Sie sind auf Seite 1von 10

Copyright 1996, American Institute of Aeronautics and Astronautics, Inc.

AIAA Meeting Papers on Disc, January 1996


A9618032, AIAA Paper 96-0056

Calculation of multielement airfoil flows, including flap wells


Tuncer Cebeci
California State Univ., Long Beach

Eric Besnard

Downloaded by UNIVERSITY OF ILLINOIS on April 14, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.1996-56

California State Univ., Long Beach

Hsun H. Chen
California State Univ., Long Beach

AIAA, Aerospace Sciences Meeting and Exhibit, 34th, Reno, NV, Jan. 15-18, 1996
A calculation method for multielement airfoils based on an interactive boundary-layer approach using an improved
Cebeci-Smith eddy viscosity formulation is described. Results are first presented for single airfoils at low and
moderate Reynolds numbers in order to demonstrate the need to calculate transition for accurate drag polar
prediction and the ability of the improved Cebeci-Smith turbulence model to predict flows with extensive separation
and therefore to predict maximum lift coefficient. Results, in terms of pressure distributions and lift and drag
coefficients, are presented for a series of two-element airfoils with flaps or slats. The method is extended to the
computation of configurations with flap wells. Results show that the same accuracy can be reached as for faired
geometries. The importance of compressibility effects and the turbulence model for stall, and the need to calculate the
onset of transition, are demonstrated. Recommendations are made for the preferred approach to predicting the
aerodynamic performance of multielement airfoils for use as a practical and efficient design tool. (Author)

Page 1

CALCULATION OF MULTIELEMENT AIRFOIL FLOWS,


INCLUDING FLAP WELLS
by
Tuncer Cebeci*, Eric Besnard" and Hsun R Chen***
Aerospace Engineering Department
California State University, Long Beach

Downloaded by UNIVERSITY OF ILLINOIS on April 14, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.1996-56

Abstract
A calculation method for multielement airfoils
based on an interactive boundary-layer approach
using an improved Cebeci-Smith eddy viscosity
formulation is described. Results are first presented
for single airfoils at low and moderate Reynolds
numbers in order to demonstrate the need to calculate
transition for accurate drag polar prediction and the
ability of the improved Cebeci-Smith turbulence
model to predict flows with extensive separation, and
therefore to predict maximum lift coefficient, (c^m^.
Results, in terms of pressure distributions and lift and
drag coefficients, are presented for a series of twoelement airfoils with flaps or slats. The method is
extended to the computation of configurations with
flap wells. Results show that the same accuracy can be
reached as for faired geometries. A slight
compressibility effect was accounted for by
introducing compressibility corrections to the HessSmith panel method. Again, the importance of the
compressibility effects and the turbulence model on
stall, and the need to calculate the onset of transition,
are demonstrated. Recommendations are then made
for the preferred approach to predicting the
aerodynamic performance of multielement airfoils for
use as a practical and efficient design tool.

1. Introduction
In recent years, there have been significant
accomplishments in computational fluid dynamics.
Whereas in the early 1960's calculations performed
for airfoil flows with panel methods and boundarylayer methods were under development for simple
flows, today the calculations are being performed
routinely with Navier-Stokes methods not only for
airfoil flows but also for complex aircraft
Professor and Chair, AIAA Fellow
Graduate Student
**
Professor, AIAA Associate Fellow
Copyright 1996 by the American Institute of Aeronautics and
Astronautics Inc. All rights reserved

configurations. Our capabilities in aerodynamic flows


have reached levels which were difficult to imagine in
the early 1960's. Progress has been so great and so
rapid that one may even say that computational fluid
dynamics has reached, or is very near to reaching, its
maturity.
Despite these advances, however, there is still
more to be done in developing design methods for
high lift configurations. The presence of high and
low Reynolds number flows on various components of
airfoils, including flap wells and significant regions of
flow separation near stall conditions as well as
possible merging of shear layers, offer significant
challenges to code developers. The required generality
and accuracy of the code and, equally important, its
efficiency as a design tool, add additional challenges.
The current development of design algorithms for
high-lift devices follows two approaches.
One
approach pursues the solution of the Navier-Stokes
equations with structured .and unstructured grids.
Some of the Navier-Stokes methods used for this
purpose are based on the solutions of the
incompressible flow equations [1, 2] while others are
based on the solutions of the compressible flow
equations [3]. The incompressible form of the
equations employing the numerical procedure of
Rogers et al. [1] provides accurate results and allows
the calculations to be performed efficiently. The
solution of the compressible equations, on the other
hand, takes a considerable amount of computer time;
their convergence rate is slower and, at this time, they
are in the research stage.
The other approach for developing methods for
high lift configurations is based on the interactive
boundary layer theory, which involves interaction
between inviscid and boundary-layer equations [4, 5].
In this approach the inviscid flow is often computed

by a panel method with or without compressibility


corrections, and the viscous flow is computed by a
compressible boundary-layer method. This approach
is very efficient but is not as general as the NavierStokes approach for more complicated geometries.

Downloaded by UNIVERSITY OF ILLINOIS on April 14, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.1996-56

Regardless of which approach is used to develop


a computational tool for high-lift configurations, it is
necessary to include the compressibility effect if the
calculation method employs incompressible flow
equations, and to calculate the onset of transition.
Both can strongly influence the accuracy of the
calculation method. Experiments on single or
multielement airfoils show that the compressibility

effect, even at Mach numbers around 0.30, has a


pronounced effect on the maximum lift coefficient.
Similarly the predicted location of transition in the
calculation method is important in properly
identifying the effects of wind tunnel and flight
Reynolds numbers. Individual components of multielement airfoils at wind tunnel Reynolds numbers can
experience relatively lower Reynolds numbers than
the main airfoil. At chord Reynolds numbers less than
500,000, the components can have large separation
bubbles, with the onset of transition occurring inside
the separation bubble [6]. As a result, the behavior of
the flow can be significantly different from the
behavior of the flow on the main airfoil at higher
Reynolds numbers. Furthermore, the transition
location can drastically influence the drag coefficient,
and therefore determining the onset of transition is
crucial for predicting drag polars, which are of major
interest to aircraft designers.
Section 2 describes an interactive-boundary-layer
approach to the calculation of high lift configurations
in two-dimensional flows. It is applied to a variety of
high lift systems, including airfoils with flaps and
slats, and airfoils with and without flap wells. Results
are presented and discussed in Section 3. Finally, in
view of the present results, a preferred approach for
predicting the flow about multielement airfoils is
discussed in Section 4.

the specification of circulation. Multielement


configurations computed with the calculation method
employing the multielement panel method with
viscous corrections showed that, while the results
were in good agreement with experimental data for
conventional airfoils, the results for airfoils with thin
trailing edges and/or supercritical airfoils were not
To improve the results, two additional options for the
vorticity distribution were incorporated into the panel
method. One option assumes that vorticity varies
quadratically with the surface distance and the second
option assumes that it varies quadratically in a small
region near the trailing edge. In this latter option, to
be referred to as the partially parabolic vorticity
distribution option, the size of the region is to be
specified in the panel method. The compressibility
correction depends upon the linearized form of the
compressible velocity potential equation and is based
on the assumption of small perturbations and thin
airfoils. The correction used in the present panel

method is based on the Prandtl-Glauert formula as


described in [7].
2.2 Inverse Boundary Layer Method
Boundary layer equations

The compressible boundary layer equations


(mass, momentum and energy) for laminar and
turbulent flows are well known and, with the
algebraic eddy viscosity (em) and turbulent Prandtl
number (Prt) formulation of Cebeci and Smith [7],
they can be expressed as follows:

(1)

i^pvir-p.u.rr-+iHGi-hp6.)ir <2>
where T is the temperature, H is the total enthalpy
given by
2

(4)

= c.T+

2. Calculation Method
2.1 Inviscid Method
The inviscid flow field is computed by the Hess
Smith panel method [7]. Three options were used in
the multielement panel method. As in a single airfoil

panel method, in one option the vorticity strength was


taken to be constant on all panels, and a single value
was adjusted to satisfy the condition associated with

and
pv=pv+pV
(5)
In the absence of mass transfer, the boundary
conditions for an adiabatic surface are
y = 0, u = 0, v = 0,

an

=0
(6a)
dy
y->oo, u->u.(x),H->H.
(6b)
In the wake, where a dividing line at y = 0 is
required to separate the upper and lower parts of the

inviscid flow, in the absence of normal pressure


gradient, the boundary conditions at y = 0 become
da
y=0,

= 0, v = 0
dy

(7)

modified. First, a fairing is placed in the flap well


region, and the Hilbert integral formulation is used to
generate initial guesses for the displacement thickness
distribution. Then, the relaxation formula

Downloaded by UNIVERSITY OF ILLINOIS on April 14, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.1996-56

Solution procedure

(5T'=(5')'

The above equations are first expressed in


transformed coordinates. Two sets of transformed
coordinates are used, one for the direct problem when
the equations are solved for the prescribed pressure
distribution, and the other for the inverse problem
with the external velocity updated during the
iterations. The Falkner-Skan transformation is used in
the direct mode and a modified version is used in the
inverse mode. Reference 7 presents and discusses the
transformed equations and their solution procedure in
detail. The airfoil is divided into upper and lower
surfaces. For each surface, the calculations start at the
stagnation point and proceed in the standard mode up
to a certain specified x-location. Then, the inverse
calculations begin from that location and continue up
to the far wake.
Interaction law

The boundary layer equations become singular at


flow separation and do not allow the calculation of
separated flows for a prescribed velocity distribution.
However, when the external velocity is treated as an
unknown, this difficulty can be overcome as discussed
in [7J. The external velocity is represented by
u e (x) = ue(x)+6ue(x)
(8)
where u is the inviscid velocity computed by the
panel method and 8ue is the perturbation velocity due
to viscous effects, which is given by the Hilbert
integral
6ue(x) = -

x-a

where 8* is actually the sum of the distance from the


fairing to the flap well cove and the displacement
thickness measured from the fairing, is used in the
inverse method to replace the Hilbert integral
formulation of the external boundary condition. The
new edge boundary conditions are given by Eq. (6b)
and Eq. (10), where u(V and u^ correspond the
external velocities computed by the boundary layer
and inviscid methods, respectively, and CD is a
relaxation parameter. At the end of the flap well, the
solution procedure reverts to the Hilbert integral
approach.
Once the boundary layer development is
computed, the blowing velocity distribution is
determined from
v B =-(u e . 5 ')

The flow calculations in the flap well region


which are similar to those for a backward facing step
require modifications. A large portion of the flow
separates immediately after the sudden change of the
airfoil geometry. The flow reattaches and gradually
recovers downstream in the flap well region or in the
wake. The calculation of such flows is difficult, and
potential theory is not adequate because of the

singularity that occurs at the geometry discontinuity


and the strong viscous effects in the separated region.
To compute such flows, the interaction law has to be

(11)

and used as a boundary condition in the inviscid


method. This interactive procedure is repeated until
convergence of the lift coefficient is achieved.
Turbulence model

An improved version of the algebraic eddyviscosity formulation of Cebeci-Smith is used here.


The eddy viscosity distribution across the boundary
layer is defined by two separate expressions,
(O, =[o.4y[l-p-^| My.

(12)

S- =

(O. =4l( u . -

(9)

in the interaction region (a, b). Its evaluation is


described in detail in [7].

(10)

1-HDP^-

where
0.0168
-1.5

A = 26i

-T

(13a)

where F is related to the ratio of the product of the


turbulence energy by normal stresses to that by shear
stress evaluated at the location where the shear stress
is maximum. As discussed in [7], it is given by
where

the

parameter

Rt=twA-puV)mtx,
represented by

which,

is

function

of

for t w 0,

is

R.^LO

Downloaded by UNIVERSITY OF ILLINOIS on April 14, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.1996-56

1+2R,(2-R,)
1+R,

(15)
R.

For TW 0, R, is set equal to zero.


Also, whereas in the original Cebeci-Smith eddy
viscosity formulation the intermittency expression was
valid only for zero pressure gradient flows, the
expression in the improved formulation is applicable
for flows with favorable and adverse pressure
gradients as well as zero pressure gradient flows. It is
based on Fiedler and Head's correlation [8] and is
given by
1
1-erf
(16)
where Y and a are general intermittency parameters
with Y denoting the value of y where y = 0.5 and CT
denoting the standard deviation. For details, see [7].
The condition used to define yc is the continuity
of the eddy viscosity, so that e m is defined by (em)(
from the wall outward (inner region) until its value is
equal to that given for the outer region by (sm)0.
The expression fa models the transition region
and is given by
Jdx
= l-exp -G(x -X*)J

(17)

Here xtt denotes the onset of transition and G is


defined by
u

(18)

where C is 60 for attached flows and the transition


Reynolds number is R^ = (u e x / v),, . In the low
Reynolds number range from R c = 2 x l 0 5 to
R c = 6 x 103 , the parameter C is given by
C2 = 213 (logR^ -4.7323J
(19)

The corresponding expressions for the eddyviscosity formulation in the wake are
(20)
where (e m ) u is the eddy viscosity at the trailing edge
computed from its value on the airfoil and (en)w is
the eddy-viscosity in the far wake given by the larger
of

u
= 0.064 J(u e -u).dy
and

(21a)

):= 0.064 J(u e -u).dy


(21b)
y*.
with y^ denoting the location where the velocity is
minimum.
For high Reynolds numbers, if the flow is
attached, the onset of transition is determined by
Michel's criterion as described, for example, in [7]. At
high angles of attack, the flow separates downstream
of the pressure peak before Michel's criterion can be
satisfied. Therefore, the onset of transition is chosen
to coincide with laminar separation.
At Reynolds numbers less than 106, where large
separation bubbles may be present even at low angles
of attack, it is necessary to calculate the onset of
transition by the en-method as discussed in [7]. For
high lift configurations, flaps usually have low
Reynolds numbers. Fortunately, the pressure
distribution on the flap upper surface does not vary
greatly with angle of attack, and therefore the
computation of the transition location does not have
to be performed at each angle of attack.
3. Results and Discussions

3.1 Single Airfoils

Before we present a sample of results for high lift


configurations, it is useful to discuss the role of
transition by examining the results shown in Figs. 1
and 2 for two airfoils at high and low Reynolds
numbers. The results in Fig. 1 are for the NACA
0012 airfoil at a chord Reynolds number of 3 x 106.
As can be seen, there is a significant difference
between the calculated lift coefficients in which the
transition location was fixed near the stagnation point
for all angles of attack or computed. A similar
observation can be made for the drag coefficients.
Those calculated with the transition computed show a
much better agreement with experimental data than
those in which the transition was fixed.
Fig. 2 shows the results for the Eppler airfoil at a
chord Reynolds number of 200,000. Here, the location
of transition does not play an important role in

predicting lift coefficients (Fig. 2a). However, the


calculations with a fixed transition location can be
performed until a = 13.5. As before, the drag
coefficients obtained with the transition location
computed show better agreement with data than those
with a fixed transition location (Fig. 2b).

3.2 Two-element Airfoils without Flap Well

A sample of results is presented below for three


two-element configurations without flap wells.
Figs 3 and 4 show the results for the MLR 7301
supercritical
airfoil/flap
configuration.
The
experimental data of Van den Berg and Oskam [9]
include pressure distributions at a = 6 and 13.1, a

lift curve and a drag polar. A flap of 32% chord was


0.0

3.0

10.0 15.0 20

Downloaded by UNIVERSITY OF ILLINOIS on April 14, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.1996-56

0.040

0.030

meaturcmenU

transition computed
traniitioo it the
leading edge

0.020
0.010
0.000
0.0

0.5

1.0

used at a deflection angle of 20 and with a gap of


2.6% chord. The experimental freestream Mach
number was M = 0.185, and the chord Reynolds
number was 2.51 x 10*. Fig. 3 shows a comparison
between measured and computed pressure
distributions at a = 6.0 and a = 13.1, and Fig. 4
shows a similar comparison for the lift coefficient and
drag polar. The inviscid flow calculations were
performed using the partially parabolic vorticity
option, and the viscous flow calculations were
performed with the onset of transition location
calculated on the main element and flap. As can be
seen from the results in Fig. 4a, the calculated results,
including drag, agree well with experimental data.

Fig. 1. Force coefficients for the NACA 0012 airfoil at


Re = 3 x 10*, (a) lift and (b) drag.
o Measurements

1.5 r

Calculations

o measurements
- transition calculated
transition it the leading edge
0.0

-2,0

3.0

8.0

13.0

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4


x/c

1S.O

(a)

0.10
0.08

0.06
0.04
0.02

-2.5
0.00

-5.0

0.0 0.2

0.0

(b)

5.0

10.0 15.

Fig. 2. Force coefficients for the Eppler airfoil at Re '

200,000, (a) lift and (b) drag.

(b)

0.4 0.6 0.8


X/C

1.0 1.2 1.4

Fig. 3. Pressure distribution for the NLR 7301


supercritical airfoil/flap configuration for (a) a * 6* and
(b) a-13.1'.

3.5 r

6), which causes a greater predicted flow acceleration


on the flap upper surface than the actual acceleration.

oMouurcmenti

-Cp

0.0

2.0 -1

20.0

Downloaded by UNIVERSITY OF ILLINOIS on April 14, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.1996-56

(a)
-2.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
X/c

0.10 r

20.0
16.0

Critical Pretsure Coefficient

12.0
8.0
4.0
0.00

0.0

(b)

'

Fig. 4. Force coefficients for the NLR 7301 airfoil/Hap


configuration, (a) lift coefficient, and (b) drag polar.

Figures 5 to 9 show the results for two NASA


high lift configurations tested by Omar et al. [10] for
a freestream Mach number of Mo = 0.201 and chord
Reynolds number of 2.83 x 106. As in the NLR 7301
multielement configuration, the inviscid flow
calculations were performed using the partially
parabolic vorticity option. In addition the
compressibility corrections were made to the inviscid

-4.0
0.0 0.2 0.4 0.6 0.8

(b)

Fig. 5. Pressure distribution for the NASA airfoil/flap


configuration at (a) a - 0.01* and (b) a 8.93*.
wake center Una

flow using the Prandtl-Glauert formula.


Figure 5 shows the pressure distributions for the

meuured merging

airfoil/flap configuration at a = 0.01 and 8.93. It is


seen that the calculated results are in good agreement
with experimental data, although there is some
discrepancy on the upper surface of the flap. This
discrepancy is due to two phenomena, the merging of
shear layers and the inaccurate wake center line
prediction, as shown in Figure 6. It is seen that in the
experiments, the merging occurs at the trailing edge
of the flap for both a = 0 and a = 8, and that it is
not predicted by the present method. Also, in the
present method, the wake center line is assumed to be

the dividing streamline merging from the main


element trailing edge and is closer to the flap upper
surface than the measured one (thick line in Figure

1.0 1.2 1.4

Fig. 6. Comparison of measured and computed shear


layers for the NASA airfoil/flap configuration, (a) a - 0*
and (b) a - 8*.

Downloaded by UNIVERSITY OF ILLINOIS on April 14, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.1996-56

Figure 7 shows the lift and drag coefficients for


the same configuration. The lift coefficient is slightly
over predicted due to the discrepancies on the flap
upper surface, and similarly the drag coefficient is
slightly under predicted. Also, the calculated
incompressible stall angle is rather different from the
calculated compressible one. The critical pressure
coefficient for M = 0.201 is indicated in Fig. 5b. At
a = 8.93, the measured stall angle, there exists a
small region of supersonic flow and a shock may
occur. Even though this shock cannot be predicted by
the panel method, the compressibility corrections that
are applicable at lower angles of attack provide a
significant improvement to the stall prediction.
3.0 r

26.9. This gives a 0.7% error for the maximum lift


coefficient and an error of 2.2% for the stall angle.
18.0 r
14.0

10.0
-Cp

6.0
2.0
-ZO

-0.2

0.0

0.2

0.4

0.6

0.8

1.0

x/c
Fig. 8. Pressure distribution for the NASA slat/airfoil
configuration at a = 26.88*.
3.0 r

2.0

Measurements
Incompressible
Compressible

1.0

-10

10

Skt

15

00

0.0

10.0

0.15 r

15.0

20.0

23.0

30.0

a
0.130 -

0.10

0.125

o Mcurananti
Cilcubtion

0.100

Cd

0.075

0.05

0.050 0.00
0,0

(b)

2.0

1.0

0.025 r

3.0

Figures 8 and 9 show the results for a slat/airfoil


configuration. Figure 8 presents a comparison
between measured and calculated pressure
distributions at 26.88. While the agreement is
excellent for the slat, the pressure peak on the main
airfoil is slightly under-predicted. Figure 9 shows the
lift and drag coefficient comparisons for the same
configuration. The agreement with data is excellent
up to stall for the lift coefficient and the drag polar is
predicted satisfactorily. The computed maximum lift
coefficient and stall angle are 2.88 and 27.5,
respectively, and the measured ones are 2.90 and

o)
^^S^

n tm\

Fig. 7. Force coefficients for the NASA airfoil/Hap


configuration, (a) lift coefficient, and (b) drag polar.

(b)

0.0

1.0

ZO

3.0

Fig. 9. Force coefficients for the NASA slat/airfoil


configuration, (a) lift coefficient and (b) drag polar.

3.3 Two-element Airfoil with Flap Well

The method described in Section 2 was applied to


an airfoil/flap configuration tested by Omar et al. [10]
similar to that of the previous section, but comprising
a flap well cove to illustrate the applicability of the
method to airfoils with flap wells.
As mentioned in Section 2, the calculations are
performed in several steps. First, a fairing is assumed
in the flap well region. Inviscid calculations are then
performed for this faired geometry and an external

velocity distribution is obtained. The boundary layer

o.oio

equations are solved over the airfoil for this pressure


distribution with the Hilbert integral inverse
formulation of Section 2 so that an initial
displacement thickness in the flap well region can be

Downloaded by UNIVERSITY OF ILLINOIS on April 14, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.1996-56

a-8.59

Cf

obtained. Then, in the flap well cove, the inverse


formulation employing the relaxation formula of Eq.
(10) is used starting with this initial displacement
thickness distribudon added to the distance from the
flap well cove to the fairing. After several iterations, a
new displacement thickness distribution in the flap
well cove is determined. The corresponding
distribution over the faired airfoil is calculated and
simulated into the inviscid method with the blowing
velocity. As in the cases without flap well, several
inviscid/viscous iterations are performed until
convergence.
Figure 10 presents the geometry in the flap well
region and the fairing used for the calculations.

a--8.11

0.005

dement

0.000

-a 0050.65

trailing edge

0.70 0.75 O.K

0.85

x/o

Fig. 11. Computed skin friction coefficient in the flap

well region for a - -8.11* and a - 8.59 *.


measurements
with flap weflcalc.
- without flap well calc.
Cp

Results in terms of computed displacement thickness

are shown for a = -8.11 and a = 8.59. As expected,


the recirculation region is much greater for a = -8.11

than for a = 8.59. Figure 11 presents the


0.5 0.6 0.7 0.8 0.9
Xfc

corresponding skin friction coefficient variation in the

flap well region and demonstrates the ability of the

1.0 1.1 1.2

method to compute large regions of separated flow.

Cp

0.5 0.6 0.7 0.8 0.9

1.0

1.1 1.2

X/c

Cp

Fig. 10. Computed displacement thickness in the flap


well region for (a) a - -8.11* and (b) a = 8.59 *.

Figure 12 shows a comparison of measured and


computed pressure distributions in the neighborhood

of the flap well cove and on the flap for a = -8.11, a


= -0.02 and a = 8.59. Comparing measurements
with those for a smooth geometry (Fig. 5), it is seen
that the presence of the recirculating flow essentially
causes an almost constant pressure distribution in the
flap well cove and increases the flow velocity in the
vicinity of the flap leading edge.

0.5 0.6 0.7 0.8 0.9

1.0 1.1 1.2

(c)

Fig. 12. Comparison of measured and computed


pressure distributions hi the flap well neighborhood for
(a) a - -8.11*. (b) a - -0.02* and (c) a - 8.59 *.

Downloaded by UNIVERSITY OF ILLINOIS on April 14, 2015 | http://arc.aiaa.org | DOI: 10.2514/6.1996-56

Figure 12 shows that flap well results are better


than those obtained with the fairing which ignored
the presence of the flap well. On the main element,
while the agreement is satisfactory for a = -8.11, the
region of recirculating flow is under predicted for a =
-0.02 and a = 8.59. This may be due to several
factors, such as the inappropriateness of the
interaction law and of the turbulence model for such
recirculating flows. Work is in progress to improve
the results. However, the pressure distribution on the
flap upper surface, in particular close to the leading
edge, is satisfactory, which suggests that the
displacement thickness prediction is accurate near the
trailing edge of the main element.
Lift and drag coefficients were computed but are
not shown here since they are essentially the same as
for the case without flap well (Fig. 7).
4. Concluding Remarks

An interactive boundary-layer method with an


improved Cebeci-Smith eddy viscosity formulation is
used to calculate the aerodynamic performance
characteristics of high lift devices. The results show
good agreement with measurements for a variety of
multielement airfoils comprising slats, main element
and flaps. The method has also been extended to the
computation of airfoils with flap wells, and the results
show the same overall level of accuracy. In general,
the predictions are excellent for relatively low angles
of attack and very satisfactory up to stall. The study
shows that the onset of transition location plays a
significant role in predicting drag and that its
calculation must be a part of the computational
method. The study also shows that for some flows
merging of shear layers takes place and, since the
present method does not account for this merging, the
calculated results near stall conditions, while
satisfactory, are not as good as those at lower angles
of attack.
The stall prediction is satisfactory for most cases.
However, when the pressure peak on the main
element reaches values close to or greater than the
critical pressure coefficient, a truly compressible
inviscid method, such as a full potential or Euler
method, should be coupled with the present
compressible boundary layer method in order to
compute stall with reliability. Work is in progress to
incorporate these capabilities, including the merging

of shear layers.

References

1. Rogers, S.E., Wiltberger, N.L. and Kwak, D.,


"Efficient Simulation of Incompressible Viscous
Flow Over Single- and Multielement Airfoils,"
AIAA Paper No. 92-0405,1992.
2. Rogers, S.E., "Progress in High-Lift Aerodynamic
Calculations," AIAA Paper No. 93-0194, Jan.
1993.
3. Valarezo, W.O. and Mavriplis, D.J., "NavierStokes Applications to High-Lift Airfoil Analysis,"
AIAA Paper No. 93-3534, Aug. 1993.

4. Cebeci, T., "Calculation of Multielement Airfoils


and Wings at High Lift," AGARD Conference
Proceedings 515 on High-Lift Aerodynamics,
Paper No. 24, Banff, Alberta, Canada, Oct. 1992.
5. Cebeci, T., Besnard, E. and Messie, S., "A
Stability/Transition-lnteractive-Boundary-Layer
Approach to Multielement Wings at High Lift,"
AIAA Paper No. 94-0292, Jan. 1994.
6. Cebeci T, "Essential Ingredients of a Method for

Low Reynolds-Number Airfoils," AIAA Journal,


Vol. 27, pp. 1680-1688, 1983.

7. Cebeci, T., An Engineering Approach to the


Calculation of Aerodynamic Flows, to be
published, 1996.
8. Fiedler, H. and Head, M.R., "Intermittency
Measurements in the Turbulent Boundary Layer,"
J. Fluid Mech., Vol. 25, Part 4, pp. 719-735, 1966.

9. B. Van den Berg and B. Oskam, "Boundary Layer


Measurements on a Two-Dimensional Wing with
Flap and a Comparison with Calculations,"
AGARD CP-271, Sept. 1979.

10. E. Omar, T. Zierten and A. Mahal, "TwoDimensional Wind Tunnel Tests of a NASA
Supercritical Airfoil with Various High Lift
Systems," NASA CR-2215, 1977.

Das könnte Ihnen auch gefallen