Tudeift: Numerical Methods in Aircraft Aerodynamics
Tudeift: Numerical Methods in Aircraft Aerodynamics
Tudeift: Numerical Methods in Aircraft Aerodynamics
CONTENTh Page
O. Preface O-1
doublet distributions
2.1.12 Summary of well-posed and some ill-posed boundary value 2-45
problems and integral representations
2.1.13 The application of the integral representation to the 2-47
integral equations
2.2.2 Surface-grid generation ('paneling') 2-66
over a panel
Aerodynamic influence coefficients 2-7'-t
2.2.5
2.2.6 Truncation errors; consistent approximations 2-86
compressible flow
2.4 Examples of Application (3D) of 'Ordinary Panel Methods 2-129
O. PREFACE
The contents of the course is of course limited in width and depth, in the sense
that the emphasis is on methods for stationnary subsonic and transonic flows.
The student is encouraged to consult the following literature for further study.
General References
Every student that is confronted with a new subject will probably immediately
feel the following questions rising within himself:
What is this about?
Why is it there?
How is it done?
To answer these questions completely probably takes a life-time but at least a
full course of lectures.
About 'WHAT?'
Numerical (or computational) Aerodynamics is a sub-discipline of Computational
Fluid Dynamics ('CFD' for short).
CFD is concerned with research, development and application of methods for
constructing numerical approximations to solutions ('numerical solutions') of
the partial differential equations that describe the motion of fluids and gasses
within or around bodies.
Numerical (aircraft) aerodynamics is concerned with the application of CFD in
aircraft aerodynamics.
Numerical Aerodynamics is also
- sitting at a graphics work-station to
generate and inspect geometries
generate and inspect meshes and grids (which form the basis of numerical
discretization) see figs. 1.1 - 1.5.
'post-process' (visualize) the results
- 'digging' in listings of codes to find the 'bug'
- sitting at your desk to
think about
how to attack a problem (make a plan!)
the error you made
call the guy that gave you the wrong data
call the guy that developed the code because it would not run (properly)
write your report
l-2
About 'WHY?'
Numerical Aerodynamics, as compared to windtunnel testing offers complementary
possibilities that. may improve the efficiency of the aerodynamic design process.
In particular it can lead to
- reduction of time and/or cost of design and development
- product/quality improvement
- improved accuracy of performance estimates in the early stages of the design
process (aircraft manufacturers must give performance guarantees to airlines
that 'buy from the drawing board'; hence reduced risk).
About 'HOW?'
In the process of development and application of numerical methods for solving
physical problems we may distinguish the following steps
definition and description of the physical problem
formulation of the corresponding well-posed analytical-mathematical
problem (mathematical-physical model)
formulation of a numerical model that discretises (approximates) the
analytical-mathematical model and its solution which sufficient accuracy
and efficiency
formulation of algorithm(s) (set of calculation rules) for the efficient
execution of the solution process for the numerical model
y) coding of the algorithm(s)
('writing' the computer program (or 'code'))
verification (checking that the code is mathematically and numerically
sound and self-consistent)
validation (checking and establishing the boundaries of the area of
applicability of the code)
l-3
In case (candidate) codes are already available steps iii) to vi) or vii) can be
replaced by 'selection of code'.
In the following paragraphs we will look at some of these steps in a little more
detail.
- friction effects that are limited to a relative thin layer (boundary layer)
adjacent to the surface of the configuration (high Reynolds numbers)
- 'smooth' flow 'separation' at the (sharp) trailing edge of wing-like
components and a thin 'trailing vortex wake'
- compressibility effects
- disturbances that, at least in subsonic flow (Mach number <1,) vanish at a
large distance from the configuration (except, possibly in the vicinity of the
trailing vortex wake).
The conservation laws or PDE's model the conservation of the flow quantities
mass
momentum (3 components!)
energy
supplemented with the
thermal and caloric equations of state (usually for a perfect gas)
and, for viscous, heat conducting fluids, expressions for
the stress tensor
viscosity coefficient
heat conduction coefficient
Conservation laws for flow quantities U within a volume with bounding surface
S can in the absence of internal and boundary surface sources, be written in the
integral form
f UdQ + d = O
Q S
(1.1.2)
+ O
l-5
The most complete description of the flow of a fluid continuum is given by the
Time-dependent Navier-Stokes equations. The table below lists the phenomena
contained by this flow model.
1-6
The time-dependent Navier-Stokes equations can be written in the form (see eq.
(1.1.2))
au -
= o (1.1.3)
+ v.(
at mv + visc )
au a
p (ri. + H = O (1.1»)
+ - (F.
ax mv + F
visc. )
+ ay
- mv +
(G. G
visc
. ) + az iriv visc J
.
at
au
at
+
- (.mv
a
a
+ visc ) + - (.mv
an
+ visc J + aç- (iTi.
mv + H
visc )
= 0 (1.1.5)
with U = TJ/J
= k1 + G +
(1.1.6)
G = (n F + n G + n H)/J
= + G +
l-7
It is noted that the energy equation for a viscous, heat conducting fluid
expresses that the local rate of change of total energy (kinetic plus internal)
is the result of the sum of transport of total energy through convection,
transport of heat through conduction and the mechanical energy delivered by
viscous forces.
Furthermore, the more detailed structure of the equations implies (ref. 1.2)
that the transformation of kinetic energy into heat through viscous forces
(dissipation) is irreversible. The same holds for the transformation of heat
into kinetic energy through heat conduction. As a consequence of these
irreversible processes the entropy of a (moving) particle of fluid cannot
decrease (2nd law of thermodynamics).
Numerical flow simulation based on the time-dependent N-S eqs. is often referred
to as direct simulation.
The practical and biggest problem of direct simulation is formed by the fact
that almost all flows of aerospace interest are turbulent. Turbulent flows are
characterized by the presence of time and space dependent fluctuations of all
the flow quantities. In many cases the level of' the turbulent fluctuations can
attain 10% or more of the mean values of the flow quantities. The biggest part
of the problem is, however, in the fact that space- and time-scales of the
turbulent fluctuations can vary enormously. The numerical resolution, with
sufficient accuracy, of the small-scale fluctuations in particular is a
formidable computational task. It requires such small space- and time steps that
for all practical purposes the computational effort is prohibatively large for a
long time to come, even for the biggest and fastest supercomputers.
Reynolds -averaging
The next (lower) level in the hierarchy of flow models therefore involves
application of a time averaging process to the turbulent fluctuations in order
to obtain laws for 'mean', time-averaged, turbulent quantities. The time-
averaging is to be done in such a way that the time-dependence of other
phenomena with time scales different from those of turbulence are not destroyed.
1-8
(1.1.7)
p
with
T/2
A = f A(x,y,z;t+T)d-t (1.1.8)
-T/2
and
A = A + A" (1.1.9)
pA" = 0 (1.1.10)
The RANS eqs contain all the terms of the original time-dependent N-S eqs
applied to the mean flow plus a number of additional terms. The additional terms
arise as a result of the non-linear character of the N-S equations. They appear
where (vector) products of quantities are to be taken, as a consequence of the
fact that the average of a product of fluctuating quantities,
pA"xA" 0 (1.1.11)
even if
pA" = O
way similar to the viscous stresses. Hence, the RANS eqs can be written in a
form similar to eq. (1.1.4), (1.1.5). In the energy equation additional terms
appear that are related to (turbulent) heat conduction.
Unfortunately the HANS eqs no longer form a closed system of equations; the
number of unknowns is no longer equal to the number of equations because of the
appearance of additional terms of the type (1.1.11). Apparently, the time-
averaging process gives rise to 'loss of information'. This loss of information
must be compensated for by explicitly adding external information from other
(experimental) sources.
The 'art' of turbulence modelling is now to 'close' the system of equations by
relating the additional, turbulent terms (Reynolds stresses) in some way to the
mean flow quantities.
The table below lists the flow phenomena that are modelled through the RANS
equations
1-lo
Time-dependent As N-S,
Navier-Stokes with but only for phenomena with length and time
Reynolds-Averaged As N-S,
It must also be mentioned that there exist two (simplified) subsets of the RANS
equations that are often used in (the emerging) practice. One involves a thin
shear layer approximation leading to the Thin layer N-S equations (TLNS). In
TLNS it is assumed that the dominating influence of the viscous and turbulent
terms come from the gradients transverse to the main flow direction, which would
be appropriate for thin shear layers. In terms of eq.(1.1.6) the TLNS eqs take
the form
aF. aG.
-+
au mv + + p- (H mv + FI . = O (1.1.12)
at a aç visc }
3F. 3G.
mv +
mv +
3
an inv1i visc )
. =0 (1.1.13)
Zonal Modelling
It appears that further simplifications of the (RA)NS equations are not possible
without distinguishing and separating viscous and inviscid parts in the flow
field. It was first recognized by Prandtl (1904) that at high Reynolds numbers
without significant flow separation the (direct) influence of the viscous and
turbulent shear stresses is limited to a thin layer close to the wall (boundary
layer) and that outside these layers the flow behaves as inviscid. In Prandtl's
theory a simplified boundary layer approximation (of the NS equations) suffices
for the determination of the viscous effects, while the (indirect) effects
thereof on the outer inviscid flow can be represented through the concept of
(boundary layer) displacement thickness. More recently Prandtl's theory has been
reformulated in terms of the theory of matched asymptotic expansions (see, e.g.
ref.l.3).
Clearly such zonal modelling requires some form of interaction between the
boundary layer computations and the computation of the outer inviscid flow; the
'inner' and 'outer' solutions must be matched at a common interface: the edge of
the boundary layer.
Prandtl's theory also teaches that for attached flow with thin boundary layers a
fully inviscid approximation is a valid and consistent one for many flow
properties.
1-12
The most complete inviscid flow model is obtained by setting the viscous and
heat conduction terms in the NS equations equal to zero, which gives (see eqs.
(1.1.4), (1.1.5))
aG. 8H.
-+
au mv +
mv +
mv -O (1 . 1 . 14)
at ax ay az
or
au
at
+ a
a
(.mv ) + - (,mv
an
+ !_
az
(T
) = (1.1.15)
The next lower level of approximation for inviscid flow is obtained by assuming
irrotationality, leading to the Full Potential (FP) equation:
+ .(pU) = o
p = p(U) (1.1.16)
U=
J
The most important aspect of the full potential equation is that it contains
only one dependent variable: the velocity potential . Recalling that the Euler
equations form a system of 5 equations with 5 dependent variables this is,
indeed, a considerable simplification. The price to be paid is a further
reduction of the number of phenomena that are (properly) modelled (see below).
xx
+ +
zz
=0
yy
= i -
= - U.x (1.1.17)
Ihi « ü
This is a linear partial differential equation which offers a further
computational advantage as compared to the FP equation.
,
xx
+$ +4
zz
=0
yy
or (1.1.18)
xx
+ +
zz
=0
yy
a a
(puw) + - (puw) + - pw 2 + p + H
a a
(pw) + = O
aç visc
. ) (1.1.19)
reduces to
aç
(1 . 1 . 20)
or p(,iì,ç)
where PeRfl) represents the pressure at the edge between boundary layer arid
outer inviscid flow.
The implication of (1.1.20) is that the pressure no longer appears as a
dependent variable but as a known external driving force which is to be obtained
from an inviscid flow computation.
The set of equations obtained in this way is known as the Boundary Layer (BL)
Equations. In the symbolic notation of eq.(1.1.5), (1.1.6) they take the saine
form as the TLNS equations (1.1.12). However, the vector U of conserved
quantities and generalised flux vector components contain one unknown less (the
pressure p) and less terms.
l-16
It is further noted that as with the N-S equations we may distinguish between
Time-dependent, LES and Reynolds Averaged Boundary Layer equations.
Time-Dependent as N-S,
App1icabi1it
The various flow models mentioned above all have their own area of applicability
in aircraft aerodynamics. Qualitatively this can be indicated as in fig. 1.12.
The figure illustrates the areas in the a-Mach plane within the flight envelope
of a transport-type aircraft where the various flow models are valid.
The Navier-Stokes equations are valid in the whole of the a-Mach or CL_Mach
plane (provided the representation of turbulence is adequate). The other flow
models have more limited regions of applicability in the sense that the higher
the model ranks in the hierarchy of fig. 1.11 the larger is its region of
applicability.
i-18
Instead of the differential form we can also apply the integral form of the
conservation laws directly to a volume element or cell of the space in which the
solution is sought.
For example, in case of the mass conservation law, this leads to (fig. 1.114)
defining values of the (as yet unknown) flux vector pV at suitably chosen 'nodal
points' in the computational grid and expressing the variation of the integrand
over the cell faces by means of a Taylor series expansion around the nodes.
Satisfying eq (1.2.1) for all volume elements, while satisfying the boundary
conditions for the fluxes in the cell faces at the boundary, one obtains, again,
a system of algebraic equations with a bandstructured matrix.
Methods of this kind are called FINITE VOLUME METHODS(FVM).
The variation of the integrand along the cell faces can also be expressed in
terms of (suitably chosen) polynomials with the function values in the nodes as
parameters. This case is, sometimes, associated with the notion of FINITE
ELEMENT METHODS IFEMJ. However, the notion of finite element methods is
generally restricted to a class of methods that is based on so-called
variational principles.
div. grad = O
(1.2.2)
It can be shown that solving (1.2.2) is equivalent with minimizing the volume
integral
f!! [grad ]z dQ
Q
(1.2.3)
advantage is that the manual as well as the computational effort required for
the generation of the unstructured spatial grids is significantly less than for
regular, structured grids. A disadvantage is that the structure of the resulting
matrices is also less regular than in case of FDM's. As a consequence the
computational effort associated with FEM's is generally larger.
FVM's can also be formulated such that they can make use of unstructured grids.
Such 'unstructured-grid' FVM's have gained in importance since the late 1980's.
For a more general introductory discussion on the differences and similarities
between FDM, FVM and FEM see e.g., ref. 1.2.
Certain PDE's, for which elementary solutions are known, can be transformed into
an integral equation (by means of Green's theorem, about which later, in section
2.1.6). The solution of a boundary value problem for such PDE's can then be
expressed in terms of integrals over boundary and volume distributions of
elementary solutions (sources/sinks, vortices) of, as yet, unknown strength;
e.g.
13
= JI oCx.) K(x.,x.)dS + III o(x k
S
)
13
K(x.,x.)dQ (1.2.4)
surface mesh leads to a system of (10 to 10') linear algebraic equations for
the unknown source/sink or vortex strengths. In this case the associated matrix
is full ( no zero entries).
Methods of this kind are called BOUNDARY ELEMENT or PANEL METHODS. In Numerical
(Aircraft) Aerodynamics they play an important role since about 1970.
1-21
The COMPUTER POWER required by the various classes of methods varies strongly
and depends on
- mathematical/physical model (type of PDE's)
- numerical model
- required numerical accuracy of solution
- turn-around time required
Fig. 1.17 gives an indication of the computer time required by various models
for one flow computation for a wing-fuselage configuration. A modest
'engineering' level of accuracy has been assumed and several types of existing
computers are considered.
Fig. 1.18 illustrates the computer capacity, in terms of processing speed and
memory required for a 'turn-around' time of about halve an hour, again assuming
a modest engineering accuracy.
Apart from a 'flow solver' the application of CFD requires extensive facilities
for PRE- and POSTPROCESSING such as
- GEOMETRY HANDLING
- GRID GENERATION
- GRAPHICS for VISUALIZATION of input (geometry, grids) and output (numerical
flow visualization)
The complexity of the complete CFD process requires an INFORMATION SYSTEMS
approach involving METHOD BASE and DATA BASE MANAGEMENT, EXECUTIVE and GRAPHICS
SOFTWARE, etc (fig. 1.19).
The COSTS associated with CFD development and application are substantial.
For instance, the investment associated with the development of a professional,
production oriented computer program is of the order of one to several million
US $.
The direct computation cost involved with a single flow computation (3D) may
vary between a few hundreds and several thousands US $.
l-23
1 .4 State-of-the-art
FUTURE PROSPECTS
Because of continuing developments in computer technology, informatics and
numerical mathematics, Numerical Aerodynamics still has an enormous growth
potential.
The need for both 'simple' (that is fast and cheap) and accurate (that is
computationally intensive and expensive) methods will remain, due to different
requirements with respect to processing speed and accuracy in the various stages
of the aerodynamic design process.
With the objective of' improving design integration computational aerodynamics
methods will, increasingly, be integrated with computational models from other
disciplines such as structures, flight mechanics, etc. (ref. 1.1).
l-24
References
2. PANEL METHODS
Panel methods are numerical methods for the solution of partial differential
equations governing potential flow, i.e. inviscid, irrotational flow. They are
based on surface distributions of sources and vortices or doublets. Within the
restriction of potential flows such methods can be applied to compute the flow
past complex geometries in two and three dimensions. These geometries are
approximated by a large number of surface elements or 'panels'. Depending on the
number of panels any degree of accuracy may be obtained in principle.
xx
+
yy
+
zz
=0 (2.1.1)
V = grad (2.1.2)
and the subscripts denote double differentiation with respect to x,y or z, the
coordinates of a right handed coordinate system. For steady compressible flow
the linearized potential equation may be written in terms of the perturbation
potential ' as
(1 - M2) + + = o (2.1.3)
xx yy zz
where is defined by the free stream velocity U and the total velocity
potential as
= Ux + (2.1.4)
2-2
and M is the free stream Mach number. The linearized potential equation is also
called the Prandtl-Glauert equation. The potential equations (2.1.1) and (2.1.3)
follow from the law of mass conservation
+ + = 0 (2.1.7)
xx yy zz
If the potential ' or is known, the pressure may be obtained from the momentum
equation. For incompressible flow this takes the form of the Bernoulli equation
p+ 1
p
(2 + + 2) = + ipU2
2
(2.1.8)
x y z
PP,
or, in terms of the pressure coefficient C -
pU2
+ 2 + 2 -
IvIz (2.1.9)
1 = 1
p U2 U_
<< U (2.1.10)
u,v,w = , ,
X y ¿
1 (U+ )2 + +
X y z
= = [i M2 [ U2
(2.1.11)
2-3
the components of p'? in the continuity equation (2.1.5) are then expressed in a
series expansion of 4) 4) , 4) as
X y z
4)
= p U [i + (1-M2) +
Uo
4)
p
y = p U [-y + (2.1.12)
4)
z
p
Z
= pU U[+
o
. [(i-w)
ax
4)
x]
+ay. [4)
j + 3z [ J =o (2.1.13)
)2 + 4)2 4)2
(U o + 4) +
= (i _iM2 I L
Y z]T-1
(2.1.14)
Po O
or from
- 1
(2.1.15)
P 1pU 1M2 P0
2 2°
Note: In the remaining of this chapter we will understand by and 4) the non-
dimensionalized quantities /U and 4)/U.
2-4
P=1_Mz
p
(2+ ..) 2 U (2.1.16)
C = - 2 - + (2.1.17)
p U
Since Laplace's equation is the limit for M40 of the Prandtl-Glauert equation,
eq. (2.1.3), the latter has a larger range of applicability than the first.
Laplace's equation is valid for incompressible, irrotational flow, whereas the
Prandtl-Glauert equation may be applied in compressible, irrotational flow
subject to small perturbations and in incompressible, irrotational flow with no
restrictions regarding the perturbations. Therefore most panel methods are based
on the Prandtl-Glauert equation. In general the pressure is computed fom eqs.
(2.1.14) and (2.1.15), which reduce to the exact expressions (2.1.8) and
(2.1.9), respectively, for M+O. Note that this feature is not present if
instead of eqs. (2.1.14) and (2.1.15), eqs. (2.1.16) and (2.1.17) are used.
xx V V
+
yy V V
+
zz'
=0 (2.1.18)
by the transformation
X' =
=y (2.1.19)
= z
or by
X' = X
(2.1.20)
=
=
2-5
where = 1_M2.
Types of problems
where s,t are coordinates along the boundary. In this case the normal velocity
component is obtained from the solution. The Dirichlet problem is related to
the problem of designing the shape of a body for a given velocity or pressure
distribution.
Well-posed problems
(2.1.21)
1f an
dS = O
s
or destroyed.
The solution for the potential that exists inside S under the condition
(2.1.21) is not unique, because an arbitrary constant, say, satisfying
c
2-7
- Neumann-problem for an exterior domain. This problem may be obtained from that
depicted in fig. 2.l.7b for S9. The not well-posedness of the Neumann-
problem for an exterior domain with = O on S. (fig. 2.1.7b) and S* seems,
at first sight, to constitute a severe handicap for applications to aircraft
aerodynamics where we are particularly interested in such problems. However,
we shall see shortly that this is not really the case.
The mixed or Poincar problem for an interior domain bounded by a closed surface
(fig. 2.1.8) may be shown to be well-posed. No additional condition for j needs
to be prescribed. The mass flow imposed on a part of the boundary by prescribing
is absorbed by the part where is prescribed. The latter guarantees
uniqueness of the potential and as a consequence the uniqueness of the problem.
Note that in order to satisfy Laplace's equation must be twice continuously
differentiable in points of the boundary (A and B in fig. 2.1.8) where the type
of the boundary conditions changes. Because of the relevance for the aerodynamic
computations encountered in the aeronautical design problems we will analyse the
mixed boundary value problem in some more detail.
= g(s,t) 4 X
or for S (S )
(2.1.22)
o
(4
s0 40
B
!! ds = (2.1.23)
as
A
The double brackets ] denote a jump in the value of the quantity inside.
Realizing that a potential flow does not contain rotation, eq. (2.1.23) may be
recognized as the expression for the circulation r around a cross-section 35B of
(see text books on elementary aerodynamics).
5B
For a continuous potential along as (as well as 3SB) we have
r = hm 0 (2.1.24)
B9A
between the points A and B on as and the cross-section aSB of the body. Across
the cut the potential is allowed to jump from a level to according to
2'
where we have formally assumed that the jump is a function h(s,t) of the local
coordinates. In fig. 2.1.11 a sketch is given of the model representation with a
cut. The circulation around the sectìon 35B of' fig. 2.1.11 is then
(F)as L (2.1.26)
Note that the introduction of the cut has reduced the original multiply-
connected domain into a simply-connected domain. Without proof we mention here
that a well-posed problem with circulation can only be formulated for a simply-
connected domain. Every cross-section of the type eQ needs a cut to produce a
well-posed problem with circulation; in three dimensions we thus obtain a
We may further notice, that since (s,t) should be continuous over the bounding
surfaces, we must require, at the edges of the discontinuity surface Sc, that
ir = o (2.1.27)
In addition it may be clear that because of the jump in the potential across Sc
we have to abandon the Dirichlet condition (2.1.22) on S in the vicinity of s;
we come back on this later.
2-10
we achieve this?
The physics of the flow require conservation of mass, momentum and energy. This
should also hold across S.
Conservation of mass across SC involves that no mass is created nor destroyed,
which can be stated as
=o
° (2. 1 .28)
=
[p + (-) a a
]j
=o (normal momentum) (2.2.29a)
and
a
LE aíp) j:B1 = o (2. 2. 29b)
(tangential momentum)
2
LE (p) j Ill
a 3
0 (2.2.29c)
(2.2. 30a)
= o
2-il
and
a a
P =o (2.2.30b)
a a
p j =o (2.2.30c)
respectively.
Since
a a
a a
=
(p Jl = (2.2.31a)
and
=
an2 (2 .2. 31b)
We note that eqs (2.2.31 a,b) express that S must be a stream surface.
Unfortunately we do not know the exact shape and position of a stream surface a
priori, without knowing the solution. What we do know, however, is that if the
flow separates somewhere on the body a stream surface eminates from the line at
which the flow separates into the flow field. In case of sharp-edged bodies,
like a wing the flow is known to separate at the sharp (trailing) edge. This
then suggest that a surface approximating the stream surface from the sharp
(trailing) edge represents a proper a priori choice for the discontinuity
surface S. In case of small perturbations streamlines in the flow will almost
be parallel to the undisturbed flow. In such conditions S may chosen to consist
of straight generators, parallel to the x-axis, eminating from the trailing
edge.
2-12
With such an a priori chosen shape and position of S the conditions (2.2.30a)
and (2.2.31a and b) are one too many for a properly posed boundary value
problem. In addition we have the complication that the zero pressure jump
condition (2.2.30a) is a non-linear one. The general approach is then to replace
the two conditions (2.2.31 a plus b) by the single (weaker) condition (2.1.28).
This, of course, still guarantees conservation of mass.
Using the Bernoulli eq. (2.1.8) the non-linear zero pressure jump condition
(3.2.30a) can be written as
onS0 (2.1.32)
For small perturbations (see eqs. (2.1.16), (2.1.17) this reduces to the
approximation
cP = O on S (2.1.33)
tangential coordinates as
+ 2 + = O on Sc (2.1.35)
IL-n s t 1
(2.1.36)
lt + = O OflSc
These expressions may be used if we posess a more accurate estimate of the flow
direction on Sc than the free stream direction.
2-13
The coordinates (s,t) may then be chosen such that f = 0 In this case the
s- direction, denoted by s, coincides with that of the component, in S, of the
average streamline of S (see fig. 2.1.13). Eq. (2.1.36) then reduces to the
linear expression
[L * :11=0 on Sc
s
or
*
it: Ji = const along s
Kutta condition
In the preceding section we have shown that a jump in the potential across the
discontinuity plane Sc is necessary in order to model circulation. However, the
magnitude of the jump has not yet been stipulated. It appears to be sufficient
that the jump [ ] is given along SB the intersection of 5c with SB (see fig.
2.1.12); the distribution on Sc then follows from eqs. (2.1.34) or eq. (2.1.38).
considering again the physics of our problem (fig. 1.10) we observe that the
condition that the flow separates 'smoothly' from the (sharp) downstream end SB
of' the body has not been satisfied yet. Here, basing on heuristic, physical
arguments, 'smoothly' is to be interpreted as with finite velocity () that is
continuous when passing from the body onto S across the trailing edge.
This condition is known as the condition of Kutta-Youkowski or simply as the
Kutta condition. The situation suggests that JJ along SB be chosen such that
the Kutta-condition is satisfied.
At this point it should be noted that the more formal argument that Laplace's
equation (2.1.1) or (2.1.6) must be satisfied everywhere in the flowfield, the
line
5b on the bounding surface SB+SC included, leads to the same requirement,
namely that is finite and continuous when passing from onto Sc. Hence, we
2-14
may conclude that if we make sure that Laplace's equation is satisfied at the
trailing edge we automatically will also satisfy the Kutta condition.
Recall also in this context, that, in section 2.1.3, we anticipated that special
measures were to be taken at the intersection of S with SB in order to restore
the conditions for a properly posed boundary value problem (that is that the
solution is continuous and differentiable twice ("C2") everywhere in the flow
field). Apparently, the Kutta condition serves as such.
Finite and continuous velocity requires in the first place a continuous
potential across the trailing edge, which can be expressed as (see fig. 2.1.15
for the position of the points P)
(Pj)
or
Because eq. (2.1.39a) directly involves the jump in potential across S we must
expect that this is the governing condition for determining the circulation.
(P )
1
or
(P2) - (P) +
for P1, P, P2 P2 + T (2.1.40)
where the latter equality results from the condition that ff = O across
Sc;V denotes the gradient in S. Eq. (2.1.41) puts a requirement on the jump of
the tangential velocity across S.
Addition of the expressions (2.1.40) results in
- [(P1))2 +
(()}2 - (P))2 =
2 p,
2
The condition given by eq. (2.1.43) may also be obtained by requiring continuity
of pressure, because this means (according to the Bernoulli equation, eq.
(2.1.9)), see fig. 2.1.14
+ + 2 = (2 + Z
)
(2 + + 2)
(
x y )
zP2 s tP2 s t nP2
and
(
P1PjP2P2 T
(2 + + + z) (2 + +
x y zP ) (
s tP s t
2)
np ,
J
(2.1.44)
(25 + 2) (2 + Z)
p' p 1,P24T (2.1.45)
tP
which expresses the same as eq. (2.1.43). These conditions are nonlinear; taking
for s the streamline direction (or a good approximation of it), eq. (2.1.45)
reduces to
2-16
4 P1,P2 T (2.1.47)
(o X ) Pl (0
X
)
Eqs. (2.1.143, 146, 47) are sometimes referred to as 'equal pressure' forms of the
Kutta condition. Note that they (pre-)assume that (2.1.39) is already satisfied.
a
(Pa) caS 02 (P2) sin ' (Pi)
as 2
n
a a (2.1.48)
j-;: (P2) sin 02 (P2) cas 02 (P
as
n
at
9(P2)a
j
and
at
at
(P
2
9(P2) J
(c)
and
(E'1) (c)
at ' 1 -: -J
both
(P2) sin 82 0
as
' for p1, p2 T (2.1.52)
and (P1) sin 0 O
--;; J
This can be satisfied only under either one of the following three conditions
17
for Pl, P2 T (2.l.53a)
O
::n (P1)
2-18
ii) sin 02 = 0
(2.1.54b)
sinO1zO ---(P)90 forP1-*T
as
n
iii) sin 02 z O
as
(P2) 40 for P2 - T
n
sin = O (2.1.55c)
In case i) the component of the flow normal to the trailing edge is stagnant on
both upper and lower surface and the flow may leave the trailing edge in a
direction anywhere between that of the lower and upper surface of the wing.
In case ii) it leaves the wing in a direction parallel to the lower surface and
in case iii) in a direction parallel to the upper surface.
It has been shown by Mangler & Smith (ref. 2.14) that which of the three
possibilities occurs is determined by the sign of the vorticity shed at the
trailing edge (or, in other words, by the sign of and
Unfortunately this is generally not known a priori so that an estimate must be
made of the direction in which the flow leaves the trailing edge. In case of
small trailing-edge angles a suitable estimate is the direction of the trailing
edge angle bisector. A requirement of this kind can be expressed as
Further insight into the significance of our mathematical model may be obtained
by considering once more the physics of the real problem, (fig. 1.10). In the
real flow we observe a vortical wake shed from the trailing edge of the body. In
our mathematical model a cross-section approximately perpendicular to the
undisturbed flow direction reveals the intersection of the discontinuity surface
Sc, see fig. 2.1.17. The circulation along a curve t around one of the edges of
S may be written as
B
dt (2.1.55)
= B - =
where A and B are points at the lower and upper side of S, respectively (fig.
2.1.17).
Since z 0, the circulation along t is unequal to zero and we conclude
that there exists vorticity inside the contour t. The only location where this
can be is at the discontinuity surface S, because everywhere else the flow is
irrotational. Hence, the discontiniuty surface is a vortex sheet, with vorticity
along lines of ¡t ji = constant. Apparently, the vortical wake (which is of
viscous origin) in the real flow is represented by the vortex sheet type of
surface of discontinuity in our potential flow model.
At this stage it is worth recalling (see Chapter I) that in stepping down from
the Navier-Stokes equations to more approximate flow models, ranking lower in
the hierarchy of fig. 1.11, essential information on the physics of the complete
flow problem is lost. For inviscid, potential flows such as described by the
Prandtl-Glauert or Laplace equation this concerns, a.o., loss of the ability to
implicitly model rotation (vorticity) and the generation of vorticity.
We may now conclude that we have, explicitly reintroduced the ability to model
rotation (and, through that circulation) through the introduction of the surface
of continuity (or vortex sheet) S and that the Kutta condition now, explicitly,
represents the machenism for generating the correct magnitude of vorticity in
the vortex sheet.
2-20
In general the normal velocity component at the body surface is zero, this is
expressed by
=0
an
onS B
or
U -
an
onS B (2. 1. 56b)
Uo
Furthermore
= o on S
and
The jump on SB (fig. 2.1.18) must be determined in such a way that the
O
0=0 on S o (2.1.59)
a good approximation.
As has been mentioned earlier the Dirichiet condition O = O at infinity. (Sm)
At x * we take
the cross-sectional shape of S invariant in x-direction; it then
becomes a cylindrical surface, usually a plane surface is chosen;
the vortex lines = constant on S parallel to the x-z plane,
resulting in perturbation velocities that have only components in a
plane normal to Sc;
the downstream surface ST normal to Sc;
instead of = 0 (on
(2.1.61)
anhST - - ax
Bodies with sharp trailing edges (wings, tailplanes, pylons, winglets, stub-
wings, flaps etc.) offer a fixed position for flow separation and thus the
beginning of the vortex sheet is known a priori (of course as long as the
Reynolds number is high enough to guarantee attached flow upto the trailing
edge). Examples of such bodies and their vortex sheets are shown in fig. 2.1.20.
For bodies without sharp edges at their downstream end (fuselages, tiptanks,
external stores etc.) the location of flow separation is not known a priori; how
to deal with such cases?
In practice two different approaches are encountered. The first one is based on
the empirical observation that isolated smooth bodies with rounded fore and aft
ends do not generate much lift, at least at small angles of incidence. This
suggests that in such case it is not necessary to adopt in our mathematical
model a jump in the velocity potential at the downstream end of the body, or, in
other words, there is no necessity for a vortex sheet behind the body, see fig.
2.1.21. We then have to solve the boundary value problem with a boundary
2-22
Note that eqs. (2.1.53a, b, c) require that, in general the vortex sheet must
be tangent to the body at
The flow separation occurs along a closed line, fig. 2.1.23, creating a
closed vortex sheet that subdivides the domain behind the body: an interior
and an exterior domain. In this mathematical-topological model the interior
domain Q. has no physical significance because in reality the flow inside Q.
° (2. 1 .62)
and
on ST(3Q.) (2.1.63)
(2 + 2
+ (O + 2)
s t
+
nP np
for P1,Pj,P2,P + T (2. 1 .64)
(2 + + 2) (Z
s t np , s
+
tP
where P1,P2 are on SB at the outer and inner side of S, respectively, and Pj,P
are downstream of S8, at the outer and inner side of S, respectively (fig.
2.1.23). Note that here (P1) = O but P2, P) O
With the zero jump condition (2.1.35) it follows further that
(2.1.65)
(
+ tPl nP2
This mathematical flow model implies that the total pressure of the (fictitious)
potential flow in Q. is equal to the total pressure of the external flow.
A more general formulation is obtained if we assume a jump in total
Pt
pressure across S, according to
on S (2.1.66)
= pt
(2) +
(2. 1 .67)
s
+
tP n P'2 Pt i
Sources
a(Q)
(P)
- 4nr(P,Q)
(2. 1 .68)
-1
= a(Q) VP[4nr(PQ)) (2. 1 .69)
Distribution of sources
For a distribution of sources on a surface S, with the source strength per unit
area given by a(Q), the potential induced at a point P (Q) is evaluated
integrating over the surface S
o(Q)
(P) = If dS (QS) (2.1.70)
4iir(P,Q)
-1
= JI o(Q) V dS (2.1.71)
4nr(P,Q)
S
In eqs. (2.1.70,71) the case P-*Q is to be excluded, since then r-0. This problem
is dealt with as follows.
Consider a circular region X with radius e around Q on the surface S (fig.
2.1.27a) and divide the integral (2.1.70) into two parts: one part over the
surface outside the circular region and the other over the circular region
itself. Now let P-*Q together with c-*0, then eq. (2.1.70) may be written as
2-26
dS
14nr(F,Q')
2n C
f j
-o(p,0)
pdpd0 (2.1.73)
0=0 p=Q L11 Jpz+nZ
where n is the distance along the normal on S from Q to P. For c0, n-*0 and a
continuous source strength o(p,O) this integral vanishes.
Thus the following conclusion may be drawn: the potential of a source
distribution on a surface S is continuous across S and takes the form
a(Q) QS (2.1.74)
'(PS) =
i4nr(P,Q)
dS ;
In analogy with the preceding analysis the normal component of the velocity may
be derived as
a -1
+ 1f o(Q') an 4nr(P,Q')
dl] (2.1.75)
2-27
L
an
P) o(Q')
t'4iir(PQ')
dS - o(P ) ; P,Q' S (2.1.76b)
s
o(P) = L
an
(P) - L
an
(P) =
+
(2.1.77)
It can further be shown that for an arbitrary closed surface there exist a
nonuniform source distribution that, uniquely, induces a constant potential
inside the surface.
Doublets
From the source (or sink) type of elementary solution for Laplace's equation
another elementary solution may be derived by differentiation in a given
direction. Such an elementary solution is known as a doublet or dipole (sources
and sinks are also known as monopoles). It follows, that unlike a source, a
doublet posesses directional properties.
The potential in a point P induced by a doublet in a point Q at a distance r is
given by
where flQ denotes the direction of differentiation and ii(Q) is the doublet
strength.
From this the velocity induced by a doublet may be derived as
Doublet distributions
(P) = fi'p(Q)
a -1 dS
arìQ t4nr(P,Q) (2.1 80)
.
As for the source distribution the potential may be evaluated for the limit P-*Q.
Again we have to distinguish two possibilities, PQ or P and PtQ or P. The
result is found to be
(2. 1 .82)
(2.1.83)
2-30
ENII = (2.1.84)
Note that the jump is given by the negative value of the doublet strength in the
point under consideration. In the case of a source distribution on a surface the
normal velocity component was found to be discontinuous and to jump by an amount
equal to c, see eq. (2.1.76). The minus sign in (2.1.814) is due to the
difference in the differential operator: -p--- =
8flQ
From eqs. (2.1.82) and (2.1.83) we may derive expressions for the tangential
velocity on S, which, consequently is also discontinuous
and
a 1
= i(Q)
¡ (- 1) dS + (2.1.86)
!__ a a 1
[(FeS)) = dS (2. 1 .87)
anP 8n anQ (
4nr
Some particular examples of doublet distributions are given in figs. 2.1.34 and
2.1.35. Fig. 2.1.34 depicts a uniform distribution on a closed body. It may be
shown that inside the body the perturbation potential is constant and equal to
minus the doublet strength per unit area (for an inward normal). Outside the
2-31
Consider again the expression (2.1.81) for the perturbation velocity induced by
a doublet distribution. This may also be written as
-1
= if P(Q) VptIflQVQ (i(P Q))) (2. 1 .88)
S,
It turns out that eq. (2.1.88) can be transformed in a different but equivalent
expression that can be associated with yet another type of elementary solution
of Laplace's equation (see also ref. 2.5).
Making use of the following rules of the differential calculus of vectors
Va(ab) + Vb(a.b)
and
it follows that the essential part of the integrant of (2.1.88) can be written
as
=0
(t)) =
p Q+
=0
+ (1) flQ X
(1
Q P) Q r (2.1.91)
+
2-32
and that =0 =0
'1
X flX v0()J = Q -
-
n +
_=0
;)JflQ
---1
- InQ.Vp)VQ(_] (2.1.92)
(_l))
= - px [iQX (2.1.93)
V1D = x E
JI
4
)X fl dS - II (Q(P) X flu) dSJ (2.1.96)
If (iixA) dS = f A ds (2.1.97)
S as
which is related to Stokes theorem. (aS representing the curve bounding the
It gives
=
X [ f
as
-+
ds
4nr
[QPxQ)ds] (2. 1.98)
2-33
=
rxds 1
- p(Q) + Jf [V (p)xn ) x dS (2. 1. 100)
S r S r
In the first term of (2.1.100) we recognize Blot & Savart's law (ref. 2.1)
applied to a closed line vortex bounding the surface S with strength equal to
the local doublet strength at aS. In case of an unbounded surface this term
vanishes. In the second integral we recognize Blot & Savart's law applied to a
distribution of vorticity on the surface S with the vorticity 'r equal to
:;
= VQ(P) XflQ (2.1.101)
=0
= (px) = cp) = O (2.1.102)
ff:;(Q)
(P) = x dS (2.1.103)
The latter two equations involve that the tangential velocity across a vortex
distribution on a surface is discontinuous, the jump in the velocity is equal to
the vortex strength T.
Due to the inherent singular behaviour of sources, sinks, doublets and vortices,
they are also referred to as singularities and their distributions as
singularity distributions.
In the next section we will see how solutions of' the boundary value problems
discussed in paragraph 2.1.2 can be expressed in terms of singularity
distributions.
1ff
Q
s
-
s
¿) dQ = Jf (
S
(i') - s
)J dS (2.1.106)
i
(2.1.107)
s - 4ir'(P,Q)
This means that satisfies Laplace's equation everywhere inside Q, except for
the point Q, where D becomes singular. We now apply Green's theorem to the
domain Q while excluding a small spherical region w with surface and radius c
Then we have
s
- s ] dQ = - ii r an
- a
an
1
dS +
Qw
+ L 1
r3n
-
anr1)) dz (2.1. 108)
ITTlz
hm
c-o
If r an
dz = C (2.1. 109)
him
° _
L jj
a
(- ) 1
dz] =- (P) (2.1. 110)
z
= L
4n
¡j
r an - an
(_
1))
r
dS (2. 1. 111)
-
ff((
S
- 1
ran an (- )) dS = O (2.1. 112)
2-37
(PeQ) = 0 (2.1.113)
We already know that for a well-posed boundary value problem (either inside or
outside Q) we must prescribe on the boundary S of Q either the potential
(Dirichlet problem) or its normal derivative (Neumann problem). Apparently
Green's identity describes a wider class of problems than the boundary value
problems in which we are interested. The difference is that Green's identity,
describes the solution both inside and outside Q, whereas the boundary value
problem is formulated either for the interior region or for the exterior region.
As we have pointed out earlier, the boundary value problems we are confronted
with are of the mixed type: we seek the solution of a boundary value problem in
a domain Q bounded by a closed surface S, where on a part of S the potential
is prescibed and on another part the normal derivative . In this section we
analyse how we can use Green's identity to find the solution of such boundary
value problems. Considering eq. (2.1.111) it is clear that prescribing only on
S (interior Dirichet problem) is insufficient to determine in a point P inside
Q, because is still unknown on S. Similarly a Neumann problem cannot be
solved because also has to be known on S.
To solve either two problems an additional and important step has to be made.
For this purpose we consider Green's identity for the case that the (internal)
point PEQ approaches the boundary S, see fig. 2.1i2. Then Green's identity
gives in the limit
i a
(EeS) = H- ) - r
dS, qs (2.1.1114)
2-38
The factor 1/2 arises from the fact that now around P a half-sphere ci of radius
-o has to be applied; the intersection of the half-sphere with the boundary S
1 1 3
- dS = (P) + '(Q) -p--
3flQ r
dS (P,QES) (2.1.115)
The left hand side of eq. (2.1.115) may be recognized as the right hand side of
eq. (2.1.74) which was the potential on a surface S produced by a source distri-
bution with strength a on S. In the case of eq. (2.1.115) we have a =
Q
Eq. (2.1.115) is known as a Fredhoim integral equation of the first kind. (The
(general) theory of integral equations may be found in ref. 2.7).
In general the numerical treatment of Fredholm integral equations of the first
kind causes difficulties and should be avoided (see ref. 2.8); we return to this
later.
Now consider the problem where is given on S (Neumann problem). This leads to
-' on that same surface S and with the dipole axis pointed inwards. We know that
a uniform doublet distribution on a closed surface induces a zero potential
inside that surface (see section 2.1.6), resulting in a zero (normal) velocity
on the inner side of the surface. This implies that we may superimpose any
constant on without changing the righthand side of eq. (2.1.116), thus
reflecting the non-uniqueness.
The considerations just given show that a well-posed problem does not automati-
cally lead to the preferred type of integral equation (Fredholm integral
equation of the second kind) and also that a, at first sight, suitable integral
equation (of the second kind) does not necessarily have a unique solution.
Once more we consider Green's identity, but now with scalar functions and
defined for the interior region Q and scalar functions " and ' for the
exterior region Q' of a closed surface S, see fig. 2.1.43.
If satisfies Laplace's equation inside Q and ' satisfies Laplace's equation
outside Q, i.e. inside Q', the potential in a point P in Q may be expressed
as, see eq. (2.1.111)
(p)=Lff{
1- -r l)3a(
3n 3n
1))ds
r
(2.1. 117a)
o=fJ[(-1
S
r an'
a
3n' r
(2. 1. 117b)
2-40
Here n' is the unit normal on S pointing into Q'. If ' should be a solution of
a well-posed boundary value problem in Q' we formally have to bound Q' by a
second closed surface S' and apply the integral (2.1.117b) also to this surface.
In doing so we will assume that S' is infinitely far from S and that ''-*o for
S'4w; in that case the integral over S' can be shown to vanish, see ref. 2.2.
Then, superposition of eqs. (2.1.117a) ai-id (2.1.117b) yields, after replacing
a a
j-- by - j-
i a a' a 1
(P) = N- ) (j-h - j---) - - 3n
;fl dS (2. 1. 118)
The first part of the integral represents the potential due to a source distri-
bution on S with strength o
a
j-- -
a' and the second part represents the
=
potential produced by a doublet distribution on S with strength p = -( -
(P) = L jj [(- -)
r
i
+ p an- (-iS
a
(2.1.119)
4TIS
In analogy with the preceding analysis we may express the potential ' in a
-
1
i- J! ( 1 a'
(j--
a
- j-h-;-) - (' - aii'
dS (2.1. 120)
S
or
a
=
[G (_ r an
-fldS (2. 1. 121)
The integral representation (Green's identity) (2.1.118) for the potential (P)
In order to evaluate (P) from eq. (2.1.119) for a point P on S we repeat the
limiting proces that was also applied in obtaining eqs. (2.1.74) and (2.1.82).
The result is
1
(P) = (a(Q) (- -) + j.i(Q) dS - p(P), for PS = aQ
arlQ
S
(2. 1. 122)
where we have taken the + side (or Q-side) of the surface S. If (PS) is
prescribed (Dirichlet condition) eq. (2.1.122) can be interpreted as an integral
equation for o and p. However, the integral equation will not have a unique
solution as long as either o, or p or a relation between both is not prescribed.
The situation suggests that prescribing either a, or p or a relation between
both is equivalent to specifying ', about which later.
If (PS) is prescribed (Neumann condition) we may obtain an integral equation
for a and p by considering the integral representation for and by letting
N 1))
i
an
(P) =[o(Q) a
an (- -)
r
+ i-i(Q) -q---
j dS + o(P)
For the saine reason as before the solution for o and p is not unique. In this
case it is not unique even for prescribed values of o or p, because of the ill-
posedness of the Neumann problem for an interior domain.
For the mixed boundary value problem a system of integral equations may be
obtained that has a unique solution provided that either o or p or a relation
between both is prescribed.
2-42
Similar to the analysis that led to the expressions 2.1.122/123 for (P) and
expressions may be obtained for ' (P') and (P') for a point P' at
np
the Q' side of S (aQ'),
=
i
[o(Q') (- ) + i.i(Q'
anQ? (
1j
- r1
S
P'aQ' (2.1.124)
and
i
an, - 4n [(Qe) 8n
8 1
(-)+p(Q')
r
a
ann, 8nQ
a
))dS-'
r
G(P')
P'caQ' (2.1.125)
These expressions may be used for formulating a boundary value problem in the
exterior domain Q' (see fig. 2.1.43).
Conversely the doublet strength p and the potential ' (P) on S' can be
determined by solving a boundary value problem for ' in Q', i.e. by solving an
2-43
In other words: for any boundary value problem in Q there exists, for any choice
of o or p, an equivalent boundary value problem for ' in Q'. Which equivalent
boundary value problem is concerned precisely, is determined by the jump
conditions associated with the particular choice of o and p.
In principle the number of possibilities for fixing o or p (and through that ')
is infinitely large. The following possibilities are the most convenient ones
and indeed used most often:
p=O, or =' on S
In case of a Neumann boundary value problem for eq. (2.1.123) then leads
to a Fredholm integral equation of the 2 kind for o.
o0, or
a'
an
=onSa
an
In case of a Dirichiet boundary value problem for ' eq. (2.1.122) leads to
2nd
a Fredholm integral equation of the kind for p.
0=-
an
(2.1. 126)
and
u=- (2.1.127)
In case of a Neumann boundary condition on aQ for one may use (2.1.126) to fix
Then (2.1.124), with '=O as boundary condition for the (equivalent) boundary
nd
value problem in Q , leads to a Fredholm integral equation of the 2 kind for
In case of a Dirichlet boundary condition on aQ for one may use (2.1.127).
2-44
a.
boundary value problem for ' (as implied by iii)) is often used in practice.
Example: a Neumann problem in Q as shown in fig. 2.1.44. The source strength is
determined by eq. (2.1.126). This problem can be replaced by the Dirichiet
problem '=O in Q'.
It may be interesting to use this possibility in order to obtain a Fredholm
integral equation of the second kind and/or to reduce the computational effort
(see section 2.1.13).
As an example consider the Neumann problem for the exterior domain of a closed
surface S with a net mass flow through S, given by (fig. 2.1.45)
dS xO
an
dS = O
we run into problems, because we can add a constant to any solution i for the
doublet strength which does not affect the normal derivative (the potential
of a constant doublet distribution on a closed surface is zero outside the
surface *), section 2.1.6). If we consider this problem from a different
viewpoint we arrive at the same conclusion: with only a doublet distribution on
*) Note that because of this the external Dirichlet problem cannot be solved
either with only a doublet distribution!
2-145
a
S the normal derivative - is continuous across S therefore the equivalent
an
interior problem is a Neumann problem of which we know that it does not have a
unique solution.
B
r= f asdsxO
A
(s is the distance measured along the contour of integration, see fig. 2.1.46).
This problem cannot be modeled by only a source distribution on the surface S,
because the circulation around a single source or around a source distribution
is zero. In order to be able to do so a non-zero circulation a doublet
distribution or a vortex distribution is necessary on SB and S (fig. 2.1.46).
These examples illustrate that a sufficient wide class of boundary value
problems can only be modeled by singularity distributions consisting of both,
sources arid doublets.
- Mixed problem: In general well-posed for both interior and exterior domains.
2-'46
Integral representations
- For a given boundary value problem inside S and a given source strength or
doublet strength an equivalent boundary value problem outside S can be
formulated.
Integral equations
Consider the boundary value problem for the flow around a configuration as shown
in fig. 2.1.47 (depicted earlier in fig. 2.1.18).
In this section we will analyse how such a problem may be transferred into a
system of integral equations. As we have seen (in the preceding sections), the
solution of a boundary value problem can be represented by means of a source
distribution and a doublet distribution on the boundaries S of the domain Q
under consideration. Hence, the perturbation potential in a point P of Q can,
according to eq. (2.1.119), be expressed as
1 1 a
P) ((_ -) ' dS (2.1.128)
BCT
+ _
r r
S +S + +
The various parts of the boundary surface, SB$SC,S and ST, distinguished in
section 2.1.3, have been indicated in fig. 2.1.47.
Considering first the contribution of the vortex sheet we learn immediately
that the source strength on S is equal to zero because of the boundary condi-
tion = O. In addition the doublet strength is directly connected to the
jump of the potential across SC, because i=- 1.
Therefore the contribution of S to the integral is
a
p (-
1
;) dS=--ff]--(-1)dS (2. 1. 129)
tR
1 a
ISB( f ) ds) dt (2.1.130)
SB
Taking ' = O in the domain Q' outside S and ST we have, according to eqs.
(2.1.126) and (2.1.127),
p = - = O, and o = - on S
an
and
o = - = O, and p = - on ST
' (FS,) =
TTS
pp
an (_ rJ
dS + p(P) + fi J dS
T
=0
,. if as + if pj-
a 1j dS (2.1.131)
- r
S S
C
a lj
(P) dS (2.1. 132)
r
T
1
((
1)
r
+ an
p a
r
dS - L jj ['t'i (- )
dS, pQ
4TTS
(2.1. 133)
PSB)
= L [(_ r) +
i a
anQ
(_1))
r
as - 3i(P) - ¡f (- )
dS
4SB
(2.1. 134)
In that case the expression for the normal derivative of ' takes the form
= L [a
3n
3
i- ) +
i a
i
a
r )
) dS +
2
o(P) +
S
a
-
3flQ
1-)dS=-U
r
n
- (2.1. 135)
= = (_ U ) (1) dS +
1'a1Q
(.1) dS +
nd
Note that this is (also) a Fredhoim integral equation of the 2 kind (provided
is known).
C
A further important point to notice is, that in case of (2.1.136) the
perturbation potential on SB follows directly from the jump relation
Q - Q' = -i-'
(2.1.137)
Q(PSB) = -
With respect to the Kutta condition it has been pointed out in section 2.1.3
that, in the first place, this requires continuity of the (perturbation)
potential across the trailing edge of the body. This can now be stated as (see
fig. 2.1.48 and recall that [[]1 -L')
G(P)
for P1,F2,P,P + T
+ Q(P) = G(P) +
1 111a 1 a
+ 1J: -
a
(-
1
) dS +
anPPT = an
SB
an
a
(
-j dS = - (U. "PP+T' lDC (2.1.140)
2 - 2
[ 5
(u x+')) - [V
5
(u x+)) -* O for P1,P2+T (2.1.141)
2
(P) = L [a(- r)
+
anQr2
a
li-
4SB
1 a
_fI3flQ C
1)dS
r
(2.1. 142)
Panel methods are based on the numerical threatment of integral equations of the
type given by eq. (2.1.135) or eq. (2.1.136); in both cases the system is
completed by the additional relations (2.1.139) plus, in case of (2.1.135), by
(2.1.140) or (2.1.141).
Since there are many possibilities for choosing the singularity distributions a
and i, together with different possibilities for the formulation of the Kutta
condition and different numerical treatments, a considerable number of different
panel methods have been developed over the years. The most important of these
are summarized in Table 2.1. The methods mentioned in the table are all
applicable to general three-dimensional configurations.
There also exists a large number of methods for the computation of potential
flows around more restrictive geometries (two-dimensional, axi-symmetric, thin
wing configurations etc.). These methods also differ not only in their basic
formulation but also in the numerical treatment, particular applicability etc.,
they will not be discussed here.
2-53
In the preceeding chapter we have seen that integral equations for boundary
value problems representing the flow about aircraft configurations can be
obtained by either applying an external Neumann boundary condition or by
applying an internal Dirichiet boundary condition.
The integral equation based on the external Neumann condition is given by eq.
(2.1.135). For a point P on the (external) surface SB of a configuration it
reads
a i a l
r(P,Q)1 dS +
i o(P) +
= o(0) -
a a i )dS+fJp(Q) a a i dS
(
an aflQ r(P,Q) n an anQ r(P,Q)
s
=-U ; (2.2.la)
= f(s) [p(0)J5
p(Q)115 = f(s)p (t)
B B
[u(0)) =p (t) (2.2. lb)
C SB
where p (t) is the doublet strength at the trailing-edge (see fig. 2.2.1).
B
The system of equations is closed by adding as 'explicit' Kutta condition either
the 'directional' form (2.l.l1tO)
2-56
30 1 a -1 a a -1
- f! o(0) Lr(P,Q)) dS +
r(P,Q )) dS
-
an 3flQ
SB
for PS -s
+ p
an0 r(P,Q)
dS = - U xi
P CB (2.2.lc)
1 a -1
= ° dS + 1_l(Q) dS +
r(P',Q) anQ r(P',Q)
SB SB
[a(Q))5 = - (U
B Q)sB
(2.2.2b)
[p(Q))5 = 1_ls
(t)
C B
p
(2.2.2c)
= p2) -i(P1) + = sB(t) for P1,P2,P' + T OrI SB
Substituting eq. (2.2.lb) into eq. (2.2.la) and rearranging the result we obtain
for a point P on SB
2-57
Lc a(0)
a
an (
r(P,Q)
dS + o(P)
SB
+ L jj
4T1
f(SQ)
a
(t0) -;:;;
an0
a -1
r(P,Q)
dS +
SB
SB
tR
a a 1
[ ds(Q)] dt0= - U (2.2.3)
an anQ r(P,Q)
tL
Analogously we may derive by substituting eq. (2.2.2b) into eq. (2.2.2a) for P
on SB
a -1
dS + p(P) +
an0 r(P,Q)
s
B
t
R
a i
)[ f anQ L r(P,Q)1 ds(Q)] dt0 =
B0 sß(t)
L (U J
r -1
dS (2.2.L)
p r(P,Q)
SB
-1
potential of a source
r(P,Q)
a i
L
potential of a doublet
r(P,Q)
(2.2.5)
a i
normal velocity component due to
an '- r(P,Q)
a source
r -1
normal velocity component due to
an -r(P,Q)
a doublet
In panel methods the integral equations (2.2.3) and (2.2.4) are discretised
through the following steps.
2-58
First the surface of the configuration and its vortex sheet are subdivided in
a number of elements or panels, see fig. 2.2.2. In most cases it suffices to
approximate the geometry of the vortex sheet by a rather simple geometry on
which the doublet strength is assumed to be constant in streaniwise direction.
Therefore, the panel distribution on the vortex sheet can be limited to only
one or a few panels in the streaxnwise direction.
The integrals in eqs. (2.2.3) and (2.2.14) are replaced by summations of the
integrals over the individual panels. For example
dS
r(P.,Q)
SB i
is replaced by
NB
'
Z f! a.(Q) (- r(P.,Q) J dS + o(Q) r(P.
dS (2.2.6)
n i
j=1 ÓS
jB i
iB i,Q)
ixi
and
dS
3flQ r(P.,Q)
is replaced by
NB
L
4n
jj (Q) aÇ (- r(P1Q)
dS +
j=l ô.S
ii
L a
aflQ r(P
-1
,Q)) dS (2.2.7)
4T1ÔS i
iB
Here ó.S and ä.S denote the surfaces of individual panels with index i
iB
and j, respectively. NB and N are the total number of panels on SB and S,
respectively. Note that because of the fact that we are dealing with
2-59
principle value integrals the integrals over panels with index j=i must be
treated separately.
Because of the fact that, on Sc, 1.1 is constant in the streainwise direction
it is convenient to rewrite the integral over S as
tn
1
SQ [f anQ
1
tr(P.,Q)1 ds(Q)] dtQ
sB(t)
and replace it by
Nc
1
z f ds(Q)] dtQ (2.2.8)
k=l äkt
ij
530 aflQ r(P.,Q)
sB(t)
dS
dd 2 3P)2]1/2
= [i + +
an
ddn (2.2.10)
where n is the Ç-component of the normal in a point on the panel, see fig.
2.2.3.
2-60
00 00
- °j-1
= ( = 0, n = 0). + 0(h2)
j+1 -1
etc.
c. Now the integral over a panel can be worked out in its local coordinate
system by substitution of eqs. (2.2.10) to (2.2.12) into eqs. (2.2.6) etc.
As an example we carry this out for the integral of eq. (2.2.6)
-1
JI o.(Q) ,Q)) dS
r(P
os
jB i
-1 dÇdn
II (o°° + 010 +
os j J {()2 + (n.-n)2 + (.)2)h/2 InI
j B
where
2-61
If
- 1 ddn
6jBS (()o + (-n)2 + (çç)2)'/2
A°.°. =
lnI
ddn
()Z
-
= JI (2.2.14)
'J ô.S (R_2 i
+
iP )2)/
(Ç.Ç InI
ddi
A°.'.
'3
=
6 S [z + ()2 +
i
f! -
(ç -ç )Z)l/2 Inçi
NB
a(- 1) dS = (A°° 000 + A1° + A°' 0.1 + (2.2.15)
)
4ii r ii j ii j
SB i =1
N
L0(
1-hi r
) dS
B
(A.. 00.0
13 3
+ ) (2.2.16)
S5
10
where A. . + 10
13
= A°.°.
13
+ A'.°. b.
13 3 i,j+1 j+l a. + A.
i,j-1 c.j-1
. +
It is possible to derive similar AIC's for all the terms of eqs. (2.2.3)
and (2.2.4). Consider for example the perturbation velocity-components on
the i-th panel due to the source distribution on the j-th panel. The
2-62
(!_)
a.
-
f!
3B
a(Q)
if i
r(P., Q) ) dS = (A i.. o.
13 3
i
a(Q) dS = (A ).. o.
ar = r(P., Q) )
T1J 3
i
'
o.s i i
Since
3n. j .
j
= n.
-
i
,-
(V
P.
j .
3
we may write
. ,a
j.
an. j
= n.
-
i
-
A.
13
. a., where A.
3
-
ij
=
i.
[(A
(a
j
ij.,
. (A
T)ij
) .,
.
(A j . .J.
Çij
Writing further that A
13
. =
1
A .
13
we arrive at
j . = A a..
an.j
.
13 3
NB
a(Q) dS + a(P.) E A' a + (Truncation Error)
r(P.,Q) ii
i j=i
i
(2.2.17)
Nc
f(SQ)
aÇ k - r(P.,Q)
i
aS =
k=l
E B'
ik (sB)k + T.E.
i
(2.2.18)
tJ: Nc
1
a a
r(P ,Q)i dSQ] dt =
I C' T.E.
(tQ
sB(t)
an 3flQ
i k=1 ik'sBk
i
(2.2.19)
NB
dS + B p + T.E. (2.2.20)
p(Q) -p---
an _ r(P
Q)
ti(P) = I
Q i' i=1
2-63
NB
flQ) H r(P.,Q) dS A.. (U + T.E. (2.2.21)
= 3-
B
tR w Nc
1
SQ sB(t)
3flQ r(P
i
Q)) dsQ dt
k=1
c
ik SßK )
+ T.E.
(2.2.22)
5. If the integrals in eqs. (2.2.3) and (2.2.4) are replaced by the summations
given in eqs. (2.2.17) to (2.2.22) we obtain for eq. (2.2.3)
NB N0
X A . a. + I (B:
ik
+ c:ik ) (p 1 =- U n.
1
,i = 1,2,3, . . . N
B
SB 1(
k=1
(2.2.23)
N N N
B c B
X B..1.i.+ I C.(p
ik J
SB k
= I A..(U
w ).,i=l23.....NB
k=l j=1
(2.2.24)
or
'r a. = _(Uw.i:)j
[A: .1
j
(2.2.25)
2-64
[B..] p. = -[A..]
13 3 13
(U ).J
m
(2.2.26)
where [A'..] and [B..] are square matrices of which the elements are the
aerodynamic influence coefficients, o. and p. are the vectors of unknown
source and doublet densities to be solved for.
resemblance of eq. (2.2.lc) and eq. (2.2.la). Then, in matrix notation the
complete system of algebraic equations becomes
A. Bk + o.
J
i,j = 1,2,3 . .
. NB
Here the aerodynamic influence coefficients A'13., A'QJ describe the influences
.
In the case of the 'internal Dirichiet' formulation eq. (2.2.24) and the
pi * U ij = 1,2 . .
.
j
Note that the number of unknowns remains NB and that the right-hand side of
(2.2.28) is identical to that of (2.2.26).
The linear algebraic system (2.2.27) or (2.2.28) are solved using numerical
techniques that are either standard or are developed specifically for this
purpose (see section 2.2.9).
When the solutions o. and p. are known, the potential and its derivatives can
be computed throughout the entire flow field including the body surface SB.
In doing so one can, again, make use of aerodynamic influence coefficients.
For example for an arbitrary point P one may write
= if o(Q) J dS + L jj a
dS
[
r(P ,Q) 4ii r(P,Q)
ni
SB SB+SC
= A a. + X B p (2.2.29)
j
.
1113 3
k
mkk
a(P m) 8A . aB
mk
u(P
ni
)
ax 3x ax
.
k
a(P m) aA . aB
mk
v(P ) = (2.2.30)
ni ay ay j ay
k
a(P ) aA . aB
mk
w(P)= in az az G +X az
- k
2-66
=- 1.1
NB
K
x
= - f! p n
S
x
dS = -
.
i=l
Xli
(p n ) .ó.S + (2.2.31)
From the velocity components and pressure distributions one can also construct
streamlines, isobars, etc.
These geometry packages offer the possibility to describe surface parts (called
patches, fig. 2.2.4) analytically, either in terms of elementary geometrical
shapes such as plane surfaces, cylinders, spherical surfaces, ellipsoids, etc.
or in terms of more general surfaces described by polynomials (e.g. splines).
Once the analytical description of a patch is established, the coordinates of an
arbitrary point on it can be determined easily.
Panel shapes
In principle there are many options for panel shapes, e.g. triangular, qua-
drilateral (trapezoidal), pentagonal, etc. Triangular elements (fig. 2.2.5),
applied frequently in finite element methods, have the advantage that they are
easily patched together in irregular shapes. Also a continuous representation of
a surface is always possible (no gaps). The disadvantages are a rather
complicated administration of panels (i.e. indexing) and a (more) irregular
distribution of nodal or collocation points which is undesirable for aerodynamic
applications. In aerodynamic practice almost exclusively quadrilateral elements
or panels are used (fig. 2.2.6). The main reasons are: they generally fit
reasonably well into regular aerodynamic shapes such as wings, fuselages, etc.;
they can be administrated by a simple, cartesian indexing system and it is
always possible to construct a triangular panel as a special case of a
quadrilateral. A disadvantage is that not all surfaces can be represented
smoothly without gaps.
F(x ,y ,z ) = 0 (2.2.32)
X = x (s,t) i
= y (s,t) (2.2.33)
z = z (s,t)
p p j
or
= (s,t)
where X is the position vector of a point on the surface, see fig. 2.2.14,
(s,t) is a curvi-linear surface coordinate system and f(s,t) is a polynomial in
s and t.
The relations (2.2.32) or (2.2.33) describing a surface patch may be obtained as
follows:
directly from a general geometry software package of the type mentioned in
the preceeding section.
from a separate, local surface fit (polynomial) through the corner points
of a panel (and, if necessary, through those of adjacent panels).
The first option has the advantage that use is made of knowledge that is already
available. The second one may be benificial from an informatics point of view
since the part of the computercode describing the gemetry as it is required for
the aerodynamic computations is decoupled from the (complex) general geometry
software package.
Depending on the degree of the polynomial representation there are several
possibilities for the local representation of the geometry of a panel.
In figs. 2.2.15 and 2.2.16 representations are considered for planar and
quadratic panels. It is important to notice that planar representations do not
need information of neighbouring panels, whereas for the higher order repre-
sentations such information is needed, complicating matters significantly.
Once the local description of the geometry of a panel in the body reference
system is known the local coordinate system (,n,Ç) of the panel can be chosen.
In general this is done in such a way that the (,ri)-p1ane is tangent to the
panel in its 'centre' see fig. 2.2.17. According to the calculus of vectors the
2-70
a a?
n= (2.2.34)
a a
How, exactly, the 'centre' of a panel is defined is not of great importance. For
planar panels one usually takes the centre of gravity. For paraineterized panels
the centre of the rectangular panel in the parameter space is used.
The 'centre' of the panel acts also as origin of the local coordinate system
(,ri,Ç) and as the point where the boundary condition is satisfied.
The ,ri directions are usually chosen such that the ri-axis is normal to the x-
axis of the body reference system. The t-axis then follows automatically if a
righthanded coordinate system (,ri,Ç) is required. With these choices the
representation (2.2.9) of the surface in the local panel coordinate system
reduces to
(P) i
c(n) -
- 2!
_
2
2
ta3no fl + n2 + 0(h3) (2.2.36)
The second derivatives are determined in the following way, see fig. 2.2.18. As
indicated are unit vectors along the x,y,z (reference) axes, respectively;
are the unit vectors along the ,n,ç (local) axes, respectively, i is the
unit normal in the centre of a panel. The ri-axis is taken normal to the x-axis,
thus I=ü. If the vector locates the panel centre (point of expansion in eq.
çP = r1 (2.2.37)
TI = - o
ç = - x0).
it can, using differentiation rules and the calculus of vectors, be shown that
a2ç P 2 2 a2p
+ 2 n (2.2.38)
2 2 asat +
as at2
o
as 3t
The coefficients , - can be expressed as
a a
a
as 1 an
m (2.2.39a)
D at - D at
l3r _l ax P (2.2.39b)
a Das Das
ap.
as at as at
D= (2.2.40)
an
as at
The scalar products in eq. (2.2.40) can be written in terms of the components of
Q,m in the x,y an z directions as
ax1 az
£ Q
as = as + as + as
ax az
m
as
- -m
as x
+ as my + m as z
(2.2.41)
etc.
2-72
ax1, ax a2x
The partial derivatives ... etc. can be approximated by
at ' as ' as2
finite differences, for example central differencing
-
(P) (2.2.42)
as i,j - 2s
Using eqs. (2.2.36) to (2.2.42) has the advantage that the only information
that needs to be stored in the computer memory are the coordinates of the panel
centre. If polynomials like eqs. (2.2.32) or (2.2.33) are used throughout, for
each panel also the polynomial coefficients have to be stored, which amounts at
least to twice as many numbers.
In section 2.2.1 we have seen that the aerodynamic influence coefficients are
computed using polynomial-type representations in the panel coordinates ,n of
(2.2.46)
pR,n) = p(O,0) + -(0,0) + (0,0) n + 0(h2)
2-73
ao as aa at ao
(2.2.147)
a a as a at
8o
an
= as
--ao + at--
an as
ao
an at
(2.2.48)
a
as
- D at
(2 .2 .49)
a
at
fi
a - - D as
and
as_ 1ap
anDat Q
(2.2.50)
a
at
anDas
Q. . - o.
1+1,3 i-1,j
- 2s (2.2.51)
etc.
The same procedure may be followed for the doublet distribution, ji.
2-74
With the representation of the geometry and the singularity distribution for a
panel established, the next step is the evaluation of the integrals appearing in
the integral equations, in other words, the evaluation of the Aerodynamic
Influence Coefficients (AIC's) introduced in section 2.2.1.
= L -r-
o.jr.. dS = A. .o. (2.2.52)
ji ijj
Thus, if the point is the collocation (mid) point of the panel i, A..o. is
A.. =- L jj [()2 + +
ddn (2.2.53)
'j 411ôs
jB
I.
A.
13
.
= L F1
(k+1)
- F1
(k)
)
+
-
k+l)F2 - ki_k)F2(k) +
k=1
(k+1)
+B o (k+1) F - B' F31], k+1 : = 1 if k=4 (2.2.54)
2-75
where
F1
(k) ik ik -B1
i
+
1k +
ik
i
-fl k
F2' - sinh1
ikj2 + ç2Jh/2
1 ik + B (k) ( fi)
F3 =
[1 + (B1)2J*2 sinh
(k)2 (B1)2) ç2] 1/2
[(B +
o
B' = ik - B1
k
73
B=R
1
-
v.
k+1 -
-
r
Ó.. = - f! o.v(---) dS
ll
dS = A. .o. (2.2.55)
j i 14n j r.. j
jB jB
where
= k.-) + (T1.-fiJrn +
(2.2.56)
rl = r = [k1-) +
()2 + ç]l/2
J'
2-76
Hence
-- (A. .)
1J
= (A j
13 = ddn (a)
ôiS
n . -n
= ff ddn (b) (2.2.57)
(A
f13
) . .
4TTÓS
jB
ci
(A) JI -- ddn (c)
ç ij -
The integrals in eq. (2.2.57) can also be evaluated analytically. The result is
r +r -d
21
n -n
log
r +r -d
(1212) 22
-n
log
2 3 23)
- d r1i-r2+d + d r2+r3+d23
12 12 23
n4-n3
log
3434 log (2.2.58)
+ d34 r3+r4+d34 d41 r4+r1+d41
r+r-d r+r-d
4nA
12 log
(12 12) 23 log
2 23)
3
n d r1+r2+d + d r2+r3+d23
12 12 23
r4+r1-d41
log
3434 +
4l log (2.2.59)
+ d34 r3+r4+d34 d41 r4+r1+d41
m e -h 1 m e
12 2
-h 2)
-1 12 1
nA=tan tan
cr
il i2
m
23 2
e -h 2) m e -h 3)
23 3
+ tan1 - tan
çr
12 i3
m e -h 34e4 - h4
3 3) - tan1
+ tan' i3 Ç.r4
m41e4 - h4 m41e1 - h
(2.2. 60)
+ tan' j- tan' ç.r
ji4 il
with
2-77
d12 = [k 2 - + (fl2 -
1
n2 - n1 n3 - n2
m12 m23
= - - 3 -
(2.2.62)
n14 - n3 n1 - n4
ffl314 a23
- - - -
and
e
k = Ç2.
i +k - k = 1,2,3,4 (2.2.64)
h
k
= (n.i - n k ) k.i
k = 1,2,3,4 (2.2.65)
ç. o. o.
(A ).. o.
i
= ii JI o -
r
dS ± -
2
= ±
2
(2.2.66)
iB
(iii) eq. (2.2.60) expresses the spatial angle under which the panel j is seen
from point i.
In the case of Neumann boundary conditions we must know the normal velocity
component in the collocation point of panel i. This component is obtained from
2-78
o.(
3 fi). =
i J i =
1
. (VA.
13
.)a.
J
(2.2.67)
where
VA..
13
= [(A )..,
ij
).., (AciJ
(Anij )..) (2.2.68)
(2. 2. 69)
ô.4. = -
3 1
ff
o.s aç r.. dS = B..
13 3
jB
where
ddn
B.. = - (2.2.70)
13
[k-J + (n.-
+
The expression (2.2.70) is, apart from the minus sign, identical to eq.
(2.2.57c), thus
B.. = - (A
13 j..
1J
(2.2.71)
ó..
j i
= - 1f p.
j
pÇ (i--)
r..
dS = (
B..) p.
ij j
(2.2.72)
jB
i 3(.-)Ç.ddrì
.---- (B = -1f (2.2.73a)
= .)z
óS i
+
(
i T))2 +
3(n
(B)
liii
=_L j r5
ddn (2.2.73b)
jB
3ç
(Bc).. = -
r
ddn (2.2.73c)
ô .s
jB
'r x XdS
j i
= 1f ' ddri -
-
j (2.2.74)
ôiSB kI aô.S Ii:.13
= g + gin
as = gds
and
gds = ds =
gds ari
ds = dri,
2-80
(- c)
= L jj ddn + pg ds (2.2.75a)
,J
i LIuós
jB
nr3 aó.S
jB
r3
o.'
j n.
=- J!
os T!
'r
(- ç.)
r' ddn
i
ao.s
r3
(2.2.75b)
jB 3B
n.-n
o.) = - 'r ) ddn +
.
' o'
3B
-(n-n)
( 1 1
(2. 2. 75c)
pg + pg j ds,
4n r3 r3
ao s
jB
-
'r = O. Hence only the line integrals in eq. (2.2.75) are retained. In that case
the velocity components may be written as
= (s°).. p
(2. 2. 76a)
= (°).. T' ij .
j
(2.2.7 6b)
(2.2.76c)
3cl
= ijj
(B°)..
ç
g ç
(B°)..
T' ds -
r3 - 4ii g; r3
(2.2.77a)
13
aó aó s
jB
jB
2-81
r
ds= r g 13
(2.2.77b)
36 36 rL
jB JB
g -(n.-n) g
(B°).. -
ç jj - 4ii
L
r3
ds
+ -; L
r3
ds
36 S
jB ao .S
jB
_L
-
g
ÓS
L
n.-n
r3 LIn g
L
r
d (2.2.77c)
JB jB
The integrals in eq. (2.2.77) can be evaluated analytically, giving
LI r i-r +-
k k+1 k+1 'k
adS
L
r3
d =
k=l
X Qn
rr
k k+1
-
kk+1-k)
(2.2.78)
jB
rr + -
kl k +.1
k k+1 'k+1 'k
L -d
r3
X {r r in
rr (2.2.79)
k=l k k+1
jB
+
rk = ik)2 + (2.2.80)
Multi-pole expansions
As we have seen in the preceeding paragraph the aerodynamic influence coef fi-
dents of constant-strength source and doublet distributions on planar panels
can be determined in an exact way by evaluating the integrals analytically.
However, their computation is costly due to the large amount of coefficients to
be determined (N*N, N=0(103)). In addition, these coefficients contain a large
number of logarithms and inverse tangents.
The computing time can be reduced considerably by using, instead of the exact
expressions, series-expansions of
1
r
= 1( _2 + (._)2
i + (ç1_çJ2]*2 (2.2.81)
in terms of
2-82
L
r
= [2
i
+ ri2.
i.
+
ii (2.2.82)
o
i
iL{'+-
r r
o
r2
o
+
112 +
o
r
o
(2.2.83)
where h is a measure for the panel dimension (ref. 2.10). The first term at the
righthand side of eq. (2.2.83) represents the potential of an elementary point
source (monopole), the second and third term that of a point-doublet with axis
in c-direction and Ti-direction, and with strengths and n, respectively. The
terms of o(P) represent quadrupoles, etc.
Note that, when integrating over the panel, the second and third term at the
righthand side of eq. (2.2.83) vanish when =n=o represents the panel centre.
mri
= ddn ;
m,n,p = 0,1,2, (2.2. 8'fl
o
The evaluation of these integrals results in relatively simple expressions in
and n which are considerably less expensive to compute than the logarithms and
inverse tangents occurring in the exact aerodynamic influence coefficients. The
number of terms to be retained in eq (2.2.83) depends on the ratio -. In
practice it turns out that for first order methods the first term suffices if
< 0.25. This means that about 90% of all exact influence coefficients may be
replaced by monopole approximations. 0f the remaining 10% about one half can be
approximated by quadrupole expressions.
The multipole expansions are used in the following methods (see also table 2.1,
section 2.1.13)
DOUGLAS-NEUMANN (Hess & Smith)
BOEING TEA 230
NLR PANEL
MBB PANEL
BAC PANEL
PANAIR
Hess & Friedman
2-83
If the local geometry of the panels and the singularity distributions are
approximated by functions of a higher degree than piecewise constant, integrals
of the following type result (see eqs. (2.2.13) and (2.2.14))
2
m n r ]1/2
n I1 + (---) + (.---)
If +2 ddn (2.2.85)
[k i J
2
+ (.-)
i
2
+ (c.-c)
i
1p/2
Integrals of the type given by eq. (2.2.85) can be evaluated only by either
numerical integration or by approximation through suitable forms of series-
expansion.
The first approach (numerical integration) is found in refs. 2.14 and 2.23. In
these methods numerical integration methods are applied (6-or more point
Gaussian quadrature (see also ref. 2.24). The computation of the aerodynamic
influence coefficients by numerical integration is relatively expensive.
The series-expansion approach (reIs. 2.16. 2.17 and 2.25) is, although less
expensive in terms of computing time, more expensive in terms of code develop-
ment than the direct numerical integration approach. Due to the more complex
derivations and computer coding involved also the possibility of errors
increases.
2 2
(n.-n) + (ç_ç)2]P/2 = r
[k-) + (2.2.87)
where rf is the distance between the field point (.,n.,Ç.) and the projection
(,n,0) of a point (,n,Ç) of the panel on its tangent plane. Then can be
expanded as
1 Ç2
(2.2.89)
r r rfrf 2 r1 -):-f::-
This series expansion is valid for Ç/rf « 1 and therefore called 'small
curvature expansion'. Note that for r = 0(h)
rf = 0(h)
= 0(1)
Ç/r1 = 0(h)
because
recovered.
form
Substitution of eq. (2.2.89) into eq. (2.2.85) yields integrals of the
mn
I ddn (2.2.90)
mnp
r
However, the computing time increases considerably less rapidly with increasing
power of the truncation error as a result of the fact that only about 10% or
less of the aerodynamic influence coefficients are of the near-field type.
(p) = an(o) - 3s
(2.2.91)
a2
(u)
as
- anas (G) - as
(2.2.92a)
also
O(o) = J -(T) ds (2.2.94)
2-86
The conclusions (i) an (ii) should not be confused with the equivalence princi-
ple of doublet- and vortex-distributions.
A =
(2.2.95)
h-h -h (2.2.96)
Au= f
The potential induced by the source distribution on a single panel ois is given
by (see preceding sections)
=
1 1ddn (2.2.98)
3 i - kn r InI
i
where
r = [(_)2 + ()2 + (2.2.99)
If
r = [2 + n2. + (2.2. 100)
0 1 ]
then
2n.n 2ç.ç
r = r
o
[i
r2 + r2 - r2 +
2
i r2
+ j] 1/2
2
(2.2.101)
O O O Oi O O
where F,n are 0(h) and ç = 0(h2) + higher order. The terms to the right of the
vertical dotted line are non-zero only in case of curved panels.
(hZ jh/2
Since 0(ä.S) = O J
= 0(h2) (i + 0(h2) (see eq. (2.2.10)), it
ç
follows that
o(hz 1 1
= o[
h2
f10(h2)+ jl/2j
r ii h[0(1)+ 0(h) + 0(h2)+ ...0(hZ)...)'/2
(2.2.103)
1
o(hz = 0(h)
r inçi
i
o(h2 = 0(h) + 0(h2) + ... 0(h3)
r ini
Verify that the expression that is obtained contains some but not all terms of
2 2
3 3
term ç2/r gives rise to a term of 0(h3). However (truncated) cubic terms in the
geometry representation are not present.
Hence, the lowest order truncation error is 0(h3).
Multiplying (2.2.103) with the source density representation o = 0(1) + 0(h) +
the accuracy and truncation error (TE) levels of the approximation for ÓA'.
are found to be as follows:
full expression for ô are superfluous (do not really help); contributions to
the truncation error of geometry and singularity representations are not in
balance. This combination does not offer any real advantage over the preceeding
one.
superfluous terms 0(h3) and higher order in full expression for ó4; truncation
errors balanced.
Superfluous terms 0(h3) and higher order in full expression for ô; truncation
errors not balanced. This combination does also not offer a real advantage over
the preceeding one.
2. r
o
= 0(1), far field.
Here the magnitude of the subsequent terms in eq. (2.2.101) is found to be
so that
0(h2 )=o{ h2 *
r InI )1/2
r [i + 0(h) + 0(h2) + 0(h2) + 0(h )+. . .0(h3)...
o
In this 'far field' case where we have to do with many, 0(N), panels contribut-
ing to the total perturbation potential 4 = &D. Hence the summation over
0(N) = 0(j-) panels cancels the factor h2, and we find:
superfluous terms 0(h) and higher order in full expression for 0; geometry and
singularity related parts of truncation error not in balance;
superfluous terms 0(h2) and higher order in full expression for 0; geometry and
singularity related parts of truncation error in balance.
superfluous terms 0(h) and higher order in full expression for 0; geometry and
singularity related contributions to truncation error in balance;
From this 'error analysis' it appears that in order to have a local truncation
error of 0(h) in the potential it is (more than) sufficient to use planar panels
with a constant-strength source distribution. The analysis also suggests that
the computation of the near field influence coefficients as well as in the
computation of the far field influence coefficients can be approximated through
series expansions (such as multi-pole expansion).
In order to achieve a truncation error of 0(h2 ) planar panels with a linear
source distribution are (also more than) sufficient. In this case also
simplification of the integrals by means of power series expansions is possible
(in fact, planar panels with constant source density would be sufficient in the
near-field).
2-91
= _L11 a 1
()
ddn ddn
(2.2. 105)
¡n r ÍnI
i J
ç.-ç
0(ô) = o [h2 [l+0(h2)+..j'/2]
0(h2
1 0(h) + ;0(h2) + ... *
= o[h
r3
h{O(1) + 0(h) + 0(h2)+...0(h2)..j3/2
)1/2]
* (i + 0(h2) (2.2.106)
where the dotted lines have the same function as before. Then it is found that
ç.-ç
i
0[hz = 0(1) + T.E. of 0(h), for planar panels;
r3 InI
= 0(1) + 0(h) + + T.E. of 0(ha), for quadratic panels
0(d) = O(i)
0(TE) = 0(h)*0(l)+0(h)*0(l) = 0(h);
0(6e) = 0(1)+0(h)+0(h2)+....
0(TE) = 0(h3)
Note: A particular 'near field' case arise when the point (.,ri.,ç.) is located
on an (adjacent) panel. For a (continuous geometry we than have Ç= 0(h2),
rather than 0(h), and the degree of the truncation error changes
correspondingly.
0(e) = 0(1) +.. . + T.E. of 0(h), for planar panels with constant-strength
doublet distribution;
= 0(1) + 0(h) +.. + T.E. of 0(h2), for planar panels with linear doublet
.
distribution;
= 0(1) + 0(h) + 0(h2) +.. + T.E. of 0(h3 ), for quadratic panels with
.
A summary of the results for the integrals for the potential is given in Table
2.2.
Consistent representations
MCAERO *)
cubic')2) , doublet
=
1 ddn L O
r ddrt
lnI
(2. 2. 108)
r
os i
-
i
1
0(óø) = 0h2 o
o(h2 L
r2 Jn1
1
= o[hz h2{0(1)
+
1
0(h) + 0(h2)+ ..O(h2)...)
*
* 0(h2)+...J'/2)
[1+ (2.2.109)
Then
0(óV) = O(i) + T.E. of 0(h), for planar panels with constant-strength source
distribution;
= 0(1) + 0(h) + . . + T.E. of 0(h2), for quadratic panels with linear
source distribution;
0(h2 L
r2
1
)
= o[h2
r2 [i+ 0(h) + 0(h2) +
1
0(h2) + 0(h )+. . .0(h)...)
*
n o
ç
c.-c
ô..
j i-= - 4 li f1
i
ji
ji.V.
a
i
i dd
Jn1 -
i
i
j i r
i ddn
Inc1
- L i i ddn (2.2.111)
- 4ii
a.S j r5 + r3
i .c.
0(ó$) = 0(h2p + 0(hlJ
r3
2.96
ç.-ç
1 1 0(h) + 0(h2)+ ....
o(h2 - o[h * (2.2.112)
r - h0(1) + 0(h) + 0(h2)+...0(h2)..j2
1/2
* (i + 0(h)+...)
1
The saine holds for the 0(h2i term.
Then
0(ô) = o() + T.E. of 0(1), for planar panels with constant-strength
doublet distribution;
= o(i) + 0(1) + .... + T.E. of 0(h), for quadratic panels with
linear doublet distribution;
= o[) + 0(1) + 0(h) + ... + T.E. of 0(h2), for cubic panels with
quadratic doublet distribution.
=r. =ï.
3---
i
L
14n
rxds
r3
+-ff 1
L4
- r
{.pxn}x-
j r3
dçdn (2.2. 113)
j j
36 S
i i
and recall that the line integral represents the velocity induced by a line
vortex with strength p. around the circumference of the panel and the surface
integral represents the velocity induced by a distribution of vorticity with
strength p on the surface of the panel.
An order of magnitude analysis for the surface integral of (2.2.113) leads to:
2.97
o{hz 1
i o {h
0(h) + 0(h2) + *
Il
Inl
h3[0(l)+0(h)+0(h2)+...0(h3)...)3/2
1/2
* (140(h2 )+. . .)
and
o{ ¡J ]
= 0(1) + T.E. of 0(h) for planar panels with constant
ô .s
J
vorticity (linear doublet) distribution
Performing a similar analysis for the line integral it is found that this is
still 0(1/h).
Hence, it seems that we have solved the problem only partially.
The remaining problem with the line integral can be solved by combining
contributions of adjacent panels. For this purpose we consider the contribution
rxds
(2.2. ll24a)
i ,j=1 j r3
J
of the line integral of all panels to V.. Noting that adjacent panels have
partially coinciding edges ad that the ã at the common edge are of opposite
direction it follows that the expression for can be replaced by
°) ;xa (2.2.114b)
i =j=1 SnaókS
r3
Note that this represents a summation over the (common) panel edges rather than
over the panels themselves.
The virtue of using (2.2.l14) is in the fact that each of the individual
integrals of (2.2.114) remains finite, even for h40. This is due to the fact
that for each (common) edge we have
2.98
The property just discussed forms the basis of the so-called vortex-lattice-
(constant-strength doublet) method of thin-wing theory. The accuracy of the
vortex-lattice method can be increased further by shifting the position of the
vortex pair and the position of the collocation point to the 1/4 and 3/4 panel
chord point, respectively. Note, however, that the tangential velocity component
which should exhibit a discontinuous behaviour when passing through a panel, can
only be modelled properly through the local gradient of the doublet strength.
Hence, a linear doublet representation, at least locally, remains necessary for
a consistent evaluation of the near-field tangential velocity. For more details
on the vortex-lattice method the reader is referred to ref. 2.27.
= 0(1) + 0(h) + ... + T.E. of 0(h2), for planar panels with linear
doublet distribution;
= 0(1) + 0(h) + 0(h2) + + T.E. of 0(h3), for quadratic panels
with quadratic doublet distribution.
Conclusion: In general planar panels with linear doublet distribution are neces-
sary to achieve a truncation error of 0(h). Special measures (equivalent
vorticity formulation) are needed in the near field representation to keep the
perturbation velocity bounded for h -+ 0. The far-field integrals may be
simplified by power-series-expansions.
In order to attain a truncation error of 0(h2) the panel geometry and the
doublet distribution should both be quadratic. Both near-field and far-field
integrals may be simplified.
A summary of the results for the integrals of the perturbation velocity is given
in Table 2.3.
2.100
Consistent representation
DOUGLAS-NEUMANN
planar 1)2) constant2) linear2) 0(h) BOEING TEA 230+)
NLR+), MBB+), BAC+
(BAC)
In the preceding section we have seen that the concept of consistency tells us
something about the accuracy of the discretized representation of the integrals
(local truncation error) for given distribution of the source or doublet
densities. However, nothing has been said about the accuracy of the solution of
the discretized integral equations. For a judgement of this aspect we need the
concept of convergence, (ref. 2.24).
A numerical method is called convergent (see also eq. (2.2.97)) if, for h O
Iläull = I
- u-h J <kh (2.2.116)
A = (2.2.95)
-h
and u that of the discretized problem
A
h-h
u =f-h (2.2.96)
h
The difference is called the global truncation error.
-h h-h
Ilu I
<k2 lIA u J
(2.2.117)
-h h h-h
H = lEA j-' hll < I[Ah]lH IA u (2.2. 118)
2.102
(2.2. 119)
I
Hence
- -h- hi-1 h -
óu=u-u 4LAJ r
IA-Aju for h 0 (2.2.120)
o r
Ioii I
lull
{] 1AhA) H
[AhJ
I I
IAAhI I
(2.2.122)
I
lAAhl I
= kha, for h 0 , II =i
Hence
k
h°, for h 0 (2.2.124)
II{A']ll <
o
2.103
the numerical method is also convergent. In particular there holds that the
global truncation error in the solution and the local truncation error in the
integrals are of the saine order in h if
h-1
[A ] H is bounded (arid non-zero) for h O (2 .2. 125)
Eq. (2.2.125) expresses the same as eq. (2.2.119) and we may conclude that
stability guarantees convergence to the same order in h as the local truncation
error.
It may be remarked also that the norm I[Ah]_lIj is related to the condition
number of the matrix Ah (ref. 2.24) by
From the preceeding analysis it follows that a good estimation of the norm of
the inverse matrix [Ah] is an essential ingredient of the stability analysis
of a numerical method. Unfortunately, good estimates of' such norms are generally
not available in the case of panelmethods. If they exist it concerns idealized
situations in two-dimensions, in which the matrix Ah can be reduced to diagonal
from using Fourier transformations.
Because of this lack of information about the norm of the inverse matrix the
estimation of the global truncation error of panel methods is in general based
on empirical/heuristical arguments. For example, it has been established
empirically, that the stability problem of' panel methods corresponds, in certain
aspects, to the problem of approximating a function by means of' piecewise defined
polynomials. In both cases lack of stability manifests itself in high-frequency
point to point oscillations. In the following we shall consider a number of'
stable and unstable formulations. In that respect we will distinguish between
(see fig. 2.2.19)
- absolute stability, local errors are damped and do not propagate through the
entire solution
- neutral stability, a local error propagates but does not increase
- instability, local errors propagate and increase in magnitude.
2.104
The problem leds to a system of linear algebraic equations for the coefficients
f'., etc.
f°.,
J_ i
Fredhoim integral equation of the second kind
pi ç'-c
Z pK2dS K2 - (2.2.128)
i os.
J
o. --
oK4dS ; K4 (2.2.129)
i os.
J ii3
Inspection of these expressions teaches that the terms p.12 and o/2 are
strongly dominant and, in fact, independent of the panel dimension h (K2 and K14
are zero for flat surfaces).
Then, if p,o are piecewise defined polynomials of the type f.(x), mentioned
above, prescribing is (almost) equivalent with prescribing p. and, hence,
.
oK dS K =1 (2.2.130)
' ;
1 1 r
jóS
i
2.105
r.n
i
= p1K dS K3 (2.2.131)
ani jj5
J as.
J
- + Ii3
In these cases a clear, strongly dominant term is absent. In fact, inspection of
the kernel K1 teaches that (even for flat surfaces) all source panels contribute
significantly to ' with the largest contributions coming from the 'near-field'
panels. A more detailed analysis (see section 2.2.6) teaches that the local
contribution of o itself is weakly dominant (0(h); the contribution of o' being
0(h2)). The implication of this is that solving the source/Dirichlet problem
(2.2.130) is equivalent, but in a weak sense only, to solving the spline-fit
problem with f. given. The same conclusion is obtained with respect to the
doublet/Neumann problem (2.2.131). (Remember, from section 2.2.6, that the near-
field contribution of p is 0(1/h) and 0(1) for p').
= I f! X dS
II
( )
.j os
Neutrally stable
Unstable
ist & 2nd kind C linear source dist. - Neumann and Dirichlet b.c
o
C linear doublet dist. - Neumann and Dirichiet b.c
Note that ail the absolutely stable formulations are associated with Fredhoim
integral equations of the 2nd kind. This is the main reason for preferring 2nd
over ist kind equations.
2.107
Figs. 2.2.30 and 2.2.31 give comparisons of results of ist generation, ist order
panel methods (NLR PANEL, BAC/H(unt)-S(emple)) with a high accuracy 3rd order
solution (BAC/Roberts) for a 3-D swept wing. For the thick (15%) wing case of
fig. 2.2.30 the agreement is quite acceptable for both 360 and 730 panels (per
half wing). For the thin (5%) wing (fig. 2.2.31) this is no longer the case for
the lower number of panels.
This is associated with the fact that the representation of the (near-field)
tangential velocity induced at the surface panels by doublet/vortex panels on
the wing camber surface (table 2.1) becomes inconsistent for vanishing thickness
(see also the discussion in section 2.2.6). Methods employing (surface)
vorticity distributions or internal Dirichlet boundary conditions do not suffer
from such behaviour.
Figs. 2.2.32 to 2.2.3L illustrate that particular accuracy problems arise in
case of internal (duct) flows modeled with source distributions. Part of the
mechanism is sketched in fig. 2.2.32 and is associated with the fact that the
2.108
local truncation error in the normal velocity induced by flat source panels on
ducts with circular cross-sections is equal to /2u, where O is the subtended
angle of the panel. It is further understood that the global truncation error is
also affected by the fact that modelling with source distributions satisfying
discrete Neumann-type boundary conditions is not (numerically) mass-conserving.
This is due to the fact that sources produce mass and that the sum over all
sources is not necessarily exactly equal to zero. For external flows this is of
relatively little concern, because the total mass flow external to the body is
unbounded. For internal flows, in which the total mass flow is bounded, the
error may become relatively large, in particular for channels with rapidly
varying cross-section where the source strengths are locally large.
Numerical models utilising doublet distributions do not suffer from this effect,
due to the fact that doublets have zero net mass production.
Figs. 2.2.33 (flow-through nacelle or ring-wing) and 2.2.34 (duct with
contraction) illustrate the seriousness of the problem for ist order source-
based methods.
2.2.9 Solution methods for the resulting system of linear a1ebraic equations;
computational effort
We recall that after discretisation we are left with the problem of solving a
large (0(103)) system of linear algebraic equations for the unknown source or
doublet strengths. The system may be written as
A = b
Direct methods
Gauss-elimination
In this method the original system A=6 is reduced to a more convenient one
Ux=b', that can be solved readily, by successive elimination of unknowns from
the equations.
Symbolically:
A =6 = 6' (2.2.127)
A stepl step2 U
In this elimination process the righthand side also changes. The system Ux= 6'
is readily solved by proceeding from the bottom to the top of the system. A
necessary and sufficient condition for this method to work is that the pivot
points (elements on the diagonal of the matrices) are 0. A matrix that does
not satisfy this condition is called ill-conditioned.
A = . U =
fi o o
(2.2.128)
L= m21 i O
m31 m32 i O
m,43 i
\ml m142
a
where the m.. = are the multiplyers from the Gauss-elimination process
.11 ai*1)
define a vector
(the superscript denotes the step in the procedure), and we also
s such that
(2.2. 129)
Ls =
then we have
(2.2.130)
= L' = L'
hence
(2.2.131)
U =
and
(2.2.132)
= U1s = U'L1L
and therefore
(2.2.133)
(LU)x = b = Ax
(2.2.131-i)
or A = LU
and from
Once L is known, may be determined readily from eq. (2.2.129/130)
eq. (2.2.131/132).
required for
The LU decomposition procedure has advantages if solutions are
(several angles of attack,
several different sets of righthand side vectors
(once and for all)
several normal velocity distributions); with U and L known
* vector (L) multiplication for the
can be determined through simple matrix
various b.
2.111
n(n-1)
s=L -1-
-
b: in
*
2
ops.
where m is the number of righthand side vectors for which the computation has to
be repeated.
Iterative methods
With iterative methods the matrix is not decomposed or factored (e.g. A = LU)
but it is split up, in some suitable way, as
A = N + p (2.2.136)
Ax = b
Nx = - (2 .2. 137)
= (2 .2. 138)
(_1)j
N{x' = b - (2 .2. 139)
or
(v-1)
NC'
((l)
N(- ) = - p )
(2.2.140)
or
N'=
where is the error of the vth iterate.
Multiplying with N gives
(v-l)
- N1
= -
(2.2.1141)
It can be shown (ref. 2.2k) that the iterative proces converges if the matrix M
is 'convergent', or
hm M' = O
'J4
where A. are the eigenvalues of M. Since p(M) < 11Ml I this may also be stated as
(2.2.142)
lIMlI <1
I II (2.2.143)
(u-1)
ll
I
lll
-1)
I
IIMH
Hence,
(2.2.145)
p(M)IIMII < 1
Jacobi-iteration
In the Jacobi-iteration, also called simultaneous iteration, the N matrix of the
preceding method is the diagonal matrix, such that
A = N + P
Gauss-Seidel iteration
The Gauss-Seidel iteration procedure (also called successive iteration) uses as
N matrix the lower triangle of the A matrix, thus
A N P
xxxx\ x000' o X x x
Block-iteration
Panel methods for 3D flow problems generally result in large matrices, which,
possibly, do not fit in the central memory of the computer. In that case the
matrix is organized in blocks (fig. 2.2.35). The Jacobi-iteration or the Gauss-
Seidel iteration is then applied per block while the blocks on the diagonal are
inverted by means of Gaussian elimination. The blocks are chosen in such a way
that they represent the influence of strips or segments of panels. The blocks on
the diagonal represent the influence of a strip or segment on itself.
It is also possible to choose N such that in addition to the diagonal blocks it
contains blocks adjacent to the diagonal (block-tn-diagonal matrix).
Advantages and disadvantages of iterative methods
The main advantage of iterative methods is in the computational effort. The
operations count is given by k*n2 operations, where k is the number of
iterations and n is the number of panels. This figure must be compared to 0(n3)
for direct methods. For large n and sufficiently small k, the iterative methods
are superior to the direct methods.
However, in order to have a fast convergence (small number of iterations) it is
necessary that the diagonal (blocks) of N and A contain elements that are
'sufficiently large'. This is generally so for Fredhoirn integral equations of
the 2nd kind and generally not so for Fredholm integral equations of the ist
kind. The larger the blocks the greater the probability of convergence (in this
respect it may be noted that there even exist converging block-iterative methods
for ist kind Fredholm integral equations!)
The precise structure of the matrix and, hence, the rate of convergence is
determined by the geometry of the problem and by the way panels are organised in
strips, (which is reflected in the block structure). The Gauss-Seidel procedure,
in this respect, is somewhat more forgiving than the Jaconi procedure, although
both methods formally have identical conditions for convergence.
Panel methods based on Fredhoim integral equations of the 2nd kind require, for
not too complex geometries, about 10 to 20 iterations. In such a case the
iterative methods are faster than direct methods if the number of panels exceeds
200 - 300, as is shown in fig. 2.2.36. However, iterative methods are sensitive
to 'unfavourable' paneling, causing large elements outside the diagonal blocks
that have a negative effect on the rate of convergence. For complex
configurations this may be unavoidable. In other words iterative methods are
less robust than direct methods.
Examples of panel methods using iterative solution procedures are
DOUGLAS-NEUMANN (Hess & Smith, ref. 2.10), Gauss-Seidel
NLR Panel, block Gauss-Seidel
MBB Panel
BAC (Hunt & Semple, ref. 2.15)
VSAERO (optional)
DOUGLAS-NEUMANN Lifting 2nd order (Hess & Friedman, ref. 2.17)
a wing-body configuration with about 730 panels. The computation was performed
using the NLR panel method with an iterative solver on a Cyber 180-855 computer
(comparable with the IBM-3080 series):
We have discussed in section 2.2.1 that, once the solution (in terms of the
source strength distribution o and/or the doublet strength distribution p) of
the discretized integral equation is known, we are able to compute the velocity
arid pressure everywhere in the flow field and thus also in the panel centres.
Forces and moments on the body may then be determined by numerical integration
of the pressure distribution. This should be done through an integration
rule with a truncation error of the same (or higher) order as the truncation
error of the calculated velocity or pressure distribution. Usually, the simple
mid-point rule suffices. For example one may use for the force in x-direction
NB
K
X
= - f! pn
S5
X
dS = - I
i=1
xi i
(pn ). ó.S (+ 0(h2)) (2.2.1146)
K NB
z 1 (2.2.1147)
1 (Cn) ô.S
CL=pU2S s i=l
i
2 ref ref
2.117
K NB
X 1
CD=l (Cn) ó.S
i
(2.2. 148)
PUZ S S i=1
2 ref ref
In eqs. (2.2.147) and (2.2.148) Sref is a reference area, e.g. the wing area,
and C the local pressure coefficient. CL and CD may also be determined for a
(wing-) strip.
The pitching moment coefficient taken with respect to a given point (x,z) can
be written as
N
B
pzi (xO-x i
1
[ (C n J J
+ (C n
pxi )
(z -z J )o S
o i
(2.2. 149)
S Q i
ref refi=1
usually acceptable for CL, but one would like to have an error of .00O1
(1"count") in CD. Secondly the strong curvature of the surface at the wing nose
results in strong pressure gradients that may produce large errors in CD unless
a very fine panel distribution is used, see the qualitative sketch of fig.
2.2.38.
According to the Kutta-Youkowski theorem the drag must be zero for a closed body
in a 2-D potential flow. Unfortunately, 2-D panelmethods do not give zero drag,
the error depends on the number distribution and of panels and on the angle of
attack; the error may be positive or negative. First order panelmethods using 60
to 80 panels around a wing profile may show errors ranging from a few counts to
lo counts depending on the airfoil shape and the angle of attack. For a typical
value of CD = 0.01 for a wing profile with turbulent boundary layer flow this
implies an error of few percent to 10 percent or more. For higher order methods
the errors are usually smaller, in general less than 5 counts. The error tends,
of course, to zero when the panel dimension h approaches zero.
2.118
In 3-D potential flow the drag that is modeled by the panel methods is the
induced drag, in other words the drag due to the downwash of the vortex wake.
Hence, we may conclude that in 3-D potential flow the integrated pressure drag
is equal to the induced drag.
A better way of evaluating the (induced) drag (it may also be done for the lift)
is obtained by applying the momentum conservation law to a closed controle
volume the bounding surface of which is at large distance from the body, see
fig. 2.2.39.
For the component of momentum in x-direction we then find
vector and U the x-component of . From eq. (2.2.150) we find that the
(pressure) drag D can be expressed as
Eq. (2.2.151) may be simplified by letting S+; in that case p+0 and the
contribution of S vanishes. The contribution of the integral over S (the
It can be shown (see e.g. ref. 2.53) that the surface integral (2.2.152) can be
reduced to the line integral
D = - f
P
SQS,
D = - f r dt = D. (2.2.153)
P
SS i
where t is the direction along the intersection SCIÌST and the normal on S,
see fig. 2.2.39.
The integral (2.2.153) can be approximated numerically and the experience is
that evaluation of the (induced) drag along this way is more accurate than the
pressure integration procedure discussed previously. Fig. 2.2.40 provides an
impression of the differences in results for the two procedures for a simple 3-D
wing (the same wing as that of fig. 2.2.31) as computed with the NLR panel
method for 60 x 12 panels.
The Trefftz-plane procedure may also be applied to complex non-planar
configurations consisting of several lifting surfaces.
As mentioned already in section 2.1.1 panel methods may also be used to obtain
approximate solutions of the linearized potential (Prandtl-Glauert) equation
governing compressible flow problems. It was also indicated that, because of the
fact that the Prandtl-Glauert equation can be transformed to Laplace's equation
by a simple coordinate transformation, everything said so far on panel methods
for Laplace's equation is also applicable to panel methods for the Prandtl-
Glauert equation. (In fact most panel methods solve the Prandtl-Glauert rather
than Laplace's equation.)
However, the precise way in which various panel methods handle (linear)
compressibility effects varies from one method to the other. It is useful,
therefor to discuss this matter in some detail.
2.120
,2O
xx
+4> +4>
zz
=0
yy
or (2.3.1)
._(,24>J
ax X ay y
+ az-
a
(4>
zj
= 0
x=x
(2.3.2)
=
we obtain
+ <p + = O (2.3.3)
xx yy zz
(2.3.4)
(Ux + = O
or
xx +4>n +4>n
zz =-Un
4>n (2.3.5)
yy
the
Recall also that eq. (2.3.1) stems from the mass conservation law under
assumption of irrotational flow and small perturbations, i.e.
Since 4=i, it folows from eq. (2.3.5) that there must hold
(2.3.7)
n = 0(c); n ,n
z
= 0(1)
y
zz =-Un 0(c2)
(2.3.8)
4>n +4>
yy
or
2i21
+ n = - Un
oex
+ 0(c2) (2.3.9)
yy z
- - l( -n
1 1
-n (2.3.10)
n (n n- n ) = - nx
- F y z
X y z
where
F2 = ,2n2 + n2 + n2 (2.3.11)
X y z
we have
n = Fn
X -X
n = Fn (2.3.12)
y
= Fn
+ = - U= + 0(c) (2.3.13)
= (2.3.14)
+ + (2.3.15)
= - Uco
yy
- -
zz X
X
X
= ct, (2.3.16)
y
z
=-
13
2. 122
Eqs. (2.3.2), (2.3.114) to (2.3.16) represent the well known Göthert rule from
classical small perturbation theory. The Göthert rule enables us to obtain an
approximate solution of the compressible flow around a configuration from the
solution of an incompressible flow past a similar or 'analogous' configuration
that is determined by the transformation (2.3.2) with the boundary conditions
(2.3.15) (the term 0(c2) can be neglected).
For panel methods which utilize the 'exact' boundary condition (2.3.5) for the
solution of the incompressible flow problem, the question may be raised what to
do with the term n (= 0(c2)J in the compressible flow case when the Göthert
rule is applied.
One of the possibilities is to use, instead of eq. (2.3.15), the complete
boundary condition of the analogous incompressible flow, i.e.
+ + = - uÇ (2. 3 . 17)
xx yy zz X
2n
xx +n
yy +nzz=-Un (2.3.18)
which is not the same as the real boundary condition eq. (2.3.5).
However, for small perturbations the error is only of 0(c2), thus, formally, in
the asymptotic sense (c-*0) the error is of the same order as that of the
linearized boundary condition given in eq. (2.3.8) or eq. (2.3.15). More
importantly, eq. (2.3.18) unlike (2.3.15) reduces to the exact boundary
condition (2.3.5) for M+0 (-l) . This suggests a preference for eq.
(2.3.17/18).
At positions where locally nx 0(1), such as at the wing nose, the small
perturbation theory on which the partial differential equation (2.3.1) is based
is violated anyway, and neither (2.3.17/18) nor (2.3.7/8) provides a valid
approximation for compressible flow.
The Göthert rule given in the form of eq. (2.3.2), (2.3.16), (2.3.17) is
sometimes referred to as Göthert rule I; it is applied for example in the NLR
panel method.
2.123
(P)
1
= - fi
ku
1
oH -) dS +
a
(-1)d (2.3.19)
r anQ r
where the integrals are obtained from chapter 2.1, eq. (2.1.107) and its
gradient; the tildas denotes the transformed quantities according to eq.
(2.3.2).
Next the Göthert rule (eq. (2.3.16)) is applied to the components of the
perturbation velocity of the transformed problem. In vector notation this may be
written as
(2.3.21)
where
loo
[B] = O O (2.3.22)
oo
Finally the 'real' boundary conditions are imposed according to eq. (2.3.5),
thus
= - U n (2.3.23a)
1---
-n --
+ n -- + n = - Un (2.3.23b)
xx yy zz X
The procedure just described is applied in the panel methods of MBB (Kraus &
Sacher, ref. 2.13) and of BAC (Hunt & Semple, ref. 2.15). It is known as the
Göthert rule II.
2.124
The method has the formal disadvantage that the integral equation which is
eventually solved is not derived from the application of Green's theorem to the
governing partial differential equation, i.e. the Prandtl-Glauert equation. This
implies that the uniqueness of the solution cannot be proven in a formal way,
see also ref. 2.30; in practice this does not appear to be a problem.
A third possibility for solving the compressible flow problem with panel methods
is based on a 'direct distribution' on the surface of a configuration with
elementary solutions of the Prandtl-Glauert equation in psysical space. In other
words, we assume
(P) = j- Jf oH )
dS + fi p 3flQ
1
)
dS (2.3.24)
s r 11s r
or
fi o(- dS + j-- if H dS (2.3.25)
P) = )
i --p-- )
S r 11S anQ r
eq. (2.3.5).
Like in the case of the Göthert II method also the integral representations
given by eqs. (2.3.24) and (2.3.25) are not obtained from the application of
Green's theorem to the Prandtl-Glauert equation. If we do this (see ref. 2.30)
we find
(P) = L ji
-
dS + L j QV3 C- dS (2.3.27)
4rT
S r S r
where
-
= -Bx + -ay+ -az=
a a a 2
[sW'] ([s])
2.125
with the matrix [B] given by eq. (2.3.22). Then eq. (2.3.27) can be written as
= f! o[- )
dS + ff pQ.[B] ([B']) (- )
dS (2.3.28)
S r S r
3- ( 2-Y ) M2
pe[u M22 M2 (2 + 2)
x = + (1-M2e)
x
-
2 ex - 2 e x z
+
y = e [ y - M2e xy + (2.3.29)
= pe [i - M2 +
z z xz
Then, the lowest order approximation to the normal component of the mass flux
vector becomes
n
xx + n
yy + n
zz ex
= - Un (2.3.31)
or
j2 [B] ([s]V) = - U n ex (2.3.32)
Note that this is the same as the boundary condition (2.3.18) of the Göthert I
model after back transformation to the physical space.
The mass-flux boundary condition is applied in the panel methods PANAIR and
QUADPAN.
The integral representations and the imposed boundary conditions of all subsonic
compressible flow models mentioned above may be expressed in terms of quantities
2.126
= [B] i (2.3.33)
[B] (2.3.34)
I[B'] $
= 2
I[B'} :;i dS (2.3.35)
= [B] (2.3.36)
a
=
-- [B
-1
] n
-
[B1] (2.3.37)
Bn [B ] n
where
i o o
[B1] = o o
o o
i
We then obtain:
Göthert I
(P) =
1 1ff(l)d+lff (-è) d]
r anQ r
s s
= L!
1
- If o
s
r
) [] I
dS + ff
s
[B1J .
{B1] VQ (-
r
)
dS
(2.3.38)
boundary condition
2
[B'] [81] = - U n (2.3.39)
2.127
Göthert II
Integral representation as Göthert I;
boundary condition
= - U n
mx (2.3.40)
Direct I
boundary condition
'exact': eq. (2.3.23) or eq. (2.3.40)
or 'mass-flux': eq. (2.3.31) or (2.3.39)
Direct II
(P) = f! o -
1) dS + ([B] (1) dS
r I
r
s s
(2.3.42)
1)
(P) = f! o - dS + lip [B] ([B]Q) (- )
dS
r r
s s
(2.3.43)
The five models are equivalent in terms of subsonic thin airfoil theory. All
models provide (numerical approximations of the) solutions of the Prandtl-
Glauert equation.
For thick bodies those formulations are equivalent which work with the same
boundary condition. The elementary solutions forming the surface distributions
are different, but they become identical for M-0 as is also the case for the
velocity boundary condition and the mass-flux boundary condition.
2.128
The expressions, in terms of a and p, for the jump of the potential and the
normal velocity component across the surface S are different for MO and
In particular the doublet-like distributions of eqs. (2.3.38/41/42/43) do not
only exhibit a jump in the potential but now also in the normal velocity. This
is due to the fact that the kernelfunction in eqs. (2.3.38/41/42/43) differ from
those in the incompressible case. This can be interpreted as that in the
compressible case the doublet axis does not coincide any more with the surface
normal. Note however, that the jump of the normal component of the linearized
mass-flux is zero for the doublet-like distribution of eq. (2.3.43). Therefore,
this integral representation is particularly attractive in combination with
mass-flux boundary conditions. The jump behaviour in the case of a source
distribution is, qualitatively, the saine as in the incompressible case.
The difference in jump behaviour has also consequences for the equivalent
internal Dirichlet boundary conditions.
Evidently all models fail in the vicinity of blunt leading edges. Moreover, also
in regions where the approximations are valid (nx small), the error for a given
geometry increases with increasing Mach number, which is due to neglecting the
higher order terms in eq. (2.3.29). However, the thinner or more slender the
configuration the higher the Mach number can be for reasonable results.
For some panel methods (NLR, BAC) semi-emperical correction methods for non-
linear compressibility effects have been developed, which decrease the errors at
blunt noses and at high subsonic Mach numbers. Such correction methods are based
on analytical solutions of higher order approximations of the potential equation
for the compressible flow past simple geometries (ellipses etc.), see ref. 2.31.
Fig. 2.3.1 shows some examples of the application of such a correction method.
(NLR-panel method, as applied to airfoils)
2.4.1 Status and role of panel methods in the process of aircraft aerodynamic
design
In the preliminary (or even conceptual) design phase (phase I in fig. 2.4.1)
panel methods are used at an increasing rate for the purpose of obtaining rapid
estimates of the forces and load distributions on (complete) aircraft
configurations. As such there is a tendency to replace (part of) the functions
of empirical data base type methods such as the well-known USAF 'DATCOM' and
ESDU 'Data Sheets'.
In phase II (Design Development) one is concerned with the detailed shaping of
(parts of) aircraft configurations. Fig. 2.4.2 depicts the design development
process in some detail.
In the process (A) detailed shapes are generated, which are analysed in the
processes (B) and (C) . Panel methods are used in process (A) as well as in
process (B).
In phase II (B) panel methods serve to analyse details of the flow pattern, load
and pressure distributions on parts of or complete configurations. In phase II
(A) they serve to solve the 'inverse' problem: given the pressure distribution,
what is the shape that produces that pressure distribution?
Note that:
- the viscous effects, at the relatively low Reynolds number of 0.65 x 10' cause
a decrease of the lift of about 20% and an increase of the moment.
- the increase of the viscous drag with the lift should be positive,
illustrating that the integrated pressure drag in fig. 2.3.6 is not correct.
- viscous effects increase rapidly when flow separation occurs (CL 0.65).
Results of computations are shown in figs. 2.4.10 and 2.4.11. In the panel
method used, some 350 panels are applied on fuselage and fairing, and about 700
panels on the wing; for symmetry reasons only one half of the configuration is
considered.
It can be observed from the results that the effect of the presence of the
fuselage on the wing load and pressure distribution is considerable. Again the
main reason for the difference between theoretical and experimental results is
found in viscous effects.
Ssee figs. 2.4.12 and 2.4.13. The following observations may be made.
Variation of the position of the vortex sheet from the wing does not notably
effect the solution, as may be observed from the lack of difference between the
computed results for an undisplaced vortex sheet (='wake') ('initial ist wake
estimate', i.e. straight downstream) and a displaced (corrected) vortex sheet,
see fig. 2.4.12b and 2.4.13a,b,d. The corrected vortex sheet was obtained by
integrating once the normal velocities on the vortex sheet to obtain a stream
surface.
However, the results reveal clearly the effect of a sting support, especially on
The remaining causes of the discrepancies between theory and experiment are
thought to be:
- viscous effects on the various aircraft parts (in particular separation on
the rear part of the fuselage);
- absence of the modeling of the vortex sheet (wake) leaving the rear part of
the fuselage.
The configuration and paneling (1219 panels) are shown in fig. 2.4.14; the
schematized vortex sheet models and position of vortex sheet are depicted in
fig. 2.4.15. The results are shown in figs. 2.4.16 to 2.4.19.
Note that in this case the (correct) positioning of the vortex sheet of the main
wing relative to the flap has a significant effect. Note also that the absence
of the modeling of' flow separation at the side edges of wing tips and flap tips
is clearly reflected in the local pressure distributions (figs. 2.4.18/19a).
See fig. 2.4.20. This example illustrates the possibility to model the effect of
nacelle and engine control on the lift distribution on wing and flaps. No
experimental results are available.
In the preceding chapter we have observed that viscous effects can lead to
significant differences between potential flow theory (panel methods) and
experiment. From classical boundary layer theory (Prandtl boundary layer
approximations) we know that for high Reynolds numbers and attached flow
viscous effects are limited to the boundary layer and its extension (wake)
behind the airfoil. In such conditions, viscosity has a relatively small
effect on the forces and pressure distribution acting on the body.
According to the Prandtl's boundary layer theory, the high Reynolds number flow
about a body can be modelled, approximately, by distinguishing two 'zones': an
iriviscid part governed by the potential eq. (or Euler eqs.) and a viscous part,
governed by the boundary layer equations. We call this zonal modeling, see also
section 1.3.2.
That such a subdivision is allowed follows also from more formal consideration
based on the theory of matched asymptotic expansions, ref. 2.33 of
characteristics of the Navier-Stokes equations at high Reynolds numbers.
In the theory of matched asymptotic expansions solutions of the Navier-Stokes
equations for high Reynolds number are expressed in terms of series expansions
with the Reynolds number (Re = UL/') as expansion parameter and the airfoil
chord (or body length) L and boundary layer thickness ô (- Re )
as
characteristic lengths.
2.133
- 1) 1/2 (i)
u (x,y,z; Re) = u0 (x,y,z) + Re u1 (x,y,z) + ...; Re (2.5.1)
with similar expressions for the pressure and density. The leading (that is
first) term of (2.5.1) must satisfy the inviscid flow equations which are
obtained by substituting (2.5.1) into the Navier-Stokes equations and retaining
only the lowest order ternis. The boundary condition on the body is
(i)
u0(x,y,O) .n=0.
(y) (i)
hm u0 (x,y,z/ó) = lu u0 (x,y,z) (2.5.3)
z/ó+ z+0
Extending the matching technique to the second order term in the 'outer
expansion' , it is found that this second order term represents the effect of the
displacement thickness of boundary layer and wake. The displacement thickness
influences the velocity and pressure distribution of the inviscid outerflow (the
outer flow has to be adapted to a new shape including the displacement
thickness). In turn, the modified pressure distribution has its influence on the
boundary layer development, etc. In this way an iterative proces may be thought
of, where, alternatingly, the inviscid outerflow and the boundary layer and wake
are computed until the pressure distribution and displacement thickness in
innerfiow and outerflow are matched together. In the block diagram of fig. 2.5.1
this process is illustrated. In the remaining part of this chapter we shall
confine ourselves to the modeling of the displacement effect in the outerflow of
2.1314
boundary layer and wake. The computation of the boundary layer itself is not
considered.
The displacement effect on the outerfiow may be modeled in various ways, see
ref. 2.3k. Two of the most important methods are discussed below.
1. The displacement thickness of boundary layer and wake, in 2D flow given by
* o
0 f l U
pU
j dz (2.5.14)
O ee
is added to the body and the dividing streamline behind the trailing edge (see
fig. 2.5.2). In the expression for the displacement thickness, O is the boundary
layer or wake thickness; the subscript e refers to the edge of the boundary
layer or wake.
The potential flow is then computed for the thickened body and wake
(displacement surface). The boundary conditions are: zero normal velocity at the
surface of the displacement body and plus the additional condition of zero
pressure difference across the wake (j p = O). Also a modified Kutta condition
must be satisfied at the transition between displacement body and wake at the
trailing edge. The additional condition at the wake causes the problem to be
overdetermined if the position of the (thick) wake is kept fixed (see section
2.1.3); it requires that the wake position is determined iteratively (non-linear
problem).
2. An alternative is to model the displacement effect by prescribing a given
in/out flow at the original profile contour and at the (infinitesemally thin)
vortex sheet (transpiration model).
Through asymptotic theory it can be shown that the displacement effect is
simulated with the same accuracy as in the first method if the in/out flow is
prescribed as
2D:
an
= li
p
-
ds
U
ee
0*) (2.5.5)
e
or,
ee + ô[as (p ee
U
3D: p - () = p W J + (PeVe)] (2.5.6)
Here s,t are orthogonal coordinates on the surface and Ue ,Ve ,W e are velocity
components at the edge of the boundary layer (e) in the directions s,t and
normal to the surface, respectively. The 2D case is sketched in fig. 2.5.3.
2.135
In linearized flow theory the lefthand side of eq. (2.5.6) may be approximated
by (see section 2.3.3)
-
3n
(p
n )
= (u +
x
)n
x
+ c
n
yy
+ n
zz (2.5.7)
Also in the case of the transpiration model we need an adapted Kutta condition.
The corresponding (jump) conditions on the vortex wake now also require a source
distribution on Sc.
The transpiration model is preferred over the first model in which the
displacement thickness is just added to the actual model and trailing vortex
sheet contours. The reason is that the (modified) boundary condition is imposed
at the original contour and therefore the aerodynamic influence coefficients
remain unaltered during the proces of iteration shown in fig. 2.5.1.
Panel methods with simulation of viscous effects are commonly applied for the
analysis of the flow past wing profiles (2D flow) with flaps (airfoils in take-
off and landing configurations), see figs. 2.5.4 and 2.5.5. For 3D wings there,
to the author's knowledge, are only a few methods available in which viscous
effects are taken into account; results from ref. 2.16 are shown in fig. 2.5.6.
In addition to the panel methods for subsonic flow past thick, lifting
configurations of given shape described uptil now there are also methods for
special applications. The most important of these will be discussed briefly in
the following sections.
In thin-airfoil theory also one looks for solutions of the linearized potential
equation. However, instead of satisfying the full zero mass flux boundary
condition at the surface of the body an approximate boundary condition of the
form of eq. 2.3.8 is used, which is imposed at the chord of the profile (2D) or
at the wing reference plane S(3D), see fig. 2.6.1. In the latter case the
approximated boundary condition reads
at the upperside of S w of S w
+ + + + +
y y
n
z
n =-Un
z
(2.6.1)
n + n = - U oex
n (2.6.2)
y y z z
+ + +
Here n, n, n z are the components of the normal on the actual wing surface.
x y
Usually the coordinate system is chosen such that the x,y plane coincides with
or is parallel to the wing reference plane S. Then, for thin wings, in addition
to also yfly may be neglected (since then n =0(E)), and eqs. (2.6.1) and
(2.6.2) reduce to
2.137
+
- +
+ n -
'
-
z
=-U =Uax x
+
aZ
onS w (2.6.3)
n
z
aZt
-ii =
ax
= 2U
ax
az - aZ
+
w
= +
z
+
=
U Caz
ax
+ 1
ax
= u ax
aZ
where is the chordwise slope of the thickness distribution of the wing and
ax
az
the slope of the camberline plus the angle of attack.
In order to model the lift we have, again, to introduce a discontinuity surface
extending downstream of S. The boundary conditions on S are the same as
those in the 'thick' wing problem.
From the application of Green's theorem it follows that the solution of the thin
wing problem may be represented in terms of a source plus a doublet distribution
on SW and a doublet distribution on S. The jump in normal velocity over S
given by eq. (2.6.4) can be related directly to the source distribution on S;
its strength follows from
+ aZ(x,y)
G(x,y) = FL(x,y)]J = 2U_ (2.6.6)
= J' a dS
(
r ) (2.6.7)
S
w
2.138
With the source strength on S known directly from eq. (2.6.6) only the doublet
distribution on S + S remains as unknown. This unknown doublet distribution is
w C
determined from eq. (2.6.5) by expressing the mean normal velocity.
:; . = (2.6.8)
+
w 2 z z
as
w p
=
)4
If pFi
w p an -- (1) r
dS + fío w p (
1) dS
r (2.6.9)
wC w
Here the index p refers to the point on S + S where the gradient is taken and
ii is the normal on S + S see fig. 2.6.1. Note that if S is a planar surface
,
w w C w
the second integral (the source distribution) vanishes.
Substitution of eqs. (2.6.8) and (2.6.9) into eq. (2.6.5) results in the
following integral equation for the unknown p
w pan w
a
- r ) dS=U
aZ
W_jj0j. (-)dsr
S+S
wC S
w
(2.6.10)
In practice, methods based on pure thin-wing theory are seldom used. On a large
scale, however, methods are used which combine thin-wing theory for lifting
parts (wings, tailplanes, winglets etc.) with distributions on the actual
surface of fuselage-like parts, see fig. 2.6.2. The integral equation that has
to be solved in that case is of a mixed type (ist kind with regard to the
lifting parts and 2nd kind with regard to the fuselage parts). The number of
2. 139
unknowns is less than in the case of ordinary panel methods, because on the
lifting parts we have reduced the number of unknowns with one half.
3F2 3F2 = Q i
2
(;) - ,2 = M2 - (2.6.12)
Also in the case of supersonic flow the linearized potential equation allows the
solution, by means of application of Green's theorem, to be expressed in terms
of surface singularity distributions. In this case the integral relation takes
the form
2.141
(P) =
1 1
If [o (- -)
r,
+p a
anQ
(- ir-)] dS (2.6.13)
where
r = [(x_xt)2 - 2 (yy,)2 - (z_z')
ji/2 (2.6.14)
- i
=
n = (n n, - n)
z
(2.6.15)
and x>x'
The integration has to be performed only over the part S' of S lying within the
forward Mach cone of P, resulting in a factor in front of the integral
instead of as in the subsonic case.
4ii
The asterix above the integral sign indicates that we have to take the finite
part of the integral (in the sense of Hadamard, see ref. 2.44 or 2.45) , because
on the surface of the forward Mach cone emanating from P the integrand of eq.
(2.6.13) is singular (this in addition to the singular behaviour in x=x' , y=y'
z=z').
Based on eq. (2.6.13) an integral equation may be formulated by satisfying the
boundary conditions on the configuration (normal velocity component zero). The
discretization and solution procedures are, in principal similar to those for
subsonic panel methods. (Note: due to the limited extent of the domains of
influence (see fig. 2.6.5) we no longer have full matrices!).
Almost all supersonic panel methods are based on the application of (supersonic)
thin-wing theory to lifting surfaces in combination with surface distributions
on, body-type components. The computer codes are usually organized in such a way
that subsonic as well as supersonic flows can be treated. Examples are the
earlier mentioned Woodward method (ref. 2.41), USSAFRO (ref. 2.42) and NLRAERO
(ref. 2.43). Also PANAIR (refs. 2.18 and 2.19) contains a supersonic option.
More recently a higher order subsonic/supersonic code has been developed by MBB,
(HISS-code, ref. 2.46).
In a numerical sense supersonic panel methods are less accurate than subsonic
methods. This is due to the property that 'physical' as well as numerical
discontuities are propagated undamped along the surfaces of Mach cones (fig.
2.6.6). It appears to be difficult to represent accurately the 'physical'
discontinuities, while the numerical discontinuities should, of course, be
absent. Because of the latter, higher order methods (ref. 2.46) are more
desirable in the supersonic case than in the subsonic case.
2. 142
Slender wings, with separated flow from the side edges which influences
relatively large regions of the wing, fig. 2.6.9 (see also figs. 2.4.14-19).
Wings with large sweep kdelta wings with flow separation from the (sharp)
leading edges (fig. 2.9.10).
configurations with canards, where the trailing vortices from the canard
interfere strongly with the wing flow (fig. 2.6.11).
In all these cases the rolling-up vortex sheet has a dominating influence on the
pressure distribution and aerodynamic forces acting on the wing.
In the 1980's considerable advances have been made with the modeling by means of
panel methods of such type of flows, see for example ref. 2.47.
However, the advances have not yet led to applications in every day practice for
complicated practical configurations. Reasons for this are
2.1143
A survey of panel methods having some form of capability for modeling of vortex
sheet roll-up can be found in ref. 2.47. They can be distinguished according to:
results obtained with a special version of the NLR PANEL method of the computed
shape of the vortex sheet behind a swept wing after 4 iterations.
Newton iteration represents a more general method for solving systems of non-
linear equations that may be found in text boóks on numerical analysis, see for
example ref. 2.24. In this technique the boundary conditions on the solid part
of the configuration (wing, fuselage) and those on the vortex sheet, i.e.
- zero normal velocity and
- vortex lines are (average) streamlines,
lead to a system of nonlinear algebraic equations with as unknowns the doublet
(vortex) distribution and the parameters describing the geometry of the vortex
sheet.
This nonlinear system can, symbolically, be written as
Ñ() = 0 (2.6.16)
(n) (-(n)'
x J
=R-(n) (2.6.17)
(')
n1) of may be
where is the residual vector. Then, an improved estimate
obtained as follows. In the neighbourhood of the solution the residual vector R
can be expanded in a Taylor series. For each component R. of , we write
i
aR.
dx (n)
) + o[(o x (2.6.18)
n) j
j ix=x
or, with eq. (2.6.17) and neglecting second and higher order terms
aN.
The system (2.6.19) may be solved with one of the methods discussed in section
2.2.9. If ô is known a new estimate of follows from
-(n+1)
X
(n)
=X-(n) +ôx (2.6.20)
(n+1)
and a new residual vector may be computed from eq. (2.6.17).
Note that according to eq. (2.6.18) a condition for the convergence of the
Newton procedure is that the initial estimate should be sufficiently close to
the solution, a condition which may be difficult to fulfil in practice.
We further note that an iterative solution of the linear system (2.6.19) allows
a blending of iteration techniques. For example, if the Gauss-Seidel iteration
technique is applied to eq. (2.6.19), it is possible to perform already a Newton
step after one or a few Gauss-Seidel sweeps; this is called Quasi-Newton. Even
during the Gauss-Seidel process which is carried out row by row, one can
aN.
determine, for each individual row, ---i values based on the most recent
ax.
'J
For completeness we mention that panel-type methods have also been developed for
- unsteady (periodic) flows described by Helmholtz' equation; in particular for
the purpose of prediction of unsteady airloads for flutter analysis
- evolutionary time-dependent flows
- the simulation of massively separated flows
- the prediction of flows in a rotational frame of reference (propellers,
helicopter rotars)
- the prediction of' propulsion-airframe interaction effects
- compressible potential flows (see part II)
The scope of this course does not permit to go into any detail of such
developments. The interested reader may find entry to the literature through
reference 2.0.
on f(x,z) = O (2.7.1)
= O
an
and
C(s) = g(s) on f(x,z) = O (2.7.2)
= i (2.7.3)
U2
2.147
where
(+ sign for s>s, - sign for s<s, where s represents the position of the
stagnation point (C =1) at the leading edge).
Note that the boundary condition (2.7.4) may be transformed by integration into
the Dirichiet condition
s
(s) = f h(s') ds' + (2.7.6)
o
s, =0
C <1
p-
C = 1 in one or two (stagnation) points (2.7.7)
The situation with only one stagnation point occurs for a wing section with a
cusped tail were we have a stagnation point at the leading adge and no
stagnation point at the trailing edge (fig. 2.7.16).
The conditions given above are necessary but usually not sufficient to guarantee
a well-posed inverse problem with a unique and acceptable solution. Foir the
case of two-dimensional flow it is known (ref. 2.49, 2.55) that additional
conditions must be satisfied, explicitly or implicitly to obtain a valid and
unique solution, i.e.
- a 'closure' condition (or conditions)
- a 'regularity' condition.
A closure condition is required because the Neumann condition eq. (2.7.1) by
itself does not guarantee a closed contour or a given trailing edge thickness.
For example shapes like that of fig. 2.7.lc may satisfy (2.7.1) for any width of
the trailing edge gap. If there is a gap, there is also a resulting nett mass
flow from the airfoil to infinity. The level of this mass flow and, hence, the
width of the trailing edge gap, can be controlled through the level of the
potential at the airfoil which is determined by the constant of integration in
eq. (2.7.6). Alternative ways for imposing closure (through free parameters in
the pressure distribution) are discussed in ref. 2.55.
(n)
Let f (x,z) = O, where x = x(s), z = z(s), and j-- (s) = V(s), j (s) =
Vt(s) represent a uniformly valid approximation to the solution of the inverse
problem. The slope e of the airfoil contour is then given by
dx
= cos 8
ds
(2.7.8)
dz
= sin O
ds
V
n
a = 8 + arctan (2.7.9)
t
From the requirement that the airfoil is a stream surface it follows that
V
arctan - O
t
Vn sO
i
F = J [w (c -c )2 + w (z-z )2 ds (2.7.10)
p g o
s=0 Pt
s
dÇ(s) J (-/-1 ds (2.7.11)
s
o
where dÇ(s) is the correction normal to the profile contour as a function of the
arc length.
2.151
A variation of the procedure depicted in fig. (2.7.3) has been developed at NLR
(code INSYST, ref. 2.51). In ref. 2.51 the geometry correction at each iteration
cycle is determined by means of inverse thin-wing theory type panel methods
(source-and doublet/vortex distributions on the wing reference plane). The
driving force in the inverse thin-wing module is the difference (residual)
between the required (target) pressure distribution and the pressure
distribution on a current estimate of the geometry. The latter is determined by
means of an 'ordinary1 panel method. The approach just described is called a
residual-correction method. A flow chart is shown in fig. 2.7.4.
The method allows for geometry constraints in a weighted least square sense.
Examples of' application of the NLR INSYST code are given in figs. 2.7.4/5. Fig.
2.7.4 illustrates a wing-alone case, without (a) and with (b) geometry
constraints. Fig. 2.7.5 deals with the same wing, but now attached to a body.
These examples illustrate, a.o.:
- that convergence to engineering accuracy is obtained in about 3 to 4
iterations
- the necessity to be able to apply geometry constraints.
2.152
References
2.0 Katz, J. Low Speed Aerodynamics; From Wing Theory to Panel
Plotkin, A. Methods.
McGraw-Hill, Inc., 1991
ISBN 0_07_0S0L46_6.
2.3 Kellog, 0.D. Foundations of Potential Theory, New York, Dover, 1953.
2.4 Mangler, K.W., Behaviour of the Vortex Sheet at the Trailing Edge of
196, p.22-1414.
2.153
2.12 Labrujère, Th.E. An Approximate Method for the Calculation of the Pressure
Loeve, W., Distribution on Wing-Body Combinations at Subcritical
Slooff, J.W. Speeds, AGARD CP, No 71, 1970.
2.13 Kraus, W., Das Panelverfahren zur Berechnung der Druckverteilung von
Sacher, P. Flugkörpern in Unterschallbereich, Z. Flugw., 21 Jahrg.,
Heft 9, September 1973.
2.16 Hess, J.L. Calculation of Potential Flow About Arbitrary 3-D Lifting
Bodies; Final Technical Report, Mc Donnell-Douglas Rep.
No. MDC J 5679-01, 1972. Also: Comp. Meth in Appl. Mech.
Eng., Vol. ¿4,
1974.
2.19 Johnson, F.T. A General Panel Method for the Analysis and Design of
Arbitrary Configurations in Incompressible Flow, NASA CR-
3079, 1980.
2.154
2.21 Youngren, H.H. Comparison of Panel Method Formulations and its Influence
Bouchard, E.E. on the Development of QUADPAN, an Advanced Low Order
Coopersmith, R.M. Method, AIAA Paper 83-1827, 1983.
Miranda, L.R.
2.22 Bristow, D.R. Development of Panel Methods for Subsonic Analysis and
Design, NASA CR-3234, 1980
2.23 Botta, E.E.F. Calculation of' Potential Flow Around Bodies, Ph.D.
Thesis, Groningen, 1978.
2.24 Isaacson, E. Analysis of Numerical Methods, John Wiley and Sons, New
Keller, H.B. York, 1966.
2.26 Hess, J.L. Consistent Velocity and Potential Expansions for Higher
Order Surface Singularity Methods, Mc Donnell Douglas
Rep. MDCJ 6911, 1975.
2.28 Joosen, C.J.J. Application of' the NLR Panel Method to a Swept-Wing/Body
2.31 Labrujère, Th.E. An Approximate Method for the Calculation of the Pressure
2.33 Van Dyke, M.D. Perturbation Methods in Fluid Mechanics, Parabolic Press,
Stanford, 1975.
2.36 Piers, W.J. Calculation of the Flow Around a Swept Wing, Taking into
Schipholt, G.J. Account the Effect of the Three-Dimensional Boundary
Vai-i den Berg, B. Layer; Part 1: Wing with Turbulent Boundary Layer, NLR FR
75076 U, 1975.
2.37 Oskam, B. A Calculation Method for the Viscous Flow Around Multi-
Component Airfoils, NLR TR 79097 U, 1980.
2.42 Woodward, F.A. An Improved Method for the Aerodynamic Analysis of Wing-
Body-Tail Configurations in Subsonic and Supersonic Flow,
NASA CR-2228, 1973.
2.50 Labrujère, Th.E. MAD-A System for Computer Aided Analysis and Design of
Multi-Element Airfoils, NLR TR 83136 U, 1983.
2.51 Fray, J.M.J. A Constraint Inverse Method for the Aerodynamic Design of
Slooff, J.W. Thick Wings with Given Pressure Distribution in Subsonic
Flow, AGARD CP No. 285, Paper 16, 1980.
2.54 Courant, R. & Supersonic Flow and Shock Waves. Interscience Publishers,
Friedrichs, K.O. Inc., N.Y., 1967.
Fig. 1.1
Exmp1e of d surface (computationat) grid.
Fig. 1.2
DATE uç/II/2g. TUlE
X X X s EPRUG s 05/I /29.
X
ItRPIES USERSBYTSP5A
Fig. 1.3
Examp'e of a spatia[ grid.
Fig. 1.4
$IVI$FA A
Ì 'q $J?44d_____
I
tZ
I
pi,:,.
Fig. 1.5
UPPER SURFACE ISOBARS
A1I CATE TDE
177'
J! ALPHA- .ß.4 . CQTaR DirERVN.- . TICAL LT)
Fig. 1.6
Fig. 1.7
Fig. 1.8
Hardware developments continu fo be the driver for CFO
-persistent performance trend factor 10 er 9 years
106
Q conventional computers
Performance peak
(M flops)
O massively parallel
sustained
tera lo'
peak
/ xl000
speed up
lo'
giga
sustained
102
peak scalar
10°
peak scatar
10_2 I. I
Fig. 1.9
z
Fig. 1.10
I
REYNOLDS AVERAGED
NAVIER STOKES EQS.
IN N N /
SYMPTOTIC THEORY
ZONAL MODELLING COMPLETE
INVISCID TURBULENCE
MODEL
EULER EQS.
IRROTATÇONALITY
LINEARIZED
POTENTIAL
EQ. HIERARCHY OF MATH./PHYSJCAL MODELS
Fig. 1.11
ONSET OF B.L. SEPARATION
MACH NUMBER
Fig. 1.12
Fig. 1.13
vol. e
surf. Se
Fig. 1.14
Fig. 1.15
RE-AVERAGED N-S
RELAT IVE \
COMPTL. \
EFFORT - EULER
-
i
LINEAR INVISCID
(PANEL)
I I I I
.1
1970 1975 1980 1985 1990 YEAR
(NL SITUATION)
Fig. 1.16
X
L1
LJ
o'rn.
L) I
'- NX
QE
Eo zL)o
>-LJ
>
ro
>_" >-, VI
.-
LJ L) L) ro r.j
10'
- TR-
io7 rn
C ompu t a Hon
time per flow -MO-
z
(seconds) 3-D wing-body Reynolds average N-S
10'
-WK-
io'
-FIR
1/2
10-
min
- Euler
102
min
I i
10
0.1 1 10 102 io3 io' 10'
Fig. 1.17
10
memory
(words)
1015
lo
Fut) Pot.
W Wing
Euler AC Aircraft
io
.. Cray YMP/832
l0
Nec SX-3 at NLR
Cray i
/<PaneL '
i'
J J
102
1018 1021
io3 io' lO9 1012
speed (flops)
Fig. 1.18
HOD BASE
PREPROCESSORS
(GEOMETRY/MESH FLOW SOLVER(S) POSTP ROC ESSO RS
GENERATION)
Fig. 1.19
Fig. 2.1.1
Fig. 2.1.2
V.
f (s,t)
Fig. 2.1.3
Fig. 2.1.4
or
G) b)
Fig. 2.1.5
or
Fig. 2.1.6
=;7. =f )s,t)
òn
onS
(CONTINUOUS I
Fig. 2.1.7
Fig. 2 .1.8
n g(s,t)
so
Q
òn
(st)
Fig. 2.1.9
Fig. 2.1.10
g(st) òS
(st)
ds
ÒQ
(f:gffì,,j:,4 B'
A
òSc
= f(s,t)
òn
Fig. 2.1.11
Fig. 2.1.12
Fig. 2.1.13
Fig. 2.1.14
Fig. 2.1.15
òn
P, T
Fig. 2.1.16
t
Sc
Fig. 2.1.17
Fig. 2.1.18
[P]1const. r
I s3
'j Sc
X st
Fig. 2.1.19
Fig. 2 .1. 20
Fig. 2.1.21
if
o .ònJi
Sc
Fig. 2 .1. 22
IT ò
òn òs
T Sc
Pl
Fig. 2.1.23
con st
òn
Fig. 2.1.24
Fig. 2.1.25
òn - n
ILine []const.
Fig. 2.1.26
n
/
alo, i
/
source strength per unit area
Fig. 2.1.27a
Fig. 2.1.27b
Fig. 2.1. 29
0= const.
Fig. 2.1.30
*- -
Fig. 2.1.31
Fig. 2.1. 32
tangential normal
Fig. 2.1.33
Fig. 2.1.34
Fig. 2.1.36
PzO
no Fig. 2.1.37
S iconst.
C-
Fig. 2.1.38
I z const.
Fig. 2.1.39
Fig. 2.1.40
Fig. 2.1.41
Fig. 2 .1. 42
-1
n
.Pl
i:i
/ s
Ql
(2
I
Q
\
//
Fig. 2.1.43
Fig. 2.1.44
dSO
s
t net rncisspraduction)
Fig.,2 .1. 45
B
'
òs
Fig. 2.1.46
Fig. 2.1.47
's
Fig. 2.1.48
Fig. 2 .1. 49
Fig. 2.2.1
Fig. 2.2.2
Fig. 2.2.3
Patch
Grid Point Network
Conti g u ration
Fig. 2.2.4
Fig. 2.2.5
Fig. 2.2.6
Z ref
7 'T'ref
Fig. 2.2.7
Fig. 2.2.8
Upper
IT I. 3
Body
'u Lower
I" I"
I
4 3 AiLeron
III
'u 'ulu
In Ti
2
Aft- Body
2
Mid Win
Upper
Aileron
Fig. 2.2.9
n
vector (13)
vector (2. 4) }
is the positive normal
vector defined in the
panel collocation point
pI
/4
1,
/ ..--
c 4/-
Fig. 2.2.10
Z ref
A
' ref
Fig. 2.2.11
PLANAR representations
Flat surface
Straight
Line
Fig. 2.2.15
O.UADRAT(C representations
Quadratic Surface
Plane
Quadratic
Curve Edge
Fig 2.2.16
X
- -k - unit vectors
Fig. 2.2.17
Fig. 2.2.18
Fig. 2.2.19
)
fIx)
X2 X3 X4 X5
Fig. 2.2.20
s Constant sp(ine
f(x)
f7
f6 -a f8
s u
f 5-_
u- \ slope f expressed as slope of
I
f2
tine through neighbouring values
f4
f3
X0 xl X2 X3 X' X5 X6 X7 X
TotaLLy LocaL
Discontinuous
fix)
I I
X0 X2 x3 x X5 X6 X7 Xe
Local
Continuous
Fig. 2.2.22
Least square linear spline
fix)
t N X-
fo
X0 xl \ X2
f3
X3 X4 X5 X6 x7 X0
Nearly continuous
Fig. 2.2.23
C q uadratic spline (derivative is C0 linear sptine)
f(x)
X2 X3 X' X5 X6 Xl Xe
Not recommendabLe
Continuous with continuous slope
Unstable for f0 specified along with f' at midpoint and X8: vortex/Neumann
Fig. 2.2.24
s C0 corner point least square quadratic spline (derivative is Least sq. Linear sp(ine)
f(x)
X0 Xl X2 X3 X4 X5 X6 X7 X8
e Oenerally unstable for f specified at midpoints along with any other condition:
source/Neumann ; doubtet/Oirichlet : source/Dirichlet Doublet/Neumann
Fig. 2.2.25
define f at corner points by least square
C0 least square quadratic spline
fitting parabola to 4 neighbouring values
(heavy weights on 2 closest values)
then fit parabola to f7 and resultant
fIx) corner values.
Fig. 2.2.26
[east square quadratic spline
f(x)
fi
f3
X0 X1 X2 X3 X x5 X6 X7 X0
Fig. 2.2.27
100
-9 L2-E
C
(Vt)
-8
1st ORDER METHOD
-7 I 21 ORDER METHOD
-6
-5
10
-4
-3
-2
-1 AREA 0F INTEREST
o X/C
near-optimal (curvature adapted) spaci
10-'
100 iO1 102
2
- -
NUMBER 0F PANELS
100
Fig. 2.2.28
L2-E
(Vï
st ORDER METHOD
10.2
AREA OF
INTEREST
non-optimal spacing
1°-4
100
J
101
I ill 102 N io3
NUMBER OF PANELS
Fig. 2.2.29
cP
- 0.8
-0.4
x/C
0.0
0.8
CONVERGENCE STUDY
CI-IOROWISE PRESSURE DISTRIBUTION
RAE WING TIC = .15. a= 5.0 .q= .549
1.2
Fig. 2.2.30
CI,
-0.4
-0.2
x/C
0.1 0.2 03 0.4 05 O6 0.7
0.0
f u H-S(SHEETS) (60x12) 1%
BAC
'L x H-S(SHEETS) (30x12) 3%
o NLR (60x12) 1.5%
0.6 (30x12) 4%
+ NLR
CONVERGENCE STUDY
Fig. 2.2.31
influence of da(0)
Curved panel w=2.[°i
2 remote sources dx
subtending an
angle o
This source induces
normal and
tangential velocity Error analysis for linearly
components at varying o strengths
boundary
point
curvature
effect
Error in w is 0(0) and uncouptes from
dx
Error in u is 0(82) and uncouptes from a
-!2iS?2 [ i - error =0(82)
2n dx
NoterRoles of u & w reversed for vortex panels L
curvature
Remark Flat source panels present a serious effect
problem for circular duct simulation.
Fig. 2.2.32
[p
3
-- ROBERTS (DATUM) .)
-0.8 o H-S(SHEETS (60x10) Ì
0 NLR (60x10)J
-0.4
outside
0.8
00
0.4
3rd
) order source-based method
0.8 1St
order source-based methods
Fig. 2.2.33
CALCULATED SURFACE VELOCITY DISTRIBUTIONS IN A CIRCULAR DUCT
3-DIMENSIONAL iST ORDER METHOD SOLUTION 756 EFFECTIVE SOURCE PANELS (42x18)
2 3 4
X
PLANE OF
S Y MME TRY
vo
AXIS OF SYMMETRY
Fig. 2.2.34
A N P
EED 000
[]flo
0EDD
rioDD
E1D r' r' r'
I I
i.._i L_J
+ r-i II
I
e---'
L....J LJ
I
r--iIir--, r-, o
EDDD I I I
L_J L_J L_J
I I I
L_J L._J t_J i I
Fig. 2.2.35
loo
C
E Boeing TEA23O
ni
60
.4-
C-
o
I, DIRECT METHOD
In
w
e 40
C- ITERATIVE METHOD
o.
X X
X
20
X NLR Panel
o
o 500 1000 1500 2000
Number of singularities
Comparison central processor times on CDC 6600 computer of vortex source method
for a direct and for an indirect method of solution.
Fig. 2.2.36
500
0.2
Fig. 2.2.37
z
s--
Fig. 2.2.38
0.0 08
0.006
60x12 pane's
o Pressure integr. C
(COpO.00O2 at CLO)
2
0.1 LL
Fig. 2.2.39
- - - - Corrected slon. of lin, pot. eqn.
-- -- Corrected soin, of tin, pot. eqn.
Exact( SELLS) full potential equation
Exact (NIEUWLANIJ) full potential equation
- - Solution linearised potential equation (64 contour
-1.0 -1.0 points) GUTHERT II
- Solution iinearised potential equation (68 contour
Cp p
points ) GOTHERT if local1
-0.8 -0.8
Cp*(M(ocaL:: 1)
r
-0.6 -0.6
\
-0.4 -0.4
-0.2 -0.2
0.0 00
0.2 0.4 0.6 0.8 1.0
0.2 0.2
0.6 0.4
0.6 0.6
//
0.8
I Symmetrical non-lifting aerofoil 0.8 ¡ Lifting aerofoil
1.0 1,0
auasi-etliptical aerofoit section t/c=0.17; M=0.714; cx=0° NPL 3111 aerofoit section t/crO.34 M0.667;a'1.2°
Fig. 2.3.1
DESIGN
/1OBJECTWES
PHASE I PRELIMINARY
DESIGN
/ MAJOR
( DIMENSIONS
IPLANFORM
ETC.)
L.
i
DESIGN
PHASE U DEVELOPMENT CURRENT FOCUS
/ DETAILED
/ GEOMETRY
T
/
Major phases of aerodynamk design process.
Fig. 2.4.1
// /PERFORMANCE GOALS,
CONSTRAINTS
.ThI
J I
f/ AERODYNAMIC
/ OBJECTIVES,
CONSTRAINTS
AERODYNAMIC
DESIGN CONTROL
SYSTEM
CONTROL
AERODYNAMIC
ANALYSIS
SYSTEM
-J
/GEOMETRY,
AERODYNAMICS OBTAINED
Fig. 2.4.2
r)
A
C
3
PRESSURE
WII 413 SECTION
PLOTTING 2
RA E 101,90/ THICK
SECTIONS
BOIJY PRESSURE ON WING
PLOTTING SECTION
-V
Fig. 2.4.3
Ø experiments
/
CL
0.6
/4 onset of boundary
// Layer separation
in experiment
0.4-
/o
0.2 0
-0.4
o
-06
o
,
Fig. 2.4.4
Ø experiments
C
25
0.4
02 0
of separation
CL
:I:
Fig. 2.4.5
r2
LL
Trefftz integrated
00
pressures
plane
o
¶
0.6
I
0.5
4
i
/ viscous drag
X
0.4
X
0.3 I I
'I I
02
X
0.1 X
C01(comp.)
O experiments
Fig. 2.4.6
PRESSURE
PLOT TI N G WING SECTION
SECTIONS RAE 1019I. THICK
BODY PRESSURE ON WING
PLOTTING SECT IONS 31
5
6
i
X NLR METHOD
body '6 o EXPERIMENT DFL
OEwing a6 Rec s 0.65 x iO6
0.6
C
0.4
0.2
Fig. 2.4.7
M0
3
PRESSURE
PLOTTING OEbodyr 6G
SECTIONS 2
NLR METHOD
ON WING o EXPERIMENT DFL
Rec z 065 x106
-x
Cp
-2.0 SECTION 1 -2.0 SECTION 2
X IC
SECTION 3
+1.0
,i o
M'O
bodyl6 awjnga6°
0.4
Fig. 2.4.9
PRESSURE
PLOTTING
SECT N S
4 ON WING
BODY PRESSE
PLOTTING SECTKDNS
Y
II__i,i,IIIIHIIUUIIIIUI____
Ma O4 WIflQ.body I CALCULATKDN
a .637 -wIflg alon. J
EXPERIMENT
Re.l 3x1O6
C{ 8
t6
4
o +
ßOOY SIDE TIP
Panet arrangement for a typical short range,subsonic Spanwise Lift distribution for a typical short rangesubsonic
transport configuration. transport configuration.
Fig. 2.4.10
PPSSuRE
PWTTI3
SEC
ON WING
X
Ma .04 wriQ.body
U :837 - - wig alone ) CALCULATON
EXPEPIMENT Iz
Peel 53x106
-04
/S(Cl t (TOP)
.02
.02
.04
CD
06
.Ø4
-02
.02
Wing pressure distributions for a typical short range Body pressures for a typical short range.
subsonic transport configuration. subsonic transport configuration.
Fig. 2.4.11
-
a)
WAKE POSTIOP4( --
f--- U637
t - 1ITIAL EST IMAÎE
eo
z
ST AB Ij E R
40
o
o 200 300 400 5Q0
'Y
-40
-60
b)
Fig. 2.4.12
- .T_.... - 200 panels on sting
015 C
0.12
CM U US(LAOC Sa1ER
0.12
0.06
,rnuu
.. u...
....
oe
CL
02 0.4
0.04
0.04
0.4
L
......rn
-- - - ein STPG
o - W1Th40Ut STP4O } 5pt.ACCO
0.2 0.4\ z
O o8 1.0
E
lt WA.E ES ATE IW1T4JT STING)
}CMCULATI0.
CL
-0.04 b)
Coritriution to the pitching oorent
of parts of the corifigu.ratior..
-006
-0.12 0.4
WITH STING
a) WITHOUT STINGS
DISPLACED WAKE
1
CALC
Total pitching coer.t versus tota..]. -. - 1st WAKE ESTIMATE(WITHOUT STING) J
lift. e L EXPERIMENT
-o
r 0.4
WITHOUT STING }CALCULATION
--WITH STING
EXPERIMENT
02
Ct
0.1
o
r63
200
AAA
120 160 2
08 OY
-0.1
04
O
-02
Y
c)
CI.an.ise liftdistributior. or wiri. -0.3
u u
d)
Spanvis .]]ftdistritutlon on
j ter.
Fig. 2.4.13
TOP VIEW
1500
0 200
ELLIPSOID
R= oo\yH ._(.__)
11%C 0.5%
k\oo
MOMENT REFERENCE
POINT
300
PARABOLOID
R 100 [1
BOX-TYPE ADDITION TO
FUSELAGE
Fig. 2.4.14a
(I
'h
'A
'lIIwaMffD'DDMM.DMW 1I111
Fig. 2.4.14b
Body centerline
Section A-A
Body centerline
Section A-A
Configuration WBF25
MODEL I
4/
Fig. 2.4.15
CALCULATION MODEL
CONFIGURAT!ON WBF25
WING + FLAP
1.5 /
7)<
CLW + CLF,
7
7
CLw 7
7 X WING
CLF 7
.7
7X7
1.25
7
7 -,
7
7
.7 7/
7 ,-
7
X 7
7
7- /4<
X
0.75 7
7 X
7
7
X
0.5
0.25
FLAP
O
-2.5 0 25 5 75 10
a
Fig. 2.4.16
CALCULATION MODEL
Is CALCULATION MODEL
U IB
CONFIGURATION W8F25
EXPERIMENT X n
EXPERIMENT X
CONFIGURATION WBF25
[a a7i°
0.4
1.5
0.2
--- X
L25
0.25 05 0,75
1
X 0.4 La =
X
SEPARATION AT
Xi TIP EDGE
0.2
0.75
a \
II
X
\
\
I
I
8 o
O 0.25 0.5 0.75
\ r)
(a = 87501
0.5
0480
0.4 La 1
CD
\0
0.25 La r 0480
0.2
O
O
o 0.25 05 0.75
o 0.25 05 0.75
7) '1
TE
PRESSURE SECTION
-4 u
EXPFRIMENT 7
CD
-3
CONFIGURATION WBF20
77 = 0.208 a = 4.6201
-2
X'X'
-t
-3
CONFIGURATION W8F25
77 = 0.500
-2
cP
I I I I
-2 77 = a = 4.620
SEPARATION / VORTEX
FORMATION AT TIP EDGE
Fig. 2.4.18
CONFIGURATION WBF?S
CALCULATION MODEL
L PRESXURE SECTIoNS
CR EXPERIMENT CU j
= 4.62°
-2
= 0.407 q u 0.743
I I I I i t I I I
0.2 0.4 0 0.2 0.4
o 0.2 0.4 0
X/C X/CW X / C,
Finp prennuro diWtributiona
a)
2.8 I 308
j I T
CONFIGURATION WBF25
(CI.W CLF)2 CONFIGURATION WB
2.4
0.8 (j. 0A EXPERIMENT i
2 CALO. MODEL T
LW1
0.4
1.6
044 0 0.02 0.04 0.06
CO,
f, CALCULATION MOOSL
is
-N
1.2
CONFIGURATION WBF0
EXPERIMFNT
0.8
4 4
0.4
o
0
44- 0.02 0.04 0.06 0.08 0.10 0.12 0.16 0.18
C0
0.2
b)
Fig. 2.4.19
.-j...uIU..UU ulîi;; -
I//I"
I/I/fill
'I//fl,', a) CONFIGURATION PANELING
11111111
/1/lili!
ill"
/1111111
.V///I/llu,lIf
,,,/I,I,,lI,u
ii.
IIIIuIa___
liuLIilLil_Il -a..
NACELLE OFF
CZ
rn.f.r. 2.3 NACELLE
m.f.r. 0.4 ON
2.5
/ \\
I
m.f.r. = MASS FLOW RATIO
4/
// \°-. .
2.0
b ) SPANWtSE LIFTDISTRIBUTION .
1.5
CENTRE-LINE NACELLE
1
o .2 4 .6 .8 10v
Effect of naceL'e flow ori lift (take-off configuration a = 18°) (Ref. 2.29).
Fig. 2.4.20
M..O.e cL.3
wrTH PYLOI I STO4E } EXR (Rt.14..40A1
r WITHOUT PYLOII I STORE
wrTH PYLON ¡STORE
WITHOUT PVWN/ } CALCULATION
cp
- 0.4
PYLON
-0.2
o
A
o X
4_
o X d UPPERSOC
SS
0.2
(from: Ref. 2.32)
0.4
Store pressure distributions
(Ma 0.8 30)
Fig. 2.4.21
a)GEOMETRICS OF CONFIGURATION
t etC...I.tIe., s. S a.
4..'. 3
n, .a..,e -.
cp*
- 1-
Fig. 2.4.22
F LIGHT CONF.
WINDTUNNEL CONF
LOCATION OF TUNNELWALL
Cp
a ) CONFIGURATION PANELING
o
C-
Cp
0.2
1.0 X/C
Fig. 2.4.23a
C.
RO'
ROS
75 i
C EXP
ROS RO? - CALC
O NACELLE
STAR N R O)
NACELLE PRESSURES
r
PRESSURE DISTRIBUTION
DISPLACEMENT THICKNESS
I
Fig. 2.5.1
new body
IpI=O
oriqinat body
Fig. 2.5.2
- dividing streamLine
SURFACE BLOWING
Fig. 2.5.3
cp
a 60 , Re 2.5x io6
o EXP. UPPER SURFACE REF. 33
EXP. LOWER SURFACE McOl9
- INVISCID / VISCOUS THEORY
I I i I I I I
0.1-
I I I I I I.
0 0.2 0.4 0.6 0.8 1
Fin
3.8
NLM 7301 + FLAP
3.4
INVISCID THEORY
C,
riREDICTEO
TURBULENT
f3 o BOUNDARY 3.6
'I LAYER NIR 7301 + FLAP
SEPARATION
C,
WING
UPPER
3.0 - t3.4
SURF ACE THEORY
g0-.----- F LAP WAKE
o
2.8 - 3.2 WING WAKE
2.6 -
/ 3.0
2.4 - 28
/ 0.-o
2.2 - 2.8 /
/
2.0 2.4 -o
R. 2.5 MILLION Re 2.5 MILLION
1.4e 1.8
4 B 12 18 lB 2(8) 300 400 500 500
o 100
(D EGH E E S) . 1OC
Fig. 2.54b
CONFIGURATION 2
t . . .
0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
rX
4.5
CONFIGURATION 2
Re = 2.6 MILLION
INVISCID THEORY
M= O
4.0-
O
t
3.5-
3.0-
2.5
0 2 4 6 8 10 12
- a(DEGREES)
Comparison of Lift versus angle of attack between viscous theory and
experimental data for confiquration 2 (Ret.2.0).
Fin. 2.5 S
4.0
CONFIGURATION 2
C4 Ri - 2.6 MILLION
5.
8
7. i1
3.5
VISCOUS THEORY
=0
EXP. DATA
/ M = 0.19
3.0
/6 OSPANWISE AVERAGE
/ I ISPANWISE VARIATION
I I
2.5
o 200 400 600 800
- 1O Cd
Comparison of lift-versus-drag curve between viscous theory and
experimental data for configuration 2
Ci
CONFIGURATION 2
-2
SEPARATION
Fig. 2.5.5b
A SUPERCRITICAL TRANSPORT CONFIGURATON
Fig. 2.5.6a
SPAN WISE DISTRIBUTIONS OF SECTION LIFT
COEFFICIENT ON THE WING FOR A
SUPERCRITICAL TRANSPORT CONFIGURATION
0.7
0.6
0.5
0.4
, /p
CQ
CL
0.582 - INVISCID CALCULATION
0.484 --- VISCOUS CALCULATION. DI'LACEMENT
20 40 60 80 100
PERCENT SEMISPAN
CL
cp
DISPLACEMENT
THICKNESS WAKE
NOT MODELED
Ref. 2.16)
z wìng-reference pLane SW
Fig. 2.6.1
BODY SEGMENTS
RING
tr.f
INTERNAL
LIFT -CARRY -OVER
STRIP
STRIP
Fig. 2.6.2
O experiments
/
CL "thick" wing theory
'thin" wing theory
10.8
¡0
/0
0.6
/ / onset of boundary
Layer separation
in experiment
_ L __L0
/
- 02 o
/
40 40
/
-8° 8° 12° 16°
0.2
o
-0.4
/0
-0.6
or
Fig. 2.6.3
O exp ?riments
- - - NLR AERO
CM25
0.8
0.2 "
"thick' wing theory
onset of sepaafion
0.2
0.4
Fig. 2.6.4
flow direction
sin i= 1/M_
domain of L domain of influence
dependence
characteristic cone
Fig. 2.6.5
x/C
Mach Wave
Panel Corners
Internal Wave Create Spurious
Reflection s Mach Waves Unless
Singularity Strengths
are Continuous
Fig. 2.6.6
AIRFOIL : NACA 65A004
L=36.5, Dmax=333
s = 12.0 SUP. L.E.
Yref c=8.89 SUP. T.E.
Ct = 2.0
15.0
-® y/s=.9428
10.0
.7448
-ay', w
ç,,"
- .5469_,.-
5.0
15. 5925 49.40° ..349
-N,
- ------I---- ref
5.0 10.0 15.0 25.0 30.0 35.0 40.0
Fig. 2.6.7a
PRESENT METHOD
O UPPER SIDEÌ
EXP. REF. 1451
D LOWER SIDEI
Fig. 2.6.7b
AIRFOILS FLAT PLATE
L -54.86 Dm = 3.048 /168 PANELS ON BODY
WING s = 597 c cr7.42 c=l .45 25 PANELS PER WING SURFACE
TAIL s = 4.62 Cr =7.09 Ct=O.O 20 PANELS PER TAIL SURFACE
Az rei
lo - lo
«:p= 00 (= 450
10 rei
YreI
WING
L_HINGE LINE WING
i/ill
&iia TAIL
6.86 F221
"la
lo
23.21
20
'II 30 40
MOMENT REFERENCE POINTx
re (=28.87
60
Fig. 2.6.8a
6pitch = angle of wing/fuse'age
0
-- pitch
=100
PRESENT
METHOD
O = Ø0 EXP. REF.1471
o =100
LO
CA.CA1' 001
A CACA«) = 0°'
.2 -
/ D
.2
'D
00
o 2 4
( .'l
D
_,zj
2
I
4
o
M 1,5
Fig. 2.6.8b
Fig. 2.6.9
Fig. 2.6.10
Fig. 2.6.11
"WAKE RELAXATION"
/
COMPUTE SOLUTION
I,
IMPROVED WAKE ESTIMATE
I
Fig. 2.6.12
i
30
1'te
CONSTANT CHORD
CL :0501
o CL :082
..________tALCuLA0OM
.0
-025
000000000 o
0 025 050 075 10 0
2 _-025 "
025
4 o
O
0000 00 oc o
Fig. 2.6.13
a) THREE_DIMENSIONAI VIEW
56
D cx' xijv*'cL 57
- vO'Sc
Fig. 2.6.14
= O, Cp f(s)
a)
stagnation point
cusp
ç- stagnation points
/
b)
C open tait
c) d)
Fig. 2.7.1
1,400
1.200
1.000
0.800
0.600
0.4 00
0.200
0.000
- 0.200
-0.400
-0.600
-0.800
-1.000 ITERATION STEP
- 1 .200
- 1.400
-1400 -1.000 -0.600 -0.200 0.200 0.600 1.000 1.400
-1.200 -0.800 -0.400 0.000 0.400 0.800 1.200
-3 -
ITERATION STEP
cp
-2 - Q TARGET
0-
Fig. 2.7.2
/ Cp(s)-..èj
/ os
/ STARTING
/ GEOMETRY
SOLUTION OF
WIRICHLET)
BOUNDARY VALUE
PROBLEM
CORRECTION
ON
GEOMETRY
NEW GEOMETR'Y
Fig. 2.7.3
I START GEOMETRY
/ C TARGET
/ WEIGHT FACTORS
NON PLANAR
PANEL METHOD
/ Cp PANEL
DETERMINATION
RES IDUAL
I
fi Cp TARGETCp PANEL
I
(LINEAR)
INVERSE METHOD
NEW GEOMETRY
Fig. 2.7.4
lop ìsw and p.n,I .rrang.msnt of .x.mpl. win;
C, - C,
If. .347
q .075
q. 57S
q. .274
Ip. .274
q. .012
q. .012
.) HIGH WEIGHT QN TRAILING - EDGE ANGLE I0) b) EFFECT OF CONSTRAINTS ON THICKNESS AND TWIST
LOW WEIGHT 034 LEADING-EDGE RADIUS I2.5)
ZERO WEIGHT 034 THICKNESS AND TWIST
Fig. 2.7.5
Top end front view and panel arrangement of wingbody configuration (body at
zero angle of attack)
.847
'wET1 - .847
'NETT
NETT -
'tNETT
Fig. 2.7.6