Vous êtes sur la page 1sur 35

A COMPUTER PROGRAM FOR THE ANALYSIS OF MULTIELEMENT AIRFOILS IN TWO-DIMENSIONAL SUBSONIC, VISCOUS FLOW By Harry L. Morgan, Jr.

NASA Langley Research Center SUMMARY A computerized analytical model which computes the performance characteristics of multielement airfoils in subsonic, viscous flow has been developed under NASA contract to the Lockheed-Georgia Company. The model computes the viscous pressure distributions, lift, moments, and local boundary-layer properties on each element of an arbitrarily arranged slotted airfoil in attached flow. The final viscous solution is obtained by an iterative technique for successively combining an inviscid solution with boundary-layer displacement thicknesses. The surface of each airfoil element is approximated as a closed polygon with segments represented by distributed vortex singularities. The ordinary boundary-layer solution is comprised of mathematical models representing state-of-the-art technology for laminar, transition, and turbulent boundary layers. An additional boundary-layer model has been incorporated to compute the characteristics of a confluent boundary layer which reflects the merging of the upper-surf ace boundary layer with the slot efflux. This computer program has been used extensively at Langley and throughout the industrial and academic communities for both the design and the analysis of airfoils. Presented in this paper are summary descriptions of the general operation and capabilities of this program and a detailed description of the major improvements that have been made to the program since its initial formulation. Sample comparisons between theoretical predictions and experimental data are presented for several types of multielement airfoils. Areas of agreement and disagreement are discussed with recommendations for areas of needed program improvement. INTRODUCTION During the initial design phase of an airfoil the effects of various modifications can be easily evaluated by describing the potential (inviscid) flow around the airfoil. Many methods are available to compute the potential flow and most generally require rather small computer storage and execution times, which make them very desirable during the initial trial-and-error design phase. However, during the final design phase a more
713

accurate assessment of the selected modification can be evaluated by describing the viscous (viscid) flow around the airfoil, which means simply that the boundary-layer properties have been included in the analysis. No general, mathematically closed form solution presently exists which describes the viscous flow around an airfoil. Until very recently, airfoil design has relied mainly on potential-flow theory to obtain the theoretical pressure distribution from which the boundary-layer properties were approximated, with the interrelationship between the two being largely ignored. However, the use of iterative techniques has provided a practical solution for defining the interrelationship between the potential and viscous flows around an airfoil. State-of-the-art technology in the areas of potential flow and boundary layers, plus the availability of high-speed large-capacity computers, has provided the background capability essential to the formulation of computer codes to compute accurately the performance of an airfoil in viscous flow. One such computer code used extensively at the'Langley Research Center and throughout the industrial and academic communities was developed by the LockheedGeorgia Company under NASA Contract NAS 1-9143 in 1969. This computer program, entitled "2-D Subsonic Multi-Element Airfoil Program," was formulated to handle singleelement and multielement airfoils (a maximum of four elements) in subsonic, viscous flow. A complete description of this original program is given in detail in. reference 1. Since the initial formulation of this program, there has been a continuing effort to improve and modify the program to handle an ever increasing range of airfoil geometries. The initial debugging and general program maintenance and improvements have been performed through in-house efforts at Langley. The only other major contribution to the program improvement has been the work done by Delbert C. Summey and Neill S. Smith under NASA Grant NGR 34-022-179 to North Carolina State University. Their work was concerned primarily with the single-element version of the program and consisted of program modifications to improve the lift and drag predictions, to reduce .the computer storage requirement, and to reduce the computer execution time. A detailed description of their work is presented in reference 2. . . A summary description of the general operation and capabilities of the airfoil program and detailed descriptions of the major changes to it are presented in this paper. Sample comparisons between theoretical predictions and experimental data are also presented for several airfoil geometries. Areas of agreement and disagreement are discussed with recommendations for areas of needed program improvement. SYMBOLS
a

l a 2> a 3

coefficients of quadratic equation of the form f(x) = ajx2 + aX + a.%

714

AH c C(j Cf GI cm cn .

aerodynamic, influence coefficient chord of airfoil, cm (in.) section profile-drag coefficient skin-friction coefficient section lift coefficient section quarter-chord, pitching-moment coefficient section .normal-force coefficient
r i

Cp . H K M N p q^ RQ s Ue ,

local static-pressure coefficient, boundary-layer form factor, d*/0 local curvature, cm~* (in-1) Mach number

Pstat

f? " P

number of corner points for polygon approximation of airfoil -static pressure, N/m2 (lb/ft2) .(lb/ft ) ^ .

free-stream dynamic pressure, N/m

Reynolds number based on momentum thickness, surface distance along airfoil contour, cm (in.) velocity at edge of boundary layer, m/sec (ft/sec)

715

Uoo v x z a y 6
5

free-stream velocity, m/sec (ft/sec) local velocity on airfoil surface, m/sec (ft/sec) airfoil abscissa, cm (in.) airfoil ordinate, cm (in.) angle of attack of airfoil, deg strength of vortex singularity, m2/sec (ft^/sec)

boundary-layer thickness, cm (in.) angular deflection of flap, slat, or vane, deg boundary-layer displacement thickness, cm (in.). boundary-layer momentum thickness^ cm (in.) kinematic viscosity, mVsec (ft^/sec) cosine distribution angle, 0i = |?, deg stream function, m^/sec (ft^/sec)

f s or v

6* , , . . - 9 v 0 i// Subscripts: c 1 j 1 te "' ' . .

control point matrix row matrix column lower


-

. - . . .

'

-1 '

trailing edge

716

u
00

upper free stream

Abbreviations: B. L. NC TE boundary layer number of components trailing edge PROGRAM OPERATION AND CAPABILITIES The original airfoil program developed by the Lockheed-Georgia Company was designed to handle single-element and multielement airfoils. As many as four elements can be handled which typically consist of an airfoil with a leading-edge slat device and a double-slotted trailing-edge flap system. The airfoil program is composed of three main parts: (1) geometry specification, (2) potential flow, and (3) boundary layer. The nonoverlay version of the program requires a computer storage of approximately 200 000 octal locations (CDC 6600) and an execution time of approximately 200 CPU seconds for a typical four-element airfoil. An overlay version has reduced the required storage to approximately 65 000 octal locations with only a slight increase in execution time. The single-element version developed by North Carolina State University requires a computer storage of approximately 110 000 nonoverlay octal locations and 53 000 overlay octal locations with execution times of the order of 30 CPU seconds for a single case. A flow chart of the single-element version of the airfoil program is presented in figure 1. After data input and geometry specification (subroutine READIT and GEOM), the program enters an iterative cycle which involves the determination of interrelationship between the potential flow and the boundary layer (subroutines MAIN2 and MAINS). After each iteration a convergence check is made which consists of a simple comparison of the computed normal-force coefficients. Experience has shown that only five iterations are necessary to obtain a converged solution. This rapid convergence is possible because of the unique method used to combine the physical airfoil geometry and the computed displacement thicknesses from the boundary-layer computations in order to obtain the next iteration geometry. This convergence method will be discussed further in a following portion of this paper. The flow chart for the multielement version is very similar to that for the singleelement version, the exceptions being.an expanded geometry routine to handle the posi717

tioning of the elements relative to one another- and an additional boundary-layer routine -to handle the merging of boundary layers between fore and aft elements (confluent boundary layers). A program routine was available in the original multielement version to : correct the pressures in the slot regions between elements to account for the high frictional and surface curvature effects. This routine employed a marching-type.calculation procedure which required an accurate definition of the slot inlet conditions. An empirical formula was derived to estimate the inlet conditions from the potential-flow solution. Experience has shown that this slot-analysis routine produces erroneous results and it has, therefore, been omitted from the more recent program version. Since the inception of the airfoil program, the program input has been kept as simple as possible to make it more user oriented. The coordinates of each element of a multielement airfoil can be input with respect to a separate coordinate system and easily positioned relative to other elements by specifying the pivot point location and deflection of the element. The boundary-layer transition location (transition from laminar to turbulent) on each surface of an element can be input as either fixed or free. The total num ber of calculation points at which the pressures are desired can also be input and will be automatically allocated using the formula
=2
(1)

where Nj is the number of points allocated to the ith element, NSp is the total num.- . i ber of calculation points, Nc is number of components, ci is the chord of a given ele-; i" ment, and c<p is the summation of the values of c^. During a single machine pass, the angle of attack and Mach number can be varied for a constant Reynolds number, Prandtl number, and stagnation temperature. To represent the effects of compressibility, the well-known Karman-Tsien pressure correction law is employed which, therefore, limits the input Mach number to that producing critical flow. The laminar and turbulent boundary-layer routines contain methods to.predict boundary-layer separation, but do not contain methods to model the flow after separation. Therefore, the angle of attack should be limited to that producing only minor separation (less than one percent of the surface). PROGRAM THEORY AND MODIFICATIONS Airfoil Geometry Specification The user inputs to the airfoil program are the upper- and lower-surface coordinates of the airfoil shape. The airfoil is modeled within the program as a polygon approximation which is illustrated in figure 2. This polygon consists of N number of corner points with N - 1 number of straight line segments. Previous experience has shown that an N equal to 65 is sufficient to obtain an accurate viscous flow solution for a

-718

single-element airfoil. Of the 65 total points, 32 points are distributed, on the upper surface, 32 on the lower surface, and 1 at the leading-edge nose location. In most instances, the user desires to input more than 65 coordinates and does not want to be concerned about whether or not the input spacing will affect the computed results. For this reason, an automatic collocation method has been included in the airfoil program. The original version of the airfoil program distributed corner points with the cosine formula
6

(2)

where 0.j.= ^-, i = 1, 2, . . ., N. This procedure is illustrated in figure 3 and a sample distribution shown in figure 4(a). This method, in effect, closely spaces points near the leading and trailing edges of the airfoil. A close spacing near the leading edge is very desirable because of the high curvature and high velocities occurring in that region. Experience has shown, however, that closely spacing points near the relatively thin trailing-edge region can result in extreme oscillations in the computed velocities. The reason for these oscillations is discussed further in the potential-flow portion of this paper. To overcome this problem, a new collocation method has been formulated and incorporated into the airfoil program. The basis of the new method is that points are spaced relative to the local curvature. This method will closely space points only in regions of high curvature. To utilize this method the curvature at each user input coordinate is computed with the formula
,2
3/2

(3)

where - a^, a.%, and a3

are taken from a curve fit of the airfoil points of the form (0 s x < o.8c) (4)

z2 = ajx 2 + a2X + a3
and

z = ajx 2 + a2x + a3

'

(0.8c S x s c)

(5)

A curvature summation is then computed from the following equation and stored for backward interpolation:
Is (6)

719

where
s

= s

// T5 / \ ~ 2 i-l + \l(xi ~xi-l) + (zi ~ z i-l)

The maximum value of K is divided into N equal portions and the s value .correspond ing to each portion is then determined by backward interpolation between the Sj and Ki arrays. Additional backward interpolations are then made between the. si, 24, and Zi arrays to determine the new distributed airfoil coordinates. A sample of the results of this collocation method is presented in figure 4(b). Note the improvement in the distribution of points as compared with the cosine method. Caution by the user should be exercised regarding the number and relative positions of the input coordinates. If more input points are given, the computation of the curvature summation array will be more accurate and, therefore, the distribution of points more accurate. The'curve fit formula of equation (4), which is used in the nose. , region, is designed to approximate an infinite slope and a better estimation of the high curvature in that region can be obtained if more points are input. , .
' Potential-Flow Solution I

The potential-flow methods used to determine the velocity at specified locations on the surface of an airfoil generally fall into two categories. One category consists of conformal transformation methods and the second, singularity distribution methods. Con- . formal transformation methods have not been greatly utilized because of the difficulties encountered when trying to obtain transformation equations for airfoils of arbitrary shape and because of the inability of this method to handle blunt-base airfoils. Singularity distribution methods have been widely utilized since the advent of the high-speed, highcapacity digital computers which are needed to solve the large systems of simultaneous equations characteristic of these methods. These methods can handle arbitrarily shaped airfoils at any orientation relative to the free stream. For singularity distribution methods, either source, sink, or vortex singularities are distributed on the surface of the airfoil and integral equations formulated to deter - . mine the velocity induced at a point by the singularity. By dividing the airfoil surface into N segments and specifying either zero-normal or a tangential flow boundary condition for each segment, the integral equations can be approximated by a corresponding system of N - 1 simultaneous equations. By satisfying the Kutta condition at the trailing edge of the airfoil, the Nth equation can be formulated and then the singularity strengths determined with any one of a number of matrix-inversion techniques. The * ' *" Kutta condition usually employed is that the velocities at the upper- and lower-surface trailing edge be tangent to the surface and equal in magnitude. -',_.___
720

The singularity distribution method used in the original version of the airfoil program consisted of a distribution of vortices at the corners of the polygon approximation of the airfoil, with an additional constraint that the vortex strength vary linearly along the surface segments. A control point was selected at the midpoint of each segment and a boundary condition of no flow normal to the surface was applied. This resulted in N unknown vortex strengths and N - 1 boundary conditions and equations. An additional equation was obtained by satisfying the Kutta condition at the trailing edge as . ,.

because, for a vortex singularity, the magnitude of the tangential surface velocity is equal to the vortex strength at the corner point. Several problems were encountered with the original singularity distribution method. For airfoils with a trailing -edge cusp, the upper- and lower -surf ace vortices' near the trailing edge tend to become identical and, thereby, generate an almost singular matrix. For other types of airfoils, a too close spacing of vortices near the trailing, edge results in extreme oscillations of the vortex strengths. This method could not handle airfoils with an open or blunt -base characteristic of the recently developed Langley supercritical or "shockless" airfoils. To overcome these handicaps an improved singularity distribution method has been incorporated into the airfoil program. The new singularity distribution method was first formulated by H. J. Oellers to, compute the pressure distribution on the surface of airfoils in cascade, and is described in reference 3. Instead of working with induced velocities, characteristic of the previous method, Oellers' method employs stream functions. The stream function for a uniform free stream plus that of the vortex sheet is set to be a constant on the airfoil surface. This is represented mathematically by the Fredholm integral equation .

_LC

2?r Jf0

In r(s,) d = Ilexes) cos (a) - O^zfe) sin (a) .

(8)

where . if is the unknown stream function constant, r(s,) is the distance between two points on the airfoil surface, x(s) and z(s) are coordinates of a point on the surface, and y() is vortex strength at a point. By dividing the surface into N segments and assuming C9nstant vortex strength for each segment the above equation becomes
'
;

-- N

i// - } Ayyj = Uoofxi cos (a) - zi sin (a)J where the influence coefficient Ay is ! " ' '
S

(9)

'

(10)
721

By specifying a control point at the midpoint of each segment (denoted by a "c" subscript), the influence coefficient becomes
At, - it, and (i = j) in (r2) - tl in (r -

(12)

where
As = sj+1

= (xj+l - xc,i) 2

- z ,i) 2

n _ ^^__^___^___^__^_____^__^___^_________^_

(13)

-As

t1, 3

_ (XJ - Xcl

.As

To determine the vortex strength (y) at the intersection of two segments, the following "interpolation formula is used:
(j ..* 1 or N)

(14)

For this method, an additional equation is needed to obtain an N by N system of equations. (The unknowns are N-l number of y's and //.) A new method of applying the Kutta condition has been formulated to reduce the oscillations of the vortex strengths caused by a too close spacing of points near the trailing edge. This new Kutta condition simply requires that the vortex strengths (y) vary quadratically for the last four segment corners near the upper and lower surface of the trailing edge and that at the trailing edge

722

This quadratic variation is expressed as y(s) = ajs^ + a2s + a3

. (16)

By combining equations (14) and (16), the equation for the vortex strength at the lowersurface trailing edge becomes ">

(17)
where _ S4s2(s3 - s4) Ji = D _ 84 (S3 " S2)(S3 + S2 " S4) Jo = D
S S

2( 4 - 3) s4
S

J, =

3(S2 -S3)" 5 - 83

" S3s2(s2 - s3) (s5 - s4) _.= ._ J D(s5 - s3)

and
D = s22(s3 - 84) + s32(s4 - s2) + s42(s2 - s3) A similar expression can be obtained for the upper-surface trailing-edge vortex by replacing the subscripts in equation (17) as follows: si = SN - s N+1 _ t
J

(i = 1, 2, . . ., 5)

=J

N-i
= l,2,3,4)

Combining equations (16) and (17) with equation (15) yields the needed Nth equation in the form .
=0

(i = 1, 2, 3, 4, N-4, N-3, N-2, N-l)

(18)

The Oellers1 method with the modified Kutta condition has been incorporated into the current version of the airfoil program. This new potential-flow method combined

723

with the new geometry specification method has successfully overcome the problems encountered with t h e original methods. - . . . - ' Boundary-Layer Solution . .

-.

The pressure coefficients computed in the potential-flow portion of the airfoil program are corrected to account for compressibility with the well-known Karman-Tsien correction law. Using the isentropic flow relations, the local Mach number is computedand input to the boundary-layer portion of the program. The boundary-layer development on a typical multielement airfoil is illustrated in figure 5. The boundary layer consists of an ordinary boundary layer (nonmerging boundary layer) and a confluent boundary layer (merging boundary layer). The ordinary boundary layer is composed of laminar, transition, and turbulent regions. The confluent boundary layer is composed of core, main I, main II, and turbulent regions as shown in figure 6. The confluent boundary- " layer model was developed by Sure sh H. Goradia from the Lockheed-Georgia Company and is one of the unique features of this program. The meaningful parameters output from the boundary-layer portion of the program are (1) the,displacement thickness 6*, (2) the momentun thickness 0, (3) the form factor H, and (4) the skin-friction coefficient Cf. The theoretical development of the boundary-layer methods used in this program are quite lengthy and, therefore, only a brief description will be presented in this paper. A flat-plate boundary-layer analysis is performed on each surface of an airfoil element, and the leading-edge stagnation point is the plate leading edge. A flow chart of the boundary-layer computations is presented in figure 7. An initial laminar boundarylayer region exists from the stagnation point to the point of transition from laminar to turbulent. The laminar boundary-layer model used is the method of Cohen and Reshotko as presented in reference 4. After computing the laminar boundary-layer characteristics at a discrete point, routine BLTRAN is called to check for transition and, if transition has occurred, to check for the formation of a long or short transition bubble and for laminar stall. The sequence of calculations within BLTRAN is presented in figure 8; An initial check is made to determine if the laminar boundary layer is stable or unstable based on the instability criterion established by Schlicting and Ulrich as presented in reference 5. If the boundary layer is unstable, a transition check is then made based on an empirically derived transition prediction curve. 'If transition1 has occurred, the initial quantities needed to start the turbulent calculations are computed. If transition has not occurred, the formation of either a long bubble with corresponding laminar stall or a short bubble with corresponding'reattachment is determined. The user .can input a fixed transition location and a check will be -made at the beginning of BLTRAN to determine whether or not the fixed location has been reached. . . After computing the transition location and corresponding initial boundary-layer properties, the turbulent boundary-layer calculations are made. The original version of
724

the program used a modified Truckenbrodt method derived by Gdradia to compute the turbulent boundary-layer properties. The Goradia method was less sensitive to local pressure oscillations and was capable of traversing localized separated/reattaching flow regions. These local pressure oscillations were caused primarily by the numerical inacccuracies associated with the potential-flow method. The Goradia method underestimated the boundary-layer displacement thicknesses which, therefore, resulted in an underestimation of the lift,and pitching-moment coefficients. After incorporating Oellers' potential-flow method, the pressure oscillations were reduced which meant that the original Truckenbrodt method described in reference 6 could be used to improve the turbulent boundary-layer computations. The original Truckenbrodt. method was coded and incorporated into ,the recent program version and has resulted in an imprpvement in the estimation of the lift and pitching-moment coefficients. If a slot exiting plane is reached during the turbulent boundary-layer computations, the confluent boundary-layer computation is initiated. The confluent boundary layer is a result of the mixing between the slot efflux and the wake from the forward element, and can exist from the slot exit to the trailing edge of the element depending upon the pressure distribution. The confluent boundary-layer model was formulated by Goradia and consists of various regions and layers as illustrated in figure 6. The model is based on the assumption that-the merging of fore-and-aft element boundary layers will have "similar" velocity profiles if nondimensipnalized in a way analogous to that for a free-jet flow. By utilizing this assumption, the governing partial differential equations were, reduced to a set of.ordinary differential equations which could be easily solved with available numerical techniques. .Several empirical constants were needed to establish the similar velocity profiles and were obtained from experimental tests performed by Goradia as reported in reference 7. The only improvement made to the original confluent boundaryTlayer routines has been, minor adjustments in the values of these empirical constants. The model formulated by Goradia assumed that the core velocity exiting the slot is greater than the velocity at the upper edge of the wake layer at the slot exit. Experience has shown that this velocity relationship is not always true and, therefore, erroneous performance predictions can occur. It is generally believed that this velocity relationship should be true for an efficient design and, therefore, no additional work has been done in this area. None of the boundary-layer methods used in the airfoil program include curvature effects; All the methods used are basically integral methods which are generally less accurate than finite-difference methods but require considerably less computer time. More accurate infinite-difference methods are available, as described in references 8, 9, and.10, and can be easily incorporated into the existing program.

725

Equivalent Airfoil Geometry The airfoil program uses an iterative procedure to obtain the viscous solution and the basic steps are as follows: (1) Compute a potential-flow solution for the basic airfoil. (2) Compute boundary-layer properties based on the potential-flow solution. ... (3) Construct a modified airfoil by adding the boundary-layer displacement thickness to the original airfoil. (4) Compute the aerodynamic performance coefficients. (5) Repeat steps (1) through (4) until convergence of the performance coefficients is obtained. The most important step in the iterative procedure is step (3) which involves selecting a geometry modification method that will insure convergence for almost any input airfoil shape in a reasonable number of iterations. The method developed by Lockheed has been highly successful and has provided converged solutions after only four or five iterations. The method developed by Lockheed to modify the airfoil shape is based on the assumption that the effect of the boundary layer on the basic thickness and basic camber can be considered separately and then superimposed to determine the net effect. The addition of the boundary layer has an uncambering effect near the trailing edge which causes a reduction in the effective angle of attack and lift coefficient, and it has a thickening effect along the airfoil which causes an increase in the local surface velocities and lift coefficient. .Experience has shown that the thickness effects are of secondary importance for multielement airfoils and are, therefore, omitted in the multielement program version. Thickness effects are, however, included in the single-element program version to improve the overall accuracy of the performance predictions. The camber change is given as the difference in the magnitude of the upper- and lower-surf ace displacement thicknesses as illustrated in figure 9(a). The thickness change is given as the difference in two thickness solutions as shown in figure 9(b).. The first thickness solution is for a . symmetric airfoil at a 0 angle of attack with the same thickness distribution as the original input airfoil/; The second thickness, solution.,is also, for a> symmetric airfoil at a 0 angle of attack with the thickness distribution of the original input airfoil plus the sum of ,the upper- and lower-surface displacement thicknesses. .. ' Applying superposition and a proportioning technique to prevent over-correction during the initial iterations, the velocity distribution becomes (vtotal)i
=

726

where BT stands for the basic thickness distribution and

kt = 1
ki=f

(i * 2)
. (1 = 2)

The addition of the displacement thickness to obtain the thickness solution often produces a symmetric airfoil with a very thick trailing edge. In actual flow the wake acts as an afterbody with a rapidly decreasing thickness distribution. Based on the work of Powell as reported in reference 11 and from observations of available experimental data, Summey and Smith from the North Carolina State University formulated the following analytical expression to represent the afterbody shape: = I[(zte - Z o o )e-6.9xx + Z o o J(i _ x x ) . . (20)

where xx = c = x 2c and z^, =-iu GJ c. The Squire and Young drag formc - 1 for . .2 o ula from reference 5 is used to compute the drag coefficient at infinity and is given as Hu+5 H1+5:
2 2

.,SlkY

.'jOt*

After each boundary -layer solution, the displacement thickness distribution is smoothed three times by using a standard least-squares smoothing technique. To prevent over -correction during the initial iterations, a proportioning technique similar to that. used for the velocities is used to compute the effective displacement thickness distribution and is given as

where BL stands for the present boundary-layer solution. COMPARISONS BETWEEN EXPERIMENT AND THEORY Both the single -element and multielement versions of the airfoil program have been widely distributed and utilized by the industrial and academic communities. The program is widely used within Langley for the design and analysis of new subsonic and transonic airfoils for application to helicopters, general aviation aircraft, and transport aircraft. Shown in figure 10 is a comparison between the theoretical and experimental performance characteristics for the recently developed NASA GA(W)-1 airfoil that was designed by Richard T. Whitcomb especially for application to general aviation aircraft. .(See ref. 12.)

727

Excellent agreement is demonstrated until an angle of attack is reached at which the boundary layer begins to separate near the upper-surf ace trailing edge. Currently, a flow separation model does not exist in the program; however, a research study is being proposed by Langley in this area of boundary-layer research. Excellent agreement is also demonstrated for the drag prediction of the GA(W)-1 airfoil, but this is not generally true, for all airfoils. The method used in the program to compute drag consists of a simple integration of the pressure and shear (skin friction) forces. An improved drag computation method is currently under development by Lockheed-Georgia, under NASA Contract NAS 1-12170, which computes the drag from computed downstream wake characteristics. This improved drag method will be incorporated into the single-' element airfoil program after verification at Langley. '
' *

A Fowler^type, _single-slotted flap was also designed for the GA(W)-1 airfoil and . later tested at Wichita State University under a NASA grant. , The results of the flapped airfoil tests are presented in reference 13, and a summary of the agreement between, experiment and theory is presented in figure 11. Excellent agreement is demonstrated until a 40 flap deflection is reached where separation occurs on the upper surface of the flap. Similar agreement was obtained from tests of a blunt-base airfoil with a single slotted flap reported in reference 14. Typical agreement between the experimental and theoretical confluent boundary-layer velocity profiles for this airfoil is shown in figure 12. The multielement airfoil program can also be used to perform a gap optimization study of a given flap system. Presented in figure 13 are the results of a gap optimization study for a 10 drooped-nose airfoil with a single-slotted flap as reported in reference 15. The viscous and inviscid theory predictions are also presented in this figure, and although the agreement between viscous theory and experiment is fair, the viscous theory does predict the correct optimum gap of 2 percent. The agreement between theory and experiment for multielement airfoils with three or more elements has not been generally as good as that for single- or two-element airfoils. This can usually be attributed to the fact that an airfoil with three'or. more elements generally will have some separated flow on at least one element. Shown in figures 14(a) and: 14(b) is the agreement between viscous and inviscid theory and experiment for a typical three-element airfoil with a leading-edge slat and a single-slotted, trailing-edge flap. (See ref. 16.) Excellent agreement is demonstrated over an angleof -attack range from -4 to 12 for the force and moment coefficients. A typical pressure-distribution agreement is shown' in figure 14(b) for an angle of attack of 8. Shown in figure 15 is the agreement between theory and experiment for a typical fourelement airfoil with a leading-edge slat and a double-slotted, trailing-edge flap. (This is model C in ref. 14.) The low -5/15 vane/flap deflection case showed poor agreement because at those deflections there were large overlap areas between the vane and flap and the airfoil program does not correct the pressure distribution for slot effects. (Previous
728

discussion pointed out that the original slot flow correction method proposed by Lockheed was erroneous and, therefore, has been dropped from the program.) The high 30/65 deflection case showed very poor agreement because of the large amount of flow separation on the upper surface of the flap element. The intermediate 14/33 case, however, showed good agreement because there were no large overlap or separated flow areas present. Shown also in figure 15 is the extremely poor drag agreement which is typical for multielement airfoil cases. The sample cases presented in this paper were selected to demonstrate some of the good as well as bad features of the airfoil program. The first major area of needed program improvement is the computation of profile drag. Many researchers use the well-known Squire and Young drag formula (eq. (21)) to compute the profile drag; however, this formula was derived based on a proven erroneous assumption about the variation of the downstream boundary-layer form factor, H. The second major area of improvement is the computation of the characteristics of the flow region after boundary-' layer separation. The lack of experimental data for separated flow regions has greatly hampered the development of a theoretical model. Experimental tests have recently been completed'at Wichita State University to map the velocities and pressures in the separated regions of the GA(W)-1 airfoil. The third major area of improvement is the computation of the pressure corrections in the slot areas between overlapping elements of multielement airfoils. This area has been of lesser importance because the designer generally desires to have as much Fowler motion (increase in effective chord) as possi- ble which results in relatively small overlap regions. CONCLUDING REMARKS This paper has discussed in detail the theoretical and operational features and capabilities of the Langley "2-D Subsonic Multi-Element Airfoil Program." Several major modifications have been made to improve the applicability and prediction accuracy of this program and involve the use of: (1) Curvature instead of cosine method to distribute segment corner points for' polygon representation of the airfoil ' (2) Oellers' stream function method to obtain the potential-flow solution (3) Modified Kutta condition to reduce pressure oscillations at the trailing edge (4) Truckenbrodt method to obtain the turbulent boundary-layer characteristics Comparisons between experimental data and theoretical predictions indicate excellent agreement for single-element airfoils in attached flow and good agreement for multi'

729

element airfoils in attached flow with small overlap regions between elements. Three major areas of program improvement are needed and involve the computation of: (1) Profile drag from downstream wake characteristics (2) Flow characteristics after boundary-layer separation .(3) Pressure corrections in the slot areas between overlapping elements of multielement airfoils The airfoil program has been demonstrated as an effective tool for the design and analysis of single-element and multielement airfoils in viscous flow, and has been widely distributed to and utilized by both the academic and industrial communities.

730

' '
\ , " , "

REFERENCES

'

'
-

1. Stevens, W. A.; Goradia, S. H.; and Braden, J. A.: Mathematical Model for TwoDimensional Multi-Component Airfoils in Viscous Flow. NASA CR-1843, 1971. 2. Smetana, Frederick O.; Summey, Delbert C.; Smith, Neill S.; and Garden, Ronald K.: . - * ' . . , Light Aircraft Lift, Drag, and Moment Prediction - A Review and Analysis. NASA CR-2523, 1975. 3. Oellers, Heinz J.: Incompressible Potential Flow in a Plane Cascade Stage. NASA TT F-13,982, 1971. 4. Cohen, Clarence B.; and Reshotko Eli: The Compressible Laminar Boundary Layer With Heat Transfer and Arbitrary Pressure Gradient. NACA Rep. 1294, 1956. (Supersedes NACA TN 3326.) 5. Schlichting, Hermann (J. Kestin, trans!.): Boundary-Layer Theory. Sixth ed., McGraw-Hill Book Co., Inc., 1968. 6. Truckenbrodt, E.: A Method of Quadrature for Calculation of the Laminar and Turbulent Boundary Layer in Case of Plane and Rotationally Symmetrical Flow. NACA TM 1379, 1955. 7. Goradia, S. H.: Confluent Boundary Layer Flow Development With Arbitrary Pressure Distribution. Ph. D. Thesis, Georgia Inst. Technol., 1971. 8. Dvorak, F. A.; and Woodward, F. A.: A Viscous/Potential Flow Interaction Analysis Method for Multi-Element Infinite Swept Wings. Volume I. NASA CR-2476, 1974. 9. Callaghan, J. G.; and Beatty, T. D.: A Theoretical Method for the Analysis and Design of Multi-Element Airfoils. AIAA Paper No. 72-3, Jan. 1972. 10. Bhateley, I. C.; and Bradley, R. G.: A Simplified Mathematical Model for the Analysis of Multi-Element Airfoils Near Stall. Fluid Dynamics of Aircraft Stalling. AGARD Conf. Pre-Print No. 102, Apr. 1972, pp. 12-1 - 12-12. 11. Powell, B. J.: The Calculation of the Pressure Distribution on a Thick Cambered Aerofoil at Subsonic Speeds Including the Effects of the Boundary Layer. C.P. No.1005, British A.R.C., June 1967. 12. McGhee, Robert J.; and Beasley, William D.: Low-Speed Aerodynamic Characteristics of a 17-Percent-Thick Airfoil Section Designed for General Aviation Applications. NASA f N D-7428, 1973. 13. Wentz, W. H., Jr.; and Seetharam, H. C.: Development of a Fowler Flap System for a High Performance General Aviation Airfoil. NASA CR-2443, 1974.

731

14. Omar, E.; Zierten, T.; and Mahal, A.: Two-Dimensional Wind-Tunnel Tests of a NASA Supercritical Airfoil With Various High-Lift Systems. Volume I - Data Analysis. NASA'CR-2214, 1973. 15. Foster, D. N.; Irwin, H. P. A. H.; and Williams, B. R.: The Two-Dimensional Flow Around a Slotted Flap; R. & M. No. 3681, British A.R.C., 1971. 16. Harris, Thomas A.; and Lowry, John G.: Pressure Distribution Over an NACA 23012. Airfoil With a Fixed Slot and a Slotted Flap. NACA Rep. 732, 1942.

732

00 O

H CO

0)

<a

0)

ai .-I

CO

<u
V4

&0 rl

733

12

13

x/c

5
Figure 2.- Polygon approximation of airfoil.

Figure 3.- Cosine method for distribution of airfoil corner points.


734

CO

O P.

g
O O
0) CO

0)

I
M 4-) CO

13 O
<U

O JS
4-1 Q)

M-l

tJ O

H CO O

B cu c

0)
4J

5O
0)

rt >

I co
O O
O CO
0)

0) M

735

T- C\J

oo

736

737

co
H 4J
4-1

cd

O.

I o
M <U

-O

rt

w (0

&
'l
60

L__Jr: - ^ _ _ _

738

739

* >.
CO CO CD

.c u
o
N
(X CO

c cu o

o
H 4J
O

c o
efl
T3 C

cu >

CO o

T)

C
M
0)
^ O

0)

1s
t5 II
>> o

co co cu

JB CJ

cu a

CJ

^\
cfl
#

to a>

c
o H u
CO 0
Tl

I
CTi

ft
M

f t gs

o "o

0) VJ

00 H

Nl

740.

Experimental data Viscous theory

cm-.l
-.2
. 0 3 . 0 2

0L

2 . 0

1.6
D D

1.2

a
Start of trailing-edge ''' >' separation

. 8
V

r-.4-

-4

8 a.deg

12

16

20

Figure 10.- Comparison between experiment and theory for NASA GA(W)-1 airfoil.

741'

8 a,deg

16

20

Figure 11.- Comparison between experiment and theory for GA(W)-1 airfoil with a 29% chord Fowler flap.
742

ro

Iz
UJ

UJ o Q_ UJ
X

? UJ h-

OJ

CM

a)

I
8

743

' Gop

,:

-:,-:.:..,, 7^

8f=30o

Inviscid Theory . Viscous Theory Experimental Dato


2.8r

I..----.

....
AP/

> ,,.- 3,->.

.^^-.. '.-..-.-. -...-- - ..<, G. .,,..percent ^ .. _,. . ;1 _ _^ /, Figure 13.- Gap optimization 'of-10 drooped^nose:.;airfoil with a single-slotted, trailing-edge flap.

744':

tnviscid Theory 'Viscous Theory Experimenfol Goto

3i2r

2,8-

o.deg (a) Force and nooent comparison. '. Figure 14.- Comparison between experiment and theory for NACA 23012 airfoil with leading-edge slat and trailing-edge flap. ' \

-8-4

12

|6

20

745

Experimental.'

, Viscous Theory
pli

-5 -4
C

_
6
u! o'.

P
_ O Upper. ' Q Lower _"+_ + - . - - . . .

-3 P -2 -1

'

a = 8
'<>.
CH

:
t L

'I
1 1

'
*

..*
-/c

"

'''-.&
, ,' ' '"* 0-:.-,.^.^
p

".
^'

. , '
\ .;
o*.
*0

'

'

'

0 ~f\ '

J
J

r
la-P"

'
1 1 1

- ' . ...-6--0
1 1 1-

/
1 ' 'I

^ <y

***.. o -

1.

^tfl ^ 1

.2 .1 2

' ' .
'.
' ~^~ ^^rV

/ /C

0 U ^~~^ =^S^.\-/
=

/^*~^
*""

^~i^-~ ^aL

'-

^
\

-.1 i_2

.';':

' c)

II

II

, .:';| ; ;..--,
) C
-"! ' '' O ;!:; .2
':>

.,
Tt .3

.1

.2

C)

.1

.2

.3

.5

.6

.7

.8

/C (b) Pressure distribution comparison.

xy

Figure 14.- Concluded.

^46

in * i
it

H M

E 3 o w

g r-i 0)

4, O 60 H-l "O

O 4-1 0) O A tH
4J CO

O 0) 0 H

a) ,0

4-1 O

a "9 0)

I
CO O) 0) 0) 00
,*J *S 0) I

0)

C -H O T3

a id

O a

I
4)

747

Vous aimerez peut-être aussi