices, and their effect on the wing aerodynamics as they
roll up are explicitly computed. Mook10 and co-workers are leaders in this methodology.
The area of vortex flows in configuration aerodynamics is fascinating, and an entire
conference was held at NASA Langley36 devoted to the topic. The references cited above were
selected to provide an entry to the literature of these flows. Interest in the area remains strong.
The effects of round leading edges have been investigated by Ericsson and Reding37 and
Kulfan.33 The relation between sweep, vortex lift, and vortex strength has been given recently by
Hemsch and Luckring. 38
3/11/98
2.00
AR = 1.5 ( = 69.4)
Prediction from Polhamus
Leading Edge Suction Analogy
1.50
CL
Experimental Data of
Bartlett and Vidal (ref. 35)
Vortex
Lift
1.00
0.500
Potential
Lift
0.00
0.0
10.0
20.0
30.0
40.0
Figure 6-52. Comparison of the leading edge suction analogy with data.
6.13 Alternate and Advanced VLM Methods
Many variations of the vortex lattice method have been proposed. They address both the
improvement in accuracy for the traditional case with a planar wake, and extensions to include
wake position and rollup as part of the solution. Areas requiring improvement include the ability
to predict leading edge suction, the explicit treatment of the Kutta condition, and the
improvement in convergence properties with increasing numbers of panels.8 The traditional
vortex lattice approaches assume that the wing wake remains flat and aligned with the
freestream. This assumption is acceptable for most cases. The effect of the wake on the wing that
generates it is small unless the wing is highly loaded. However, the interaction between the wake
from an upstream surface and a trailing lifting surface can be influenced by the rollup and
position.
In the basic case where the wake is assumed to be flat and at a specified location, the primary
extensions of the method have been directed toward improving the accuracy using a smaller
number of panels. Hough7 demonstrated that improvement in accuracy could be achieved by
using a lattice that was slightly smaller than the true planform area. Basically, he proposed a 1/4
panel width inset from the tips.
3/11/98
3/11/98
(1 x/c )
x /c
i. Pick an aspect ratio 10 unswept wing at = 3 and 12 and run VLMpc.
ii. Plot (Cp)/ as a function of x/c at the wing root.
iii. How many panels do you need to get a converged solution from VLM?
iv. What conclusions do you reach?
3. Compare the validity of an aerodynamic strip theory using VLMpc. Consider an uncambered,
untwisted wing, AR = 4, = .4, le = 50, at a lift coefficient of 1. Plot the spanload, and the
Cp distribution at approximately the center section, the midspan station, and the 85%
semispan station. Compare your results with a spanload constructed assuming that the wing
flow is approximated as 2D at the angle of attack required to obtain the specified lift. Also
compare the chordloads, Cp, at the three span stations. How many panels do you need to
obtain converged results. Document your results. Do you consider this aerodynamic strip
theory valid based on this investigation? Comment.
4. Compare the wing aerodynamic center location relative to the quarter chord of the mac for the
wing in exercise 3, as well as similar wings. Consider one wing with zero sweep on the
quarter chord, and a forward swept wing with a leading edge sweep of -50. Compare the
spanloads. Document and analyze these results. What did you learn from this comparison?
5. For the wings in exercise 4, compare the section lift coefficients. Where would each one stall
first? Which wing appears to able to reach the highest lift coefficient before the section stalls.
6. For the problem in exercise 5, add twist to each wing to obtain near elliptic spanloads.
Compare the twist distributions required in each case.
7. Pick a NASA or NACA report describing wind tunnel results for a simple one or two lifting
surface configuration at subsonic speeds. Compare the lift curve slope and stability level
predicted by VLMpc with wind tunnel data. Submit a report describing your work and
assessing the results.
8. Add a canard to the aft and forward swept wings analyzed in exercise 4. Plot the sum of the
spanloads. How does the canard effect the wing spanload.
9. Consider the wings in exercise 8. How does lift change with canard deflection? Add an
equivalent tail. Compare the effect of tail or canard deflection on total lift and moment. Did
you learn anything? What?
10. Construct a design code using the 1/4 - 3/4 rule and compare with DESCAM.
11. Construct a little 2D code to study ground effects.
12. Compare wing and wing/tail(canard) results for CL with standard analytic formulas.
3/11/98
6.15 References
1
20
Lamar, J.E., A Vortex Lattice Method for the Mean Camber Shapes of Trimmed NonCoplanar Planforms with Minimum Vortex Drag, NASA TN D-8090, June 1976.
21
Loving, D.L., and Estabrooks, B.B., Transonic Wing Investigation in the Langley Eight Foot
High Speed Tunnel at High Subsonic Mach Numbers and at Mach Number of 1.2, NACA RM
L51F07, 1951.
22
Mason, W.H., MacKenzie, D.A., Stern, M.A., Ballhaus, W.F, Jr., and Frick, J., Automated
Procedure for Computing the Three-Dimensional Transonic Flow Over Wing-Body
Combinations, including Viscous Effects, Vol. I, Description of Methods and Applications,
AFFDL-TR-77-122, February 1978.
23
Cook, William C., The Road to the 707, TYC Publishing, Bellevue, 1991.
24
Irving, Clive, Wide-Body: The Triumph of the 747, William Morrow, New York, 1993.
25
Snyder, George, Structural Design Problems in the B-29 Airplane, Aeronautical Engineering
Review, Feb. 1946, pp. 9-12.
26
Hunton, Lynn W., A Study of the Application of Airfoil Section Data to the Estimation of the
High-Subsonic-Speed Characteristics of Swept Wings, NACA RM A55C23, June 1955.
27
Shevell, R.S., and Schaufele, R.D., Aerodynamic Design Features of the DC-9, Journal of
Aircraft, Vol. 3, No. 6, Nov.-Dec. 1966, pp. 515-523.
28
Nguyen, L.T., Ogburn, M.E., Gilbert, W.P., Kibler, K.S., Brown, P.W., and Deal, P.L.,
Simulator Study of Stall/Post-Stall Characteristics of a Fighter Airplane With Relaxed
Longitudinal Static Stability, NASA TP 1538, Dec. 1979.
29
Payne, F.W., and Nelson, R.C., An Experimental Investigation of Vortex Breakdown on a
Delta Wing, in Vortex Flow Aerodynamics, NASA CP 2416, 1985.
30
Polhamus, E.C., Application of Slender Wing Benefits to Military Aircraft, AIAA Wright
Brothers Lecture, AIAA-83-2566, 1983..
31
Polhamus, E.C., A Concept of the Vortex Lift on Sharp Edge Delta Wings Based on a
Leading-edge Suction Analogy, NASA-TN 3767, 1966.
32
Polhamus, E.C., Prediction of Vortex Lift Characteristics by a Leading-edge Suction
Analogy,, Journal of Aircraft, Vol. 8, No. 4, 1971, pp. 193-199.
33
Kulfan, R.M., Wing Airfoil Shape Effects on the Development of Leading-Edge Vortices,
AIAA 79-1675, 1979.
34
Kulfan, R.M., Wing Geometry Effects on Leading Edge Vortices, AIAA Paper No. 79-1872,
1979.
35
Bartlett, G.E., and Vidal, R.J., Experimental Investigation of Influence of Edge Shape on the
Aerodynamic Characteristics of Low Aspect Ratio Wings at Low Speeds, Journal of the
Aeronautical Sciences, Vol. 22, No. 8, August, 1955, pp.517-533,588.
36
Vortex Flow Aerodynamics, NASA SP-2416 (volume I), and NASA SP-2417 (volume II),
October, 1985.
37
Ericsson, L.E., and Reding, J.P., Nonlinear Slender Wing Aerodynamics, AIAA Paper No.
76-19, January, 1976.
38
Hemsch, M.J., and Luckring, J.M., Connection Between Leading-Edge Sweep, Vortex Lift,
and Vortex Strength for Delta Wings, Journal of Aircraft, Vol. 27, No. 5, May 1990, pp. 473475.
39
Albano, E., and Rodden, W.P., A Doublet-Lattice Method for Calculating Lift Distributions
on Oscillating Surfaces in Subsonic Flows, AIAA J., Vol. 7, No. 2, February 1969, pp. 279-285;
errata AIAA J., Vol. 7, No. 11, November 1969, p. 2192.
3/11/98
8. Introduction to
Computational Fluid Dynamics
We have been using the idea of distributions of singularities on surfaces to study the
aerodynamics of airfoils and wings. This approach was very powerful, and provided us with
methods which could be used easily on PCs to solve real problems. Considerable insight into
aerodynamics was obtained using these methods. However, the class of effects that could be
examined was somewhat restricted. In particular, practical methods for computing fundamentally
nonlinear flow effects were excluded. This includes both inviscid transonic and boundary layer
flows.
In this chapter we examine the basic ideas behind the direct numerical solution of differential
equations. This approach leads to methods that can handle nonlinear equations. The simplest
methods to understand are developed using numerical approximations to the derivative terms in
the partial differential equation (PDE) form of the governing equations. Direct numerical
solutions of the partial differential equations of fluid mechanics constitute the field of
computational fluid dynamics (CFD). Although the field is still developing, a number of books
have been written.1,2,3,4,5,6 In particular, the book by Tannehill et al,1 which appeared in 1997 as a
revision of the original 1984 text, covers most of the aspects of CFD theory used in current codes
and reviewed here in Chapter 14. Fundamental concepts for solving partial differential equations
in general using numerical methods are presented in a number of basic texts. Smith7 and Ames 8
are good references.
The basic idea is to model the derivatives by finite differences. When this approach is used
the entire flowfield must be discretized, with the field around the vehicle defined in terms of a
mesh of grid points. We need to find the flowfield values at every mesh (or grid) point by writing
down the discretized form of the governing equation at each mesh point. Discretizing the
equations leads to a system of simultaneous algebraic equations. A large number of mesh points
is usually required to accurately obtain the details of the flowfield, and this leads to a very large
system of equations. Especially in three dimensions, this generates demanding requirements for
computational resources. To obtain the solution over a complete three dimensional aerodynamic
configuration millions of grid points are required!
3/17/98
8 -1
3/17/98
Introduction to CFD 8 - 3
distance x away. Then we can approximate the slope at x0 by taking the slope between these
points. The sketch illustrates the difference between this simple slope approximation and the
actual slope at the point x0. Clearly, accurate slope estimation dependents on the method used to
estimate the slope and the use of suitably small values of x.
True slope
at X 0
Approximate slope
at X 0
X
x0
Figure 8-1. Example of slope approximation using two values of the function.
Approximations for derivatives can be derived systematically using Taylor series
expansions. The simplest approach is to find an estimate of the derivative from a single series.
Consider the following Taylor series:
df
f (x0 + x) = f(x 0 ) + x
dx
and rewrite it to solve for
+
x0
(x )2 d 2f
2
dx2 x
0
(x )3 d 3f
6
dx3 x
0
+ .. .
(8-1)
df
:
dx x 0
f ( x0 + x ) f ( x0 )
df
1 d 2f
=
x
...
dx x 0
x
2 dx2 x
0
or:
f ( x0 + x ) f ( x0 )
df
=
+ O
(x
12
3)
dx x 0
x
(8-2)
Truncation
Error
where the last term is neglected and called the truncation error. In this case it is O(x). The term
truncation error means that the error of the approximation vanishes as x goes to zero. * The
*
This assumes that the numerical results are exactly accurate. There is a lower limit to the size of the difference step
in x due to the use of finite length arithmetic. Below that step size, roundoff error becomes important. In most
3/17/98
+ . . ..
dx x 0
2 dx2 x
6 dx3 x
0
0
(8-3)
Solving for the first derivative in the same manner we used above, we obtain:
f ( x0 ) f ( x0 x)
df
=
+ O( x ) ,
dx x 0
x
(8-4)
cases the stepsize used for practical finite difference calculations is larger than the limit imposed by roundoff errors.
We cant afford to compute using grids so finely spaced that roundoff becomes a problem.
*
With x small, x2 is much smaller than x
3/17/98
Introduction to CFD 8 - 5
f ( x0 + x ) f ( x0 x )
df
=
+ O( x )2 .
dx x 0
2x
(8-5)
This is a second order accurate central difference formula since information comes from both
sides of x0. Numerous other approximations can be constructed using this approach. Its also
possible to write down second order accurate forward and backward difference approximations.
We also need the finite difference approximation to the second derivative. Adding the Taylor
series expressions for the forward and backward expansions, Eq.(8-1) and Eq.(8-3), results in the
following expression, where the odd order terms cancel:
f (x0 + x) + f(x 0 x) = 2 f (x0 ) + (x )2
d 2f
+ O (x ) 4
2
dx x
0
(8-7)
(8-8)
In addition, expressions can be derived for cases where the points are not evenly distributed.
In general the formal truncation error for unevenly spaced points is not as high as for the evenly
spaced point distribution. In practice, for reasonable variations in grid spacing, this may not be a
serious problem. We present the derivation of these expressions here. A better way of handling
non-uniform grid points is presented in the next chapter. The one sided first derivative
expressions Eq.(8-2) and Eq.(8-4) are already suitable for use in unevenly spaced situations. We
need to obtain a central difference formula for the first derivative, and an expression for the
3/17/98
( )
x +
d 2f
+
6
dx2 x
0
( )
x
d 2f
6
dx2 x
0
x
df
f (x0 x ) = f (x0 ) x
+
dx x0
2
( )
x +
+
+ df
f (x0 + x ) = f (x0 ) + x
+
dx x0
2
( )
d 3f
+ . ..
dx3 x
0
(8-9)
d 3f
+...
dx3 x
0
(8-10)
Define x+ = x . To obtain the forms suitable for derivation of the desired expressions,
replace x+ in Eq. (8-9) with x , and multiply Eq. (8-10) by . The resulting expressions
are:
x )
(
+
f (x0 + x ) = f (x0 ) + x
df
dx x 0
dx2 x
0
x )
(
+
x )
(
+
d 2f
2
( )
d 3f
+ . ..
dx3 x
0
(8-11)
x d 3f
d 2f
f (x 0 x ) = f (x0 ) x
+ . . . (8-12)
dx x 0
2
6
dx 2 x
dx3 x
0
0
To obtain the expression for the first derivative, subtract Eq ( 8-12) from Eq. (8-11).
df
df
dx x0
( )
2
2 2
x
x
d f
+
+...
2
2 dx2
x0
(8-13)
dx x 0
2x
(8-14)
To obtain the expression for the second derivative, add (8-11) and (8-12):
( )
2
2 2
x
x
d f
f (x0 + x + ) +f (x 0 x ) = f (x0 ) + f(x 0 ) +
+
+ O(x)3 . ..
2
2
2
dx x
0
3/17/98
Introduction to CFD 8 - 7
d 2f
f (x0 + x + ) (1 + ) f (x 0 ) +f(x0 x )
=
+ O(x )
2
2
dx x
(1+ ) x
0
2
( )
(8-15)
Note that both Eqs. (8-14) and (8-15) reduce to the forms given in Eq.(8-5) and Eq.(8-6)
when the grid spacing is uniform.
Finally, note that a slightly more sophisticated analysis (Tannehill, et al,1 pages 61-63) will
lead to a second order expression for the first derivative on unevenly spaced points:
dx x 0
( +1)x
(8-16)
Tannehill, et al,1 give additional details and a collection of difference approximations using
more than three points and difference approximations for mixed partial derivatives (Tables 3-1
and 3-2 on their pages 52 and 53). Numerous other methods of obtaining approximations for the
derivatives are possible. The most natural one is the use of a polynomial fit through the points.
Polynomials are frequently used to obtain derivative expressions on non-uniformly spaced grid
points.
These formulas can also be used to represent partial derivatives. To simplify the notation, we
introduce a grid and a notation common in finite difference formulations. Figure 8-2 illustrates
this notation using x = y = const for these examples.
Assume:
i,j+1
x = y = const.
x = ix
y = jx
y or "j"
i-1,j
i,j
i+1,j
i,j-1
x or "i"
Figure 8-2. Nomenclature for use in partial differential equation difference expressions.
3/17/98
(8 - 20)
Similar expressions can be written for the y derivatives. To shorten the expressions, various
researchers have introduced different shorthand notations to replace these expressions. The
shorthand notation is then used in further operations on the difference expressions.
8.2 Finite difference representation of Partial Differential Equations (PDE's)
We can use the approximations to the derivatives obtained above to replace the individual
terms in partial differential equations. The following figure provides a schematic of the steps
required, and some of the key terms used to ensure that the results obtained are in fact the
solution of the original partial differential equation. We will define each of these new terms
below.
Discretization
Consistency
?
Exact
Solution
System
of
Algebraic
Equations
Stability
Convergence
as x,t 0
Approximate
Solution
3/17/98
Introduction to CFD 8 - 9
Successful numerical methods for partial differential equations demand that the physical
features of the PDE be reflected in the numerical approach. The selection of a particular finite
difference approximation depends on the physics of the problem being studied. In large part the
type of the PDE is crucial, and thus a determination of the type, i.e. elliptic, hyperbolic, or
parabolic is extremely important. The mathematical type of the PDE must be used to construct
the numerical scheme for approximating partial derivatives. Some advanced methods obscure the
relationship, but it must exist. Consider the example given in Fig. 8-4 illustrating how
information in a grid must be used.
subsonic flow
supersonic flow
i,j+1
i,j+1
zone of
dependence
i-1,j
i,j
i,j-1
i+1,j
i-1,j
i,j
i+1,j
i,j-1
3/17/98
x
i+1
i
i-1
n-1
n
n+1
Figure 8-5. Grid nomenclature for discretization of heat equation.
The heat equation can now be written as:
u
2u
2 =
t424
x3
1
PDE
uin+1 uin
n
n
ui+1
2uin + ui1
+
2
t
(
x
)
1444 444 2444 444 3
FDE
2 n
4 n
2
u t + u (x ) + ... = 0
2
x 4 i 12
t i 2
(8-22)
Truncation Error
where we use the superscript to denote time and the subscript to denote spatial location. In Eq.
(8-18) the partial differential equation (PDE) is converted to the related finite difference equation
3/17/98
Introduction to CFD 8 - 11
and any errors due to treatment of BCs. A reexamination of the Taylor series representation is
worthwhile in thinking about the possible error arising from the discretization process:
f f ( x 0 + x) f ( x0 x)
=
+
x
2 x
x 2 3f
3
6 2x
1
4
4
3
(8-5a)
Thus we see that the size of the truncation error will depend locally on the solution. In most
cases we expect the discretization error to be larger than round-off error.
consistency
A finite-difference representation of a PDE is consistent if the difference between the PDE
and its difference representation vanishes as the mesh is refined, i.e.,
lim
mesh 0
(8-23)
mesh0
When might this be a problem? Consider a case where the truncation error is O(t/x). In
this case we must let the mesh go to zero just such that:
t
lim = 0
t,x0 x
(8-24)
Some finite difference representations have been tried that werent consistent. An example
cited by Tannehill, et al,1 is the DuFort-Frankel differencing of the wave equation.
stability
A stable numerical scheme is one for which errors from any source (round-off, truncation)
are not permitted to grow in the sequence of numerical procedures as the calculation proceeds
from one marching step, or iteration, to the next, thus:
errors grow unstable
errors decay stable
and
Stability is normally thought of as being associated with
marching problems.
Stability requirements often dictate allowable step sizes.
In many cases a stability analysis can be made to define the
stability requirements.
3/17/98
un+1
uni
n
n
n
i
=
2 ui+1 2ui + ui1
t
(x)
(8-25)
and the solution at time step n is known. At time n+1 there is only one unknown.
This is convergence with respect to grid. Another convergence requirement is associated with the satisfaction of the
solution of a system of equations by iterative methods on a fixed grid.
3/17/98
Introduction to CFD 8 - 13
x
i+1
x
x
i-1
n-1 n n+1
Figure 8-6. Grid points used in typical explicit calculation.
We solve for the value of u at i and the n+1 time step:
t n
n
n
un+1
= uin +
i
2 ui +1 2ui + ui1
(x )
(8-26)
x
i+1
i
i-1
n-1 n n+1
Figure 8-7. Grid points used in typical implicit calculation.
3/17/98
un+1
uni
n+1
n+1
i
=
ui+1
2uin+1 + ui1
2
t
(x)
(8-27)
where we use the spatial derivative at time n+1. By doing this we obtain a system where at each i
on n+1, un+1
depends on all the values at n+1. Thus we need to find the values along n+1
i
simultaneously. This leads to a system of algebraic equations that must be solved. For our model
problem this system is linear. We can see this more clearly by rearranging Eq. (8-27). Defining
=
t
( x )2
(8-28)
for i = 1,..., N .
(8-29)
This can be put into a matrix form to show that it has a particularly simple form:
(1 + 2 )
(1 + 2)
O
O
O
O
O
O
0
0
0
0
O
O
0
0
0
O
(1 + 2 )
O
0
0
0
O
O
0
0 u1n+1 u1n
0
0 u2n+1 u2n
O
O M M
O
O uin+1 = uin
O
O M M
n+1 n
(1+ 2)
uN
1 u N1
(1+ 2) uNn+1 u nN
(8-30)
Equation (8-30) is a special type of matrix form known as a tridiagonal form. A particularly
easy solution technique is available to solve this form. Known as the Thomas algorithm, the
details are described in Section 8.5 and a routine called tridag is described in Appendix H-1.
Many numerical methods are tailored to be able to produce this form.
The approach that leads to the formulation of a problem requiring the simultaneous solution
of a system of equations is known as an implicit scheme. To summarize:
The solution of a system of equations is required at each step.
The good news: stability requirements allow a large step size.
The not so good news: this scheme is harder to vectorize/parallelize.
A common feature for both explicit and implicit methods for parabolic and hyperbolic
equations:
A large number of mesh points can be treated, you only need the values at a
small number of marching stations at any particular stage in the solution.
This means you can obtain the solution with a large number of grid points
using a relatively small amount of memory. Curiously, some recent codes
dont take advantage of this last fact.
3/17/98
Introduction to CFD 8 - 15
2nd Example - elliptic PDE
We use Laplaces equation as the model problem for elliptic PDEs:
xx + yy = 0
(8-31)
y
j+1
j
j-1
i-1
i+1
(8-32)
yy =
(8-33)
and:
( x)
i, j+1 2i, j + i, j1
(y)
=0
(8-34)
(y)2
2 ( x)2 + (y )2
i+1, j + i 1, j +
(x)2
(i,j +1 + i,j1)
2 (x)2 + (y)2
(8-35)
where if x = y:
i,j =
3/17/98
(8-36)
3/17/98
(8-37)
Introduction to CFD 8 - 17
q=
F = u,
G = v
(8-38)
and recall that this conservation law could also come from the integral statement:
dV = V ndS .
t
(8-39)
Introducing the notation defined above and assuming two dimensional flow, the conservation
law can be rewritten as:
where
qdV + H n dS = 0
t
(8-40)
H = (F,G) = V
(8-41)
Hx = F = u
.
Hy = G = v
(8-42)
and
y, j
n
A
B
n
x, i
Figure 8-9. Basic nomenclature for finite volume analysis.
Using the definition of n in Cartesian coordinates, and considering for illustration the
Cartesian system given in Fig. 8-9, we can write:
H ndS = H xi + H y j n dS
= (Fi + Gj) ndS
(8-43)
(8-44)
H ndS = F dy
(8-45)
3/17/98
(8-46)
Using the general grid shown in the Fig. 8-10, our integral statement, Eq. (8-40) can be
written as:
DA
Aq j,k + ( Fy Gx) = 0.
t
AB
(8-47)
Here A is the area of the quadrilateral ABCD, and qi,j is the average value of q over ABCD.
j+1
k+1
j
C
j-1
k
D
cell centered
at j,k
B
k -1
x
Figure 8-10. Circuit in a general grid system.
Now define the quantities over each face. For illustration consider AB:
yAB = yB yA
xAB = xB x A
1
FAB = Fj,k 1 + Fj,k ,
2
1
GAB = Gj,k +1 + Gj,k
2
3/17/98
(8-48)
Introduction to CFD 8 - 19
Assuming A is not a function of time, and combining:
q j,k 1
1
+ Fj,k 1 + Fj,k yAB Gj,k 1 + Gj,k x AB
2
2
t
1
1
+ Fj,k + Fj+1,k yBC G j,k + G j+1,k xBC
2
2
1
1
(8-49)
+ Fj,k + Fj,k+1 yCD G j,k + G j,k+1 xCD .
2
2
1
1
+ Fj1,k + Fj,k yDA G j 1,k + Gj,k x DA
2
2
=0
Supposing the grid is regular cartesian as shown in Fig. 8-11. Then A = xy, and along:
BC:
CD:
y = 0, x AB = x
x = 0, yBC = y
.
y = 0, x CD = x
DA:
x = 0, yDA = y
AB:
(8-50)
y
k +1
k -1
j
j+1
j-1
Figure 8-11. General finite volume grid applied in Cartesian coordinates.
Thus, in Eq. (8-49) we are left with:
q j,k
1
1
Gj,k 1 + Gj,k x + Fj,k + Fj+1,k y
2
2
t
.
1
1
+ Gj,k + Gj,k +1 x Fj1,k + Fj,k y = 0
2
2
xy
Collecting terms:
3/17/98
(8-51)
(8-52)
or:
q F G
+
+
=0.
t x y
(8-53)
Thus, and at first glance remarkably, the results of the finite volume approach can lead to the
exact same equations to solve as the finite difference method on a simple Cartesian mesh.
However, the interpretation is different:*
Finite difference: approximates the governing equation at a point
Finite volume:
"
"
"
"
over a volume
Finite volume is the most physical in fluid mechanics codes, and is actually used in
most codes.
Finite difference methods were developed earlier, the analysis of methods is easier
and further developed.
Both the finite difference and finite volume methods are very similar. However, there are
differences. They are subtle but important. We cite three points in favor of the finite volume
method compared to the finite difference method:
Good conservation of mass, momentum, and energy using integrals when mesh is
finite size
Easier to treat complicated domains (integral discretization [averaging] easier to
figure out, implement, and interpret)
Average integral concept much better approach when the solution has shock waves
(i.e. the partial differential equations assume continuous partial derivatives).
Finally, special considerations are needed to implement some of the boundary conditions in
this method. The references, in particular Fletcher,3 should be consulted for more details.
8.4 Boundary conditions
So far we have obtained expressions for interior points on the mesh. However, the actual
geometry of the flowfield we wish to analyze is introduced through the boundary conditions. We
use an elliptic PDE problem to illustrate the options available for handling boundary conditions.
Consider the flow over a symmetric airfoil at zero angle of attack, as shown in Fig. 8-12.
3/17/98
Introduction to CFD 8 - 21
X
Figure 8-12. Example of boundary condition surface requiring consideration.
-
Here, because there is no lift, symmetry allows us to solve only the top half of the region. If
is a perturbation potential [see Chap. 2, Eq. (2-123)],
U = U + x
,
V = y
(8-54)
(8 55)
0 as x 2 + y 2 .
(8 56)
or
For a lifting airfoil, the farfield potential must take the form of a potential vortex singularity
with a circulation equal to the circulation around the airfoil.
The boundary condition on the surface of primary interest is the flow tangency condition,
where the velocity normal to the surface is specified. In most cases the velocity normal to the
surface is zero.
Consider ways to handle the farfield BC
There are several possibilities:
A.
go out far enough (?) and set = 0 for 0, as the distance from the body goes to
infinity (or v = 0, u = 0 where these are the perturbation velocities, or u = U if it is the
total velocity).
How good is this? This method is frequently used, although clearly it requires
numerical experimentation to ensure that the boundary is far enough from the body.
In lifting cases this can be on the order of 50 chord lengths in two-dimensions. In
3/17/98
Transform the equation to another coordinate system, and satisfy the boundary
condition explicitly at infinity (details of this approach are given in Chap. 9).
Figure 8-13 demonstrates what we mean. In the system the physical distance from 0
to infinity is transformed to the range from 0 to 1. Although this approach may lead to
efficient use of grid points, the use of the resulting highly stretched grid in the physical
plane may result in numerical methods that lose accuracy, and even worse, do not
converge during an iterative solution.
points evenly spaced in plane
x = is = 1
1.20
1.00
0.80
0.60
0.40
0.20
0.00
0.00
5.00
10.00
15.00
20.00
25.00
Figure 8-13. One example of a way to handle the farfield boundary condition.
C.
Blocks of Grids are sometimes used, a dense inner grid and a coarse outer grid. In
this approach the grid points are used efficiently in the region of interest. It is a simple
version of the adaptive grid concept, where the the grid will adjust automatically to
concentrate points in regions of large flow gradients.
D.
Match the numerical solution to an analytic approximation for the farfield boundary
condition.
This is emerging as the standard way to handle the farfield boundary conditions. It
allows the outer boundary to be placed at a reasonable distance from the body, and
properly done, it ensures that the boundary numerical solution reflects the correct
physics at the boundary. This has been found to be particularly important in the solution
3/17/98
Introduction to CFD 8 - 23
of the Euler equations. Effort is still underway to determine the best way to implement
this approach.
To summarize this discussion on farfield boundary conditions:
BCs on the FF boundary are important, and can be especially important for Euler
codes which march in time to a steady state final solution.
How to best enforce the FF BC is still under study - research papers are still being
written describing new approaches.
Consider ways to handle the nearfield BC
There are are also several ways to approach the satisfaction of boundary conditions on the
surface. Here we discuss three.
A. Use a standard grid and allow the surface to intersect grid lines in an irregular manner.
Then, solve the equations with BCs enforced between node points. Figure 8-14
illustrates this approach. In the early days of CFD methodology development this
approach was not found to work well, and the approach discussed next was developed.
However, using the finite volume method, an approach to treat boundary conditions
imposed in this manner was successfully developed (primarily by NASA Langley and its
contractors and grantees). It has not become a popular approach, and is considered to lead
to an inefficient use of grid points. Many grid points end up inside the body.
Irregular Intersection of
airfoil surface and grid
3/17/98
a. Entire geometry
b. Closeup of the trailing edge
Figure 8-15. Body conforming grid for easy application of BCs on curved surfaces.10
C. Another approach is to use thin airfoil theory boundary conditions, as described in detail
in Chapter 6. This eliminates many of the problems associated with the first two
approaches. It is expedient, but at some loss in accuracy (but very likely not that much, as
shown in Chap. 6, Fig. 6-14).
3/17/98
Introduction to CFD 8 - 25
When the solution requires that the gradient normal to the surface be specified, a so-called
dummy row is the easiest way to implement the boundary condition. As an example, following
Moran,4 consider a case where the normal velocity, v, is set to zero at the outer boundary. The
boundary is at grid line j = NY. Assume that another row is added at j = NY + 1, as indicated in
Fig. 8-17.
j = NY + 1
j = NY
j = NY - 1
i
Figure 8-17. Boundary condition at farfield.
The required boundary condition at j = NY is:
(8-57)
(8-58)
The equations are then solved up to Y NY, and whenever you need at NY +1, simply use the
value at NY-1.
Now, we present an example demonstrating the application of thin airfoil theory boundary
conditions at the surface. Recall that the boundary condition is:
v=
df
= U
y
dx
(8-59)
where yfoil = f(x). Assuming that in the computer code v has been nondimensionalized by U ,
the boundary condition is:
df
=
y dx
and the grid near the surface is defined following Fig. 8-18.
3/17/98
(8-60)
(8-61)
df
.
dx
(8-62)
Note that since j = 1 is a dummy row, you can select the grid spacing such that the spacing is
equal on both sides of j = 2, resulting in second order accuracy. Thus, as in the previous example,
anytime we need i,1 we use the value given by Eq. (8-62). Using these boundary condition
relations, the boundary conditions are identically satisfied. Note also that this approach is the
reason that in many codes the body surface corresponds to the second grid line, j = 2.
Imposition of boundary conditions is sometimes more difficult than the analysis given here
suggests. Specifically, both the surface and farfield boundary conditions for the pressure in the
Navier-Stokes and Euler equations can be tricky.
8.5 Solution of Algebraic Equations
We now know how to write down a representation of the PDE at each grid point. The next
step is to solve the resulting system of equations. Recall that we have one algebraic equation for
each grid point. The system of algebraic equations may, or may not, be linear. If they are
nonlinear, the usual approach is to form an approximate linear system, and then solve the system
iteratively to obtain the solution of the original nonlinear system. The accuracy requirement
dictates the number of the grid points required to obtain the solution. Previously, we assumed
that linear equation solution subroutines were available, as discussed in Chapter 3. However, the
development of CFD methods requires a knowledge of the types of algebraic systems of
equations.
3/17/98
Introduction to CFD 8 - 27
Recall that linear algebraic equations can be written in the standard form:
Ax = b .
(8-63)
For an inviscid two-dimensional solution, a grid of 100 x 30 is typical. This is 3000 grid points,
and results in a matrix 3000 x 3000. In three dimensions, 250,000 300,000 grid points are
common, 500,000 points are not uncommon, and a million or more grid points are often
required. Clearly, you cant expect to use classical direct linear equation solvers for systems of
this size.
Standard classification of algebraic equations depends on the characteristics of the elements
in the matrix A. If A:
1. contains few or no zero coefficients, it is called dense,
2. contains many zero coefficients, it is called sparse,
3. contains many zero coefficients, and the non-zero coefficients are close to
the main diagonal: the A matrix is called sparse and banded.
Dense Matrix
For a dense matrix direct methods are appropriate. Gauss elimination is an example of the
standard approach to these systems. LU decomposition11 is used in program PANEL, and is an
example of a standard method for solution of a dense matrix. These methods are not good for
large matrices (> 200-400 equations). The run time becomes huge, and the results may be
susceptible to round-off error.
Sparse and Banded
Special forms of Gauss elimination are available in many cases. The most famous banded
matrix solution applies to so-called tridiagonal systems:
b1
a
2
c1
b2
O
c2
O
ai
0
O
bi
O
ci
O
a N1
O
bN1
aN
x1 d1
x2 d2
M M
xi = di
M M
c N1 x N1 dN 1
bN x N dN
(8-64)
The algorithm used to solve Eq. (8-64) is known as the Thomas algorithm. This algorithm is very
good and widely used. The Thomas algorithm is given in detail in the sidebar, and a sample
subroutine, tridag, is described in App H-1.
3/17/98
i = 1,K,n
where a1 and cn are zero. The solution algorithm12 starts with k = 2,....,n:
m=
ak
bk 1
bk = bk mck 1 .
dk = dk mdk1
Then:
d
xn = n
bn
and finally, for k = n - 1,...1:
d c x
xk = k k k+1 .
bk
In CFD methods this algorithm is usually coded directly into the solution procedure,
unless machine optimized subroutines are employed on a specific computer.
General Sparse
These matrices are best treated with iterative methods. In this approach an initial estimate of
the solution is specified (often simply 0), and the solution is then obtained by repeatedly
updating the values of the solution vector until the equations are solved. This is also a natural
method for solving nonlinear algebraic equations, where the equations are written in the linear
equation form, and the coefficients of the A matrix are changed as the solution develops during
the iteration. Many methods are available.
There is one basic requirement for iterative solutions to converge. The elements on the
diagonal of the matrix should be large relative to the values off the diagonal. The condition can
be give mathematically as:
aii
aij
j=1
ji
3/17/98
(8-65)
Introduction to CFD 8 - 29
aii >
aij
(8-66)
j=1
ji
A matrix that satisfies this condition is diagonally dominant, and, for an iterative method to
converge, the matrix must be diagonally dominant. One example from aerodynamics of a matrix
that arises which is not diagonally dominant is the matrix obtained in the monoplane equation
formulation for the solution of the lifting line theory problem.
One class of iterative solution methods widely used in CFD is relaxation. As an example,
consider Laplaces Equation. Start with the discretized form, Eq.(8-31). The iteration proceeds
by solving the equation at each grid point i,j at an iteration n+1 using values found at iteration n.
Thus the solution at iteration n+1 is found from:
n+1
i,j
=
1 n
n
i+1, j + in1, j + i,n j+1 + i,j
1 .
4
(8-67)
The values of are computed repeatedly until they are no longer changing. The relaxation of
the values of to final converged values is roughly analogous to determining the solution for an
unsteady flow approaching a final steady state value, where the iteration cycle is identified as a
time-like step. This is an important analogy. Finally, the idea of iterating until the values stop
changing as an indication of convergence is not good enough. Instead, we must check to see if
the finite difference representation of the partial differential equation using the current values of
actually satisfies the partial differential equation. In this case, the value of the equation should
be zero, and the actual value of the finite difference representation is know as the residual. When
the residual is zero, the solution has converged. This is the value that should be monitored during
the iterative process. Generally, as done in THINFOIL, the maximum residual and its location
in the grid, and the average residual are computed and saved during the iterative process to
examine the convergence history.
Note that this method uses all old values of to get the new value of . This approach is
known as point Jacoby iteration. You need to save all the old values of the array as well as the
new values. This procedure converges only very slowly to the final converged solution.
A more natural approach to obtaining the solution is to use new estimates of the solution as
soon as they are available. Figure 8-19 shows how this is done using a simply programmed
systematic sweep of the grid. With a conventional sweep of the grid this becomes:
n+1
i,j
=
3/17/98
1 n
n
n+1
i+1, j + in+1
1, j + i,j+1 + i, j1 .
4
(8-68)
new values
available
X
X
n+1 1 n
n
n+1
i,j
= i+1, j + in+1
1, j + i,j+1 + i, j1
4
(8-69)
(i,n+1j i,n j ) .
(8-70)
3/17/98
Introduction to CFD 8 - 31
-0.02
-0.01
(16,2)
-0.01
-0.00
0.00
100
200
300
iteration number
400
500
3/17/98
X
X
X
X
X
X
X
X
X
starting upstream, move downstream, solving a line at a time.
Figure 8-21. Solution approach for SLOR.
200
Convergence tolerance
= 10-5
150
Number
of
Iterations
100
= 10-4
= 10-3
50
ONERA M6 Wing
M = 0.84
= 3
0
1.60
1.65
1.70
1.75
1.80
1.85
1.90
1.95
2.00
Figure 8-22. Effect of the value of on the number of iterations required to achieve
various levels of convergence.13
Another way to spread the information rapidly is to alternately sweep in both the x and y
direction. This provides a means of obtaining the final answers even more quickly, and is known
3/17/98
Introduction to CFD 8 - 33
as an alternating direction implicit or ADI method. Figure 8-23 illustrates the modification to the
SLOR method that is used to implement an ADI scheme. Several different methods of carrying
out the details of this iteration are available. The traditional approach for linear equations is
known as the Peaceman-Rachford method, and is described in standard textbooks, e.g., Ames8
and Isaacson and Keller.14 This approach is also known as an approximate factorization or AF
scheme. It is known as AF1 because of the particular approach to the factorization of the
operator. A discussion of ADI including a computer program is given in the first edition of the
Numerical Recipes book.15
Another approach has been found to be more robust for nonlinear partial differential
equations, including the case of mixed sub- and supersonic flow. In this case the time-like nature
of the approach to a final value is used explicitly to develop a robust and rapidly converging
iteration scheme. This scheme is known as AF2. This method was first proposed for steady flows
by Ballhaus, et al,16 and Catherall17 provided a theoretical foundation and results from numerical
experiments. A key aspect of ADI or any AF scheme is the use of a sequence of relaxation
paramters rathers a single value, as employed in the SOR and SLOR methods. Typically, the
sequence repeats each eight to eleven iterations.
Holst10 has given an excellent review and comparison of these methods. Figure 8-24, from
Holst,10 shows how the different methods use progressively better information at a point to find
the solution with the fewest possible iterations. The advantage is shown graphically in Figure 825, and is tabulated in Table 8-1 (also from Holst10). Program THINFOIL, described in Section
8.7, uses these methods and App. G-1 contains a description of the theoretical implementation of
these methods. Further details are given in Chapter 11, Transonic Aerodynamics.
3/17/98
Point Jacoby
Point Gauss-Seidel
Line Gauss-Seidel
SLOR
Line Jacoby
SOR
ADI
3/17/98
Introduction to CFD 8 - 35
100000
Point Gauss Seidel
& Line Jacobi
10000
Line
Gauss Seidel
1000
Point
Jacoby
n
SOR
100
SLOR
10
ADI
1
0.001
0.010
0.100
1.000
Figure 8-25. Comparison of convergence rates of various relaxation schemes (Holst10). This is
the number of iterations estimated to be required to reduce the residual by one order of
magnitude
Table 8-1
Convergence rate estimates for various relaxation schems (Holst 10)
Algorithm
3/17/98
Point - Jacobi
2/
1 / 2
1 / (2 )
Line - Jacobi
1 / 2
1 / (2 2 )
SLOR
ADI
1 / (2 2 )
log( / 2) / 1.55
(8-71)
2
2
+
.
x2 y2
(8-72)
where
L=
To solve this equation, we re-write the iteration scheme expressions given above in Equation.
(8-70) as:
N
Ci,n j
{
nth iteration
correction
L i,nj
1
23
= 0.
(8-63)
nth iteration
residual, = 0
when converged
solution is achieved
(8-64)
The actual form of the N operator depends on the specific scheme chosen to solve the problem.
8.6 Stability Analysis
The analysis presented above makes this approach to solving the governing equations for
flowfields appear deceptively simple. In many cases it proved impossible to obtain solutions.
Frequently the reason was the choice of an inherently unstable numerical algorithm. In this
section we present one of the classical approaches to the determination of stability criteria for use
in CFD. These types of analysis provide insight into grid and stepsize requirements (the term
stepsize tends to denote time steps, whereas a grid size is thought of as a spatial size). In
addition, this analysis is directly applicable to a linear equation. Applications in nonlinear
problems are not as fully developed.
Fourier or Von Neumann Stability Analysis
Consider the heat equation used previously,
u
2u
= 2
t
x
(8-17)
and examine the stability of the explicit representation of this equation given by Eq. (8-21).
Assume at t = 0, that an error, possibly due to finite length arithmetic, is introduced in the form:
3/17/98
Introduction to CFD 8 - 37
u(x,t)
123 = (t)
"error" is
introduced
jx
e{
(8-75)
where:
a real constant
j = 1
(8-21)
Substitute Eq. (8-75) into this equation, and solve for (t + t). Start with
(t + t)e jx (t)e jx
(t) j (x +x )
=
2e jx + e j (x x )
2 e
t
(x )
(8-76)
t
jx jx
e
2 + e jx
2 (t)e
(x)
14 4424 443
(8-77)
x +e jx
2+e1j
442
443
2cos x
Note that the e jx term cancels, and Eq. (8-77) can be rewritten:
t
(t + t) = (t)1 +
2 (2 + 2 cosx )
(x )
= (t) 1 2
1
cosx
424
3
(x)2 double1
angle formula
=12sin2 x
(8-78)
t
2 x
(t + t) = (t)1 + 2
2 1 1+ 2sin 2
(x )
t
2 x
= (t)1 4
sin
2
( x )2
3/17/98
(8-79)
G=
(t + t)
t
2 x
= 1 4
sin
.
(t)
2
( x )2
(8-80)
(8-81)
which means that the error introduced decays. For arbitrary , what does this condition mean?
Observe that the maximum value of the sine term is one. Thus, the condition for stability will be:
t
1 4
2 <1
(
x
)
1
424
3
(8-82)
(8-83)
(8-84)
and
1
2
t
1
2 <2
( x )
or:
t
1
2 <2.
( x )
(8-85)
(8-86)
This sets the condition on t and x for stability of the model equation. This is a real
restriction. It can be applied locally for nonlinear equations by assuming constant coefficients.
An analysis of the implicit formulation, Eq. (8-23), demonstrates that the implicit formulation is
unconditionally stable.
Is this restriction on t and x real? Rightmyer and Morton23 provided a dramatic example
demonstrating this criteria. Numerical experiments can quickly demonstrate how important this
condition is. Figure 8-26 repeats the analysis of Rightmyer and Morton,23 demonstrating the
validity of the analysis. The initial and boundary conditions used are:
3/17/98
Introduction to CFD 8 - 39
u(x,0 = (x) (given) for0 x
u(0,t) = 0, u(,t) = 0 for t > 0
Figure 8-26a presents the development of the solution and shows the particular choice of
initial value shape, , using a value of < 1/2: 5/11. Figure 8-26b-d provide the results for a
value of > 1/2: 5/9. Theoretically, this stepsize will lead to an unstable numerical method, and
the figure demonstrates that this is, in fact, the case.
Our model problem was parabolic. Another famous example considers a hyperbolic equation.
This is the wave equation, where c is the wave speed:
2u
2u
c
= 0.
(8-87)
t 2
x 2
This equation represents one-dimensional acoustic disturbances. The two-dimensional small
disturbance equation for the potential flow can also be written in this form for supersonic flow.
Recall,
(1 M2 ) xx + yy = 0
(8-88)
(8-89)
2
M
1
yy = 0
and we see here that x is the timelike variable for supersonic flow.
Performing an analysis similar to the one above, the stability requirement for Eq. (8-87) is
found to result in a specific parameter for stability:
=c
t
x
(8-90)
which is known as the Courant number. For many explicit schemes for hyperbolic equations, the
stability requirement is found to be
1.
(8-91)
This requirement is known as the CFL condition, after its discoverers: Courant, Friedrichs,
and Levy. It has a physical interpretation. The analytic domain of influence must lie within the
numerical domain of influence.
Recalling that the evolution of the solution for an elliptic system had a definite time-like
quality, a stability analysis for elliptic problems can also be carried out. For the SOR method,
that analysis leads to the requirement that the over-relaxation factor, , be less than two.
3/17/98
1.40
relative timestep
0
5
10
15
20
40
1.20
1.00
0.80
u(x,t)
0.60
initial condition
0.40
0.20
0.00
-0.5
0.0
1.5
2.0
2.5
3.0
x
a) numerical solution using a theoretically stable stepsize.
1.40
0.5
1.0
numerical solution
relative timestep
1.20
3.5
initial condition
5
1.00
stable solution
0.80
u(x,t)
0.60
0.40
0.20
0.00
-0.5
0.0
0.5
1.0
1.5
2.0
2.5
3.0
x
t
b) numerical solution using a theoreticaly unstable stepsize,
= 5.
(x)2
3.5
Figure 8-26. Demonstration of the step size stability criteria on numerical solutions.
3/17/98
Introduction to CFD 8 - 41
1.40
1.20
numerical solution
relative timestep
1.00
10
0.80
stable solution
u(x,t)
initial condition
0.60
0.40
0.20
0.00
-0.5
1.5
2.0
2.5
3.0
x
t
c) numerical solution using a theoreticaly unstable stepsize,
= 10.
(x)2
1.80
0.0
0.5
1.0
3.5
numerical solution
relative timestep
15
initial condition
1.20
u(x,t)
stable solution
0.60
0.00
-0.60
-0.5
0.0
0.5
1.0
1.5
x
2.0
2.5
3.0
3.5
t
= 15.
(x)2
Figure 8-26. Demonstration of the step size stability criteria on numerical solutions (concluded).
3/17/98
10
101
Maximum
Residual
SOR
100
10-1
AF2
-2
10
10-3
10-4
100
200
300
Iteration
400
500
600
3/17/98
Introduction to CFD 8 - 43
engineering practice is to consider the solution converged when the residual is reduced by 3 or 4
orders of magnitude. However, a check of the solution obtained at a conventional convergence
level with a solution obtained at much smaller residual (and higher cost) level should be made
before conducting an extensive analysis for a particular study.
The solution for a 5% thick biconvex airfoil obtained with THINFOIL is presented in
Figure 8-28, together with the exact solution. For this case the agreement with the exact solution
is excellent. The exact solution for a biconvex airfoil is given by Milton Van Dyke,24 who cites
Milne-Thompson25 for the derivation.
-0.30
5% Thick Biconvex Airfoil
(THINFOIL uses thin airfoil theory BC's)
-0.20
-0.10
Cp
0.00
0.10
Cp (THINFOIL)
Cp (Exact)
0.20
74 x 24 grid
0.30
0.0
0.2
0.4
0.6
0.8
1.0
X/C
Figure 8-28. Comparison of numerical solution with analytic solution for a biconvex airfoil.
The material covered in this chapter provides a very brief introduction to an area which has been
the subject of an incredible amount of research in the last thirty years. Extensions to include
ways to treat flows governed by nonlinear partial differential equations are described after some
basic problems in establishing geometry and grids are covered in the next chapter.
3/17/98
3/17/98
Introduction to CFD 8 - 45
8.9 References
1
Tannehill, J.C., Anderson, D.A., and Pletcher, R.H., Computational Fluid Mechanics and Heat
Transfer, 2nd Ed., Taylor & Francis, New York, 1997. (on the first edition the order of the
authors was Anderson, D.A., Tannehill, J.C., and Pletcher, R.H)
2
Hoffman, K.A., and Chiang, S.T., Computational Fluid Dynamics for Engineers, in two
volumes, Engineering Education System, PO Box 20078, Wichita, KS, 67208-1078. 1993.
3
Fletcher, C.A.J., Computational Techniques for Fluid Dynamics, Vol. 1: Fundamental and
General Techniques, Vol. II, Specific Techniques for Different Flow Categories, SpringerVerlag, Berlin, 1988.
4
Moran, J., An Introduction to Theoretical and Computational Aerodynamics, John Wiley &
Sons, New York, 1984.
5
Hirsch, C., Numerical Computation of Internal and External Flows, Vol. 1, Fundamentals of
Numerical Discretization, 1988, and Vol. 2, Computational Methods for Inviscid and Viscous
Flows, 1990, John Wiley & Sons, New York.
6
Anderson, J.D., Jr., Computational Fluid Dynamics: The Basics with Applications, McGrawHill, Inc., New York, 1995.
7
Smith, G.D., Numerical Solution of Partial Differential Equations: Finite Difference Methods,
3rd Ed., Clarendon Press, Oxford, 1985.
8
Ames, W.F., Numerical Methods for Partial Differential Equations, 2nd Ed., Academic Press,
New York, 1977.
9
Versteeg, H.K., and Malalasekera, W., An Introduction to Computational Fluid Dynamics: The
Finite Volume Method, Addison Wesley Longman, Ltd., Harlow, England, 1995.
10
Holst, T.L., Numerical Computation of Transonic Flow Governed by the Full-Potential
Equation, VKI Lecture Series on Computational Fluid Dynamics, Rhode-St.-Genese, March,
1983.
11
Press, W.H., Flannery, B.P., Teukolsky, S.A., and Vettering, W.T., Numerical Recipes in
FORTRAN: The Art of Scientific Computing, 2nd ed., Cambridge University Press, Cambridge,
1992.
12
Conte, S.D., and deBoor, C., Elementary Numerical Analysis, McGraw-Hill, New York, 1972.
13
Mason, W.H., Mackenzie, D., Stern, M., Ballhaus, W.F., and Frick, J., An Automated
Procedure for Computing the Three-Dimensional Transonic Flow Over Wing-Body
Combinations, Including Viscous Effects, Vol. II Program Users Manual and Code
Description, AFFDL-TR-77-122, Feb. 1978.
14
Isaacson, E., and Keller, H.B., Analysis of Numerical Methods, John Wiley, New York, 1966.
15
Press, W.H., Flannery, B.P., Teukolsky, S.A., and Vettering, W.T., Numerical Recipes: The
Art of Scientific Computing (FORTRAN Version), Cambridge University Press, Cambridge,
1989.
16
Ballhaus, W.F., Jameson, A., and Albert, J., Implicit Approximate Factorisation Schemes for
the Efficient Solution of Steady Transonic Flow Problems, AIAA J., Vol. 16, No. 6, June 1978,
pp. 573-579. (also AIAA Paper 77-634, 1977).
17
Catherall, D., Optimum Approximate-Factorisation Schemes for 2D Steady Potential Flows,
AIAA Paper 81-1018, 1981.
18
Jameson, A., Acceleration of Transonic Potential Flow Calculations on Arbitrary Meshes by
the Multiple Grid Method, AIAA Paper 79-1458, Proceedings of the AIAA 4th Computational
3/17/98
3/17/98