joa vol41 no1 lifting-line analysis for twistedwings and washout-optimizedwings

9
JOURNAL OF AIRCRAFT Vol. 41, No. 1, January–February 2004 Lifting-Line Analysis for Twisted Wings and Washout-Optimized Wings W. F. Phillips Utah State University, Logan, Utah 84322-4130 A more practical form of the analytical solution for the effects of geometric and aerodynamic twist (washout) on the low-Mach-number performance of a finite wing of arbitrary planform is presented. This infinite series solution is based on Prandtl’s classical lifting-line theory and the Fourier coefficients are presented in a form that only depends on wing geometry. The solution shows that geometric and aerodynamic washout do not affect the lift slope for a wing of any planform shape. This solution also shows that any wing with washout always produces induced drag at zero lift. Except for the special case of an elliptic planform, washout can be used to reduce the induced drag for a wing producing finite lift. A relation describing the optimum spanwise distribution of washout for a wing of arbitrary planform is presented. If this optimum washout distribution is used, a wing of any planform shape can be designed for a given lift coefficient to produce induced drag at the same minimum level as an elliptic wing with the same aspect ratio and no washout. Nomenclature A n = coefficients in the infinite series solution to the lifting-line equation a n = planform contribution to the coefficients in the infinite series solution to the lifting-line equation b = wingspan b n = washout contribution to the coefficients in the infinite series solution to the lifting-line equation C Di = induced-drag coefficient C L = lift coefficient C Ld = design lift coefficient C L = wing lift slope ˜ C L = airfoil section lift slope c = wing section chord length R A = wing aspect ratio R T = wing taper ratio V = magnitude of the freestream velocity z = spanwise coordinate α = geometric angle of attack relative to the freestream α L0 = airfoil section zero-lift angle of attack = spanwise section circulation distribution γ t = strength of shed vortex sheet per unit span ε = washout effectiveness θ = change of variables for the spanwise coordinate κ D = planform contribution to the induced-drag factor κ DL = lift-washout contribution to the induced-drag factor κ Do = optimum induced-drag factor κ D = washout contribution to the induced-drag factor κ L = lift slope factor = maximum total washout, geometric plus aerodynamic opt = optimum total washout to minimize induced drag ω = normalized washout distribution function Introduction P RANDTL’S classical lifting-line theory 1,2 can be used to obtain an infinite series solution for the spanwise distribution of vor- Received 26 November 2002; presented as Paper 2003-0393 at the 41st Aerospace Sciences Meeting, Reno, NV, 6–9 January 2003; revision received 19 February 2003; accepted for publication 26 May 2003. Copyright c 2003 by W. F. Phillips. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission. Copies of this paper may be made for personal or internal use, on condition that the copier pay the $10.00 per-copy fee to the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923; include the code 0021-8669/04 $10.00 in correspondence with the CCC. Professor, Mechanical and Aerospace Engineering Department, 4130 Old Main Hill. Member AIAA. ticity generated on a finite wing. If the circulation about any section of the wing is (z ) and the strength of the shed vortex sheet per unit span is γ t (z ), as shown in Fig. 1, then Helmholtz’s vortex theo- rem requires that the shed vorticity is related to the bound vorticity according to γ t (z ) =− d dz (1) For a finite wing with no sweep or dihedral, combining this relation with the circulation theory of lift produces the fundamental equation that forms the foundation of Prandtl’s lifting-line theory: 2(z ) V c(z ) + ˜ C L 4π V b/2 ζ =−b/2 1 z ζ d dz z = ζ dζ = ˜ C L [α(z ) α L0 (z )] (2) In Eq. (2) α and α L0 are allowed to vary with the spanwise coor- dinate to account for geometric and aerodynamic twist, such as the examples shown in Fig. 2. For a given wing design, at a given angle of attack and airspeed, the planform shape, airfoil section lift slope, geometric angle of attack, and zero-lift angle of attack are all known functions of spanwise position. The only unknown in Eq. (2) is the section circulation distribution (z ). An analytical solution to Prandtl’s lifting-line equation can be obtained in terms of a Fourier sine series. From this solution the circulation distribution is given by (θ) = 2bV n = 1 A n sin(nθ) (3) where θ = cos 1 (2z /b) (4) and the Fourier coefficients A n must satisfy the relation n = 1 A n 1 + n ˜ C L c(θ) 4b sin(θ) sin(nθ) = ˜ C L c(θ)[α(θ) α L0 (θ)] 4b (5) From this circulation distribution, the resulting lift and induced-drag coefficients for the finite wing are found to be C L = π R A A 1 (6) C Di = π R A n = 1 nA 2 n = C 2 L π R A + π R A n = 2 nA 2 n (7) 128

Upload: donzqw8194

Post on 08-Feb-2016

42 views

Category:

Documents


0 download

DESCRIPTION

Lifting-Line Analysis for TwistedWingsand Washout-OptimizedWings

TRANSCRIPT

JOURNAL OF AIRCRAFT

Vol. 41, No. 1, January–February 2004

Lifting-Line Analysis for Twisted Wingsand Washout-Optimized Wings

W. F. Phillips∗

Utah State University, Logan, Utah 84322-4130

A more practical form of the analytical solution for the effects of geometric and aerodynamic twist (washout) onthe low-Mach-number performance of a finite wing of arbitrary planform is presented. This infinite series solutionis based on Prandtl’s classical lifting-line theory and the Fourier coefficients are presented in a form that onlydepends on wing geometry. The solution shows that geometric and aerodynamic washout do not affect the lift slopefor a wing of any planform shape. This solution also shows that any wing with washout always produces induceddrag at zero lift. Except for the special case of an elliptic planform, washout can be used to reduce the induced dragfor a wing producing finite lift. A relation describing the optimum spanwise distribution of washout for a wing ofarbitrary planform is presented. If this optimum washout distribution is used, a wing of any planform shape canbe designed for a given lift coefficient to produce induced drag at the same minimum level as an elliptic wing withthe same aspect ratio and no washout.

NomenclatureAn = coefficients in the infinite series solution to the

lifting-line equationan = planform contribution to the coefficients in the infinite

series solution to the lifting-line equationb = wingspanbn = washout contribution to the coefficients in the infinite

series solution to the lifting-line equationCDi = induced-drag coefficientCL = lift coefficientCLd = design lift coefficientCL ,α = wing lift slopeCL ,α = airfoil section lift slopec = wing section chord lengthRA = wing aspect ratioRT = wing taper ratioV∞ = magnitude of the freestream velocityz = spanwise coordinateα = geometric angle of attack relative to the freestreamαL0 = airfoil section zero-lift angle of attack� = spanwise section circulation distributionγt = strength of shed vortex sheet per unit spanε� = washout effectivenessθ = change of variables for the spanwise coordinateκD = planform contribution to the induced-drag factorκDL = lift-washout contribution to the induced-drag factorκDo = optimum induced-drag factorκD� = washout contribution to the induced-drag factorκL = lift slope factor� = maximum total washout, geometric plus aerodynamic�opt = optimum total washout to minimize induced dragω = normalized washout distribution function

Introduction

P RANDTL’S classical lifting-line theory1,2 can be used to obtainan infinite series solution for the spanwise distribution of vor-

Received 26 November 2002; presented as Paper 2003-0393 at the 41stAerospace Sciences Meeting, Reno, NV, 6–9 January 2003; revision received19 February 2003; accepted for publication 26 May 2003. Copyright c© 2003by W. F. Phillips. Published by the American Institute of Aeronautics andAstronautics, Inc., with permission. Copies of this paper may be made forpersonal or internal use, on condition that the copier pay the $10.00 per-copyfee to the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers,MA 01923; include the code 0021-8669/04 $10.00 in correspondence withthe CCC.

∗Professor, Mechanical and Aerospace Engineering Department, 4130Old Main Hill. Member AIAA.

ticity generated on a finite wing. If the circulation about any sectionof the wing is �(z) and the strength of the shed vortex sheet perunit span is γt (z), as shown in Fig. 1, then Helmholtz’s vortex theo-rem requires that the shed vorticity is related to the bound vorticityaccording to

γt (z) = −d�

dz(1)

For a finite wing with no sweep or dihedral, combining this relationwith the circulation theory of lift produces the fundamental equationthat forms the foundation of Prandtl’s lifting-line theory:

2�(z)

V∞c(z)+ CL ,α

4πV∞

∫ b/2

ζ = −b/2

1

z − ζ

(d�

dz

)z = ζ

= CL ,α[α(z) − αL0(z)] (2)

In Eq. (2) α and αL0 are allowed to vary with the spanwise coor-dinate to account for geometric and aerodynamic twist, such as theexamples shown in Fig. 2. For a given wing design, at a given angleof attack and airspeed, the planform shape, airfoil section lift slope,geometric angle of attack, and zero-lift angle of attack are all knownfunctions of spanwise position. The only unknown in Eq. (2) is thesection circulation distribution �(z).

An analytical solution to Prandtl’s lifting-line equation can beobtained in terms of a Fourier sine series. From this solution thecirculation distribution is given by

�(θ) = 2bV∞∞∑

n = 1

An sin(nθ) (3)

where

θ = cos−1(−2z/b) (4)

and the Fourier coefficients An must satisfy the relation

∞∑n = 1

An

[1 + n

CL ,αc(θ)

4b sin(θ)

]sin(nθ) = CL ,αc(θ)[α(θ) − αL0(θ)]

4b

(5)From this circulation distribution, the resulting lift and induced-dragcoefficients for the finite wing are found to be

CL = π RA A1 (6)

CDi = π RA

∞∑n = 1

n A2n = C2

L

π RA+ π RA

∞∑n = 2

n A2n (7)

128

PHILLIPS 129

Fig. 1 Prandtl’s model for the bound vorticity and trailing vortex sheetgenerated by a finite wing.

Fig. 2 Examples of geometric and aerodynamic twist.

Historically, the coefficients in this infinite series solution have usu-ally been evaluated from collocation methods. Typically, the series istruncated to a finite number of terms and the coefficients in the finiteseries are evaluated by requiring Eq. (5) to be satisfied at a numberof spanwise locations equal to the number of terms in the series. Avery straightforward method was first published by Glauert.3 Themost popular method, based on Gaussian quadrature, was originallydeveloped by Multhopp.4 Most recently, Rasmussen and Smith5 pre-sented a more rigorous and rapidly converging method, based on aFourier series expansion similar to that first presented by Lotz6 anddescribed by Karamcheti.7

The general lifting-line solution expressed in the form ofEqs. (3–7) is cumbersome for evaluating traditional wing proper-ties, because the Fourier coefficients depend on angle of attack andmust be reevaluated for each operating point studied. For a wingwith no geometric or aerodynamic twist, α and αL0 are independentof θ and the Fourier coefficients in Eq. (5) can be written as3

An ≡ an(α − αL0) (8)

which reduces Eq. (5) to

∞∑n = 1

an

[4b

CL ,αc(θ)+ n

sin(θ)

]sin(nθ) = 1 (9)

The Fourier coefficients obtained from Eq. (9) are independent of theangle of attack and only depend on the airfoil section lift slope andwing planform. Using Eq. (8) in Eqs. (6) and (7), the lift and induced-drag coefficients for a wing with no geometric or aerodynamic twistcan be written as follows:

CL = CL ,α(α − αL0)

(1 + CL ,α/π RA)(1 + κL)(10)

CDi = C2L(1 + κD)

π RA(11)

where

κL = 1 − (1 + π RA/CL ,α)a1

(1 + π RA/CL ,α)a1

(12)

κD =∞∑

n = 2

n

(an

a1

)2

(13)

For a detailed presentation of Prandtl’s lifting-line theory seeRefs. 8–12.

Solution for Wings with Geometricand Aerodynamic Twist

For a wing with geometric and/or aerodynamic twist, the aerody-namic angle of attack, α − αL0, is not constant along the span andthe relation given by Eq. (8) cannot be applied. Because geometricand aerodynamic twist enter the lifting-line formulation only as asum, we need not distinguish between these two types of twist. Toavoid repeated use of the lengthy and cumbersome phrase geomet-ric and aerodynamic twist, in the following presentation the wordstwist and washout will be used synonymously to indicate a span-wise variation in either the local geometric angle of attack or thelocal zero-lift angle of attack. Washout is treated as positive at thosesections with a lower aerodynamic angle of attack than the root,and negative at those sections with a higher aerodynamic angle ofattack than the root.

The procedure commonly used for lifting-line analysis of twistedwings is based on Eqs. (3–7) and requires evaluating separate setsof the Fourier coefficients An for more than one angle of attack.Because the unstalled lift coefficient is a linear function of angle ofattack, the wing lift slope and zero-lift angle of attack for a finitetwisted wing can be determined by evaluating only two sets of co-efficients An , one set for each of two different angles of attack.However, as pointed out by Karamcheti,7 the induced-drag coeffi-cient for a twisted wing is not a linear function of the lift coefficientsquared, as it is for an untwisted wing. Thus, to obtain the induced-drag polar for a twisted wing using this procedure, several sets ofthe Fourier coefficients An are determined over a range of angles ofattack. Bertin9 as well as Kuethe and Chow11 present the details ofthis procedure together with example computations.

Glauert3 authored the first textbook to present a lifting-line anal-ysis for twisted wings. In this work, he considered only the rect-angular planform with a linear spanwise symmetric variation inwashout. In the 1930s and 1940s this type of analysis became quitecommonplace (see, for example, Refs. 13–21). In addition to mod-eling a continuous variation in twist along the span of a finite wing,lifting-line theory has also been used to evaluate the spanwise liftdistribution resulting from a step change in aerodynamic angle ofattack, which is caused by the deflection of partial span trailing-edge flaps and ailerons.22−28 Because a fairly large number of termsmust be carried to adequately represent the Fourier expansion ofa step function, lifting-line analysis for wings with deflected flapsand ailerons was extremely laborious prior to the development ofthe digital computer. This labor was amplified by the fact that theprocedure required evaluating a complete set of Fourier coefficientsfor each angle for attack and flap deflection angle investigated.

130 PHILLIPS

A more practical form of the lifting-line solution for twisted wingscan be obtained by using the simple change of variables

α(θ) − αL0(θ) ≡ (α − αL0)root − �ω(θ) (14)

where � is defined to be the maximum total washout, geometricplus aerodynamic,

� ≡ (α − αL0)root − (α − αL0)max (15)

and ω(θ) is the washout distribution normalized with respect to themaximum total washout:

ω(θ) ≡ α(θ) − αL0(θ) − (α − αL0)root

(α − αL0)max − (α − αL0)root(16)

The normalized washout distribution function ω(θ) is independentof angle of attack and varies from 0.0 at the root to 1.0 at the pointof maximum washout, commonly at the wingtips.

Using Eq. (14) in Eq. (15) gives

∞∑n = 1

An

[1 + n

CL ,αc(θ)

4b sin(θ)

]sin(nθ)

= CL ,αc(θ)

4b[(α − αL0)root − �ω(θ)] (17)

The Fourier coefficients An in Eq. (17) can be conveniently writtenas

An ≡ an(α − αL0)root − bn� (18)

where the Fourier coefficients an and bn are obtained from

∞∑n = 1

an

[4b

CL ,αc(θ)+ n

sin(θ)

]sin(nθ) = 1 (19)

∞∑n = 1

bn

[4b

CL ,αc(θ)+ n

sin(θ)

]sin(nθ) = ω(θ) (20)

Comparing Eq. (19) with Eq. (9), we see that the Fourier coefficientsin Eq. (19) are those corresponding to the solution for a wing of thesame planform shape but without washout. The solution to Eq. (20)can be obtained in a similar manner and is also independent of angleof attack.

Truncating the series, the first N Fourier coefficients bn can beobtained by enforcing Eq. (20) at N different spanwise cross sectionsof the wing. With the first and last sections located at the wingtipsand the intermediate sections spaced equally in θ , this gives thefollowing system of equations:

N∑n = 1

bn

[4b

CL ,αc(θi )+ n

sin(θi )

]sin(nθi ) = ω(θi )

θi = (i − 1)π

N − 1, i = 1, N

After applying the following relations at the wingtips, θ = 0 andθ = π , [

sin(nθ)

sin(θ)

]θ → 0

= n

[sin(nθ)

sin(θ)

]θ → π

= (−1)n + 1n

we have

N∑n = 1

bnn2 = ω(0)

N∑n = 1

bn

[4b

CL ,αc(θi )+ n

sin(θi )

]sin(nθi ) = ω(θi )

θi = (i − 1)π

N − 1, i = 2, N − 1

N∑n = 1

bn(−1)n + 1n2 = ω(π) (21)

This N × N system is readily solved for the N Fourier coefficientsbn . It should be noted that the matrix obtained for the coefficientsof bn on the left-hand side of Eq. (21) is exactly the same as thatobtained from Eq. (19) for determination of the Fourier coefficientsan . Thus, this matrix needs to be inverted only once. To obtain theplanform Fourier coefficients an , the N × N inverted matrix is mul-tiplied by an N component column vector with all components setto 1. Multiplying the same N × N inverted matrix by an N compo-nent column vector obtained from the washout distribution functionyields the Fourier coefficients bn . If both the wing planform andwashout distribution function are spanwise symmetric, all of theeven Fourier coefficients in both an and bn will be zero. However,carrying both the odd and even Fourier coefficients and distribut-ing the collocation points over both sides of the wing allows forthe analysis of asymmetric twist, such as that produced by ailerondeflection. With modern tools, the added computational burden isinsignificant.

Other methods of evaluating the Fourier coefficients5,6 could alsobe used. As was shown by Bertin,9 the different methods give similaraccuracy for a given truncation level and produce identical resultswhen the number of terms carried from the infinite series becomeslarge. The results obtained from the present analysis do not dependon the method used to evaluate the Fourier coefficients, provided thata sufficient number of terms are carried. With today’s technology,100 to 1000 terms in the infinite series can be evaluated as fast asthe result can be written to the screen. Thus, carrying more terms inthe Fourier series is the easiest way to improve accuracy.

Using Eq. (18) in Eq. (7), the lift coefficient for a wing withwashout can be expressed as

CL = π RA A1 = π RA[a1(α − αL0)root − b1�] (22)

Using Eq. (18) in Eq. (7), the induced-drag coefficient is given by

CDi = π RA

∞∑n = 1

n A2n = π RA[a1(α − αL0)root − b1�]2

+ π RA

∞∑n = 2

n[an(α − αL0)root − bn�]2

or, in view of Eq. (22),

CDi = C2L

π RA+ π RA

∞∑n = 2

n[a2

n(α − αL0)2root

− 2anbn(α − αL0)root� + b2n�

2]

(23)

Equations (22) and (23) can be algebraically rearranged to yield

CL = CL ,α[(α − αL0)root − ε��] (24)

CDi = C2L(1 + κD) − κDLCLCL ,α� + κD�(CL ,α�)2

π RA(25)

where

CL ,α = π RAa1 = CL ,α

(1 + CL ,α/π RA)(1 + κL)(26)

κL ≡ 1 − (1 + π RA/CL ,α)a1

(1 + π RA/CL ,α)a1

(27)

ε� ≡ b1

a1(28)

PHILLIPS 131

κD ≡∞∑

n = 2

na2

n

a21

(29)

κDL ≡ 2b1

a1

∞∑n = 2

nan

a1

(bn

b1− an

a1

)(30)

κD� ≡(

b1

a1

)2 ∞∑n = 2

n

(bn

b1− an

a1

)2

(31)

Comparing Eqs. (24–31) with Eqs. (10–13), we see that washoutincreases the zero-lift angle of attack for any wing. However, thelift slope for a wing of arbitrary planform shape is not affected bywashout. Equations (24), (26), and (27) show that the lift slope for atwisted finite wing depends only on the airfoil section lift slope, thewing aspect ratio, and the first Fourier coefficient in the infinite seriesobtained from Eq. (19). From examination of Eq. (19), it is seen thatthe Fourier coefficients an depend only on wing planform and airfoilsection lift slope. No approximation was made in the developmentof these equations beyond those that are already included in Eq. (2),which is the foundation for all lifting-line theory. Thus, lifting-linetheory will always predict that the lift slope for an unstalled finitewing is independent of geometric and aerodynamic twist.

Also notice that the induced drag for a wing with washout is notzero at the angle of attack for which the wing develops zero netlift. In addition to the usual component of induced drag, which isproportional to the lift coefficient squared, a wing with washoutproduces a component of induced drag that is proportional to thewashout squared and this gives rise to induced drag at zero netlift. There is also a component of induced drag that varies with theproduct of the lift coefficient and the washout. The induced drag,which is produced on a twisted wing at zero net lift, still resultsfrom the production of lift on the wing. For a wing with geometricor aerodynamic twist, there is no global angle of attack for whichthe wing carries zero lift over all segments of its span. When the netlift on a twisted wing is zero, some spanwise segments of the wingare carrying positive lift while other spanwise segments are carryingnegative lift. Induced drag is generated as a result of the productionof this lift, even though the negative lift exactly balances the positivelift so that no net lift is generated by the wing.

Elliptic Wings with Linear WashoutAn elliptic wing with no geometric or aerodynamic twist provides

a planform shape that produces minimum possible induced drag fora given lift coefficient and aspect ratio. The chord of an elliptic wingvaries with the spanwise coordinate according to the relation

c(z) = (4b/π RA)√

1 − (2z/b)2

or

c(θ) = (4b/π RA) sin(θ) (32)

Using Eq. (32) in Eqs. (19) and (20) yields

∞∑n = 1

an

(π RA

CL ,α

+ n

)sin(nθ) = sin(θ) (33)

∞∑n = 1

bn

(π RA

CL ,α

+ n

)sin(nθ) = ω(θ) sin(θ) (34)

The well-known solution to Eq. (33) is

a1 = CL ,α

π RA + CL ,α

, an = 0, n �= 1 (35)

and the solution to Eq. (34) is given by the Fourier integral

bn = 2

π

(CL ,α

π RA + nCL ,α

)∫ π

0

ω(θ) sin(θ) sin(nθ) dθ (36)

For the case of linear washout, the normalized washout distributionfunction is

ω(z) = |2z/b| or ω(θ) = | cos(θ)| (37)

Using Eq. (37) in Eq. (36) and applying the trigonometric identitysin(2θ) = 2 sin(θ) cos(θ) yields

bn = 1

π

(CL ,α

π RA + nCL ,α

)[∫ π/2

0

sin(2θ) sin(nθ) dθ

−∫ π

π/2

sin(2θ) sin(nθ) dθ

](38)

The integrals on the right-hand side of Eq. (38) can be evaluatedfrom∫

sin(mθ) sin(nθ) dθ

=

sin[(n − m)θ ]

2(n − m)− sin[(n + m)θ ]

2(n + m), n �= m

θ

2− sin(2nθ)

4n, n = m

which yields

bn = 1

π

(CL ,α

π RA + nCL ,α

){4 sin(nπ/2)

4 − n2, n �= 2

0, n = 2 (39)

Using Eqs. (35) and (39) in Eqs. (27–31) gives

κL = 0, ε� = (4/3π), κD = 0, κDL = 0

κD� =∞∑

n = 3

n

[1

π

(π RA + CL ,α

π RA + nCL ,α

)4 sin(nπ/2)

4 − n2

]2

Using these results in Eqs. (24) and (25), together with the relation

sin2

(nπ

2

)=

{1, for n odd0, for n even

the lift and induced-drag coefficients for an elliptic wing with linearwashout are given by

CL = CL ,α

[(α − αL0)root − 4

3π�

]

CL ,α = CL ,α

1 + CL ,α/π RA

(40)

CDi = C2L + κD�(CL ,α�)2

π RA

κD� =∞∑

i = 1

16(2i + 1)

[(2i + 1)2 − 4]2π2

(π RA + CL ,α

π RA + (2i + 1)CL ,α

)2

(41)

From Eq. (41) it can be seen that linear washout always increasesinduced drag for an elliptic wing. The variation of κD� with aspectratio for an elliptic planform with linear washout and a section liftslope of 2π is shown in Fig. 3. This solution is consistent with theresults presented by Filotas.29

132 PHILLIPS

Fig. 3 Washout contribution to the induced-drag factor for ellipticwings with linear washout.

Tapered Wings with Linear WashoutWhereas elliptic wings with no geometric or aerodynamic twist

produce minimum possible induced drag, they are more expensiveto manufacture than simple rectangular wings. The tapered winghas commonly been used as a compromise. A tapered wing has achord that varies linearly with the spanwise coordinate according tothe following relation:

c(z) = 2b

RA(1 + RT )[1 − (1 − RT )|2z/b|]

or

c(θ) = 2b

RA(1 + RT )[1 − (1 − RT )| cos(θ)|] (42)

Using Eqs. (37) and (42) in Eqs. (19) and (20) yields the results fora tapered wing with linear washout:

∞∑n = 1

an

{2RA(1 + RT )

CL ,α[1 − (1 − RT )| cos(θ)|] + n

sin(θ)

}sin(nθ) = 1

(43)

∞∑n = 1

bn

{2RA(1 + RT )

CL ,α[1 − (1 − RT )| cos(θ)|] + n

sin(θ)

}sin(nθ) = | cos(θ)|

(44)

The solution obtained from Eq. (43) for the Fourier coefficientsan is commonly used for predicting the lift and induced-drag co-efficients for a tapered wing with no geometric or aerodynamictwist. This solution is discussed in most engineering textbooks onaerodynamics.8−12 Figures 4 and 5 show how the lift slope andinduced-drag factors vary with aspect ratio and taper ratio for ta-pered wings with no geometric or aerodynamic twist. In these andall following figures an airfoil section lift slope of 2π was used.The results shown in Fig. 5 have sometimes led to the conclusionthat a tapered wing with a taper ratio of about 0.4 always producesless induced drag than a rectangular wing of the same aspect ratiodeveloping the same lift. As a result, tapered wings are commonlyused as a means of reducing induced drag. However, this reduc-tion in drag comes at a price. Because a tapered wing has a lowerReynolds number at the wingtips than at the root, a tapered wingwith no geometric or aerodynamic twist tends to stall first in the re-gion near the wingtips. This wingtip stall can lead to poor handlingcharacteristics during stall recovery.

When interpreting the results shown in Fig. 5, it has proven easyto lose sight of the fact that these results can be used to predict totalinduced drag only for the special case of wings with no geometric

Fig. 4 The lift slope factor for tapered wings.

Fig. 5 Planform contribution to the induced-drag factor for taperedwings.

or aerodynamic twist. This is only one of many possible washoutdistributions that could be used for a wing of any given planform.Furthermore, it is not the washout distribution that produces mini-mum induced drag with finite lift, except for the special case of anelliptic planform. When the effects of washout are included, it can beshown that the conclusions sometimes reached from considerationof only those results shown in Fig. 5 are erroneous.

The induced-drag coefficient for a wing with washout can be pre-dicted from Eq. (25) with the definitions given in Eqs. (29–31). Fortapered wings with linear washout, the Fourier coefficients bn canbe obtained from Eq. (44) in exactly the same manner as the coeffi-cients an are obtained from Eq. (43). Using the Fourier coefficientsso obtained in Eqs. (28), (30), and (31) produces the results shownin Figs. 6–8. Notice from either Eq. (31) or Fig. 8 that κD� is alwayspositive. Thus, the third term in the numerator on the right-hand sideof Eq. (25) always contributes to an increase in induced drag. How-ever, from the results shown in Fig. 7, we see that the second term inthe numerator on the right-hand side of Eq. (25) can either increaseor decrease the induced drag, depending on the signs of κDL and �.This raises an important question regarding wing efficiency. Whatlevel and distribution of washout will result in minimum induceddrag for a given wing planform and lift coefficient?

Minimizing Induced Drag with WashoutFor a wing of any given planform shape with a fixed washout dis-

tribution function, the induced drag can be minimized with washoutas a result of the tradeoff between the second and third terms in thenumerator on the right-hand side of Eq. (25). Thus, Eq. (25) can be

PHILLIPS 133

Fig. 6 Washout effectiveness for tapered wings with linear washout.

Fig. 7 Lift-washout contribution to the induced-drag factor fortapered wings with linear washout.

Fig. 8 Washout contribution to the induced-drag factor for taperedwings with linear washout.

used to determine the optimum value of total washout, which willresult in minimum induced drag for any washout distribution andany specified lift coefficient. Differentiating Eq. (25) with respectto total washout at constant lift coefficient gives

∂CDi

∂�= −κDLCLCL ,α + 2κD�C2

L ,α�

π RA

From this result it can be seen that minimum induced drag is attainedfor any given wing planform, c(z), any given washout distribution,

ω(z), and any given design lift coefficient, CLd , by using an optimumtotal washout, �opt given by the relation

CL ,α�opt = κDLCLd

2κD�

(45)

Because κDL is zero for an elliptic wing, Eq. (45) shows that anelliptic wing is optimized with no washout. From consideration ofEq. (45), together with the results shown in Figs. 7 and 8, we see thattapered wings with linear washout and taper ratios greater than 0.4are optimized with positive washout, whereas those with taper ratiosless than about 0.4 produce minimum induced drag with negativewashout. However, a taper ratio less than 0.4 with negative washoutwould result in very poor stall characteristics and would not bepractical.

Using the value of optimum washout from Eq. (45) in the ex-pression for induced-drag coefficient that is given by Eq. (25), wefind that the induced-drag coefficient for a wing of arbitrary plan-form with a fixed washout distribution and optimum total washoutis given by

(CDi )opt = C2L

π RA

[1 + κD − κ2

DL

4κD�

(2 − CLd

CL

)CLd

CL

](46)

From Eq. (46) it can be seen that a wing with optimum washoutwill always produce less induced drag than a wing with no washouthaving the same planform and aspect ratio, provided that the actuallift coefficient is greater than one-half the design lift coefficient.When the actual lift coefficient is equal to the design lift coefficient,the induced-drag coefficient for a wing with optimum washout is

(CDi )opt = (C2

L

/π RA

)(1 + κDo), κDo ≡ κD − (

κ2DL

/4κD�

)(47)

For tapered wings with linear washout, the variations in κDo withaspect ratio and taper ratio are shown in Fig. 9. For comparison, thedashed lines in this figure show the same results for wings with nowashout. Notice that when linear washout is used to further optimizetapered wings, taper ratios in the range of 0.4 correspond closely toa maximum in induced drag, not to a minimum.

The choice of a linear washout distribution, which was used togenerate the results shown in Fig. 9, is as arbitrary as the choice of nowashout. Whereas a linear variation in washout is commonly usedand simple to implement, it is not the optimum washout distributionfor wings with linear taper. Minimum induced drag for a given finitelift coefficient occurs when the right-hand side of Eq. (5) varies withthe spanwise coordinate in proportion to sin(θ). This requires thatthe product of the local chord length and the local aerodynamic angle

Fig. 9 Optimum induced-drag factor for tapered wings with linearwashout.

134 PHILLIPS

of attack, α − αL0, varies elliptically with the spanwise coordinate;that is,

c(z)[α(z) − αL0(z)]√1 − (2z/b)2

= c(θ)[α(θ) − αL0(θ)]

sin(θ)= Const (48)

There are many possibilities for wing geometry that will satisfy thiscondition. The elliptic planform with no geometric or aerodynamictwist is only one such geometry. Because the local aerodynamicangle of attack decreases along the span in direct proportion to theincrease in washout, Eq. (48) can only be satisfied if the washoutdistribution satisfies the following relation:

c(z)[1 − ω(z)]√1 − (2z/b)2

= c(θ)[1 − ω(θ)]

sin(θ)= Const (49)

Equation (49) is satisfied by the spanwise washout distribution

ω(z) = 1 −√

1 − (2z/b)2

c(z)/crootor ω(θ) = 1 − sin(θ)

c(θ)/croot(50)

For wings with linear taper, this gives

ω(z) = 1 −√

1 − (2z/b)2

1 − (1 − RT )|2z/b|or

ω(θ) = 1 − sin(θ)

1 − (1 − RT )| cos(θ)| (51)

This washout distribution is shown in Fig. 10 for several values oftaper ratio. Results obtained for tapered wings with this washoutdistribution are presented in Figs. 11–13.

When an unswept wing of arbitrary planform has the washoutdistribution specified by Eq. (50), the value of κDo as defined inEq. (47) is always identically zero. With this spanwise washout dis-tribution and the total washout set according to Eq. (45), an unsweptwing of any planform shape can be designed to operate at a given liftcoefficient with the same minimum induced drag as that producedby an untwisted elliptic wing with the same aspect ratio and liftcoefficient. This optimum spanwise washout distribution producesan elliptic spanwise lift distribution, which always results in uni-form downwash over the wingspan and minimum possible induceddrag. For example, a rectangular wing requires an elliptic washoutdistribution according to Eq. (51). With this washout distribution,an aspect ratio of 8.0, and a section lift slope of 2π , Eqs. (29–31)and (47) give

κD = 0.0676113, κDL = 0.1379367

κD� = 0.0703526, κDo = 0.0000000

Fig. 10 Optimum washout distribution for wings with linear taper.

Fig. 11 Washout effectiveness for tapered wings with optimumwashout distribution.

Fig. 12 Lift-washout contribution to the induced-drag factor fortapered wings with optimum washout distribution.

Similarly, a tapered wing of aspect ratio 8.0 and taper ratio 0.5 resultsin

κD = 0.0171896, κDL = 0.0455690

κD� = 0.0302003, κDo = 0.0000000

These results were obtained by carrying a total of 99 terms in theinfinite series, that is, 50 of the nonzero odd terms.

Consider an unswept wing with linear taper having the washoutdistribution given by Eq. (51). Using Eqs. (42) and (51) in Eqs. (19)and (20) yields

∞∑n = 1

an

a1sin(nθ) + CL ,α[1 − (1 − RT )| cos(θ)|]

2RA(1 + RT )

∞∑n = 1

nan

a1

sin(nθ)

sin(θ)

= CL ,α[1 − (1 − RT )| cos(θ)|]2a1 RA(1 + RT )

(52)

∞∑n = 1

bn

b1sin(nθ) + CL ,α[1 − (1 − RT )| cos(θ)|]

2RA(1 + RT )

∞∑n = 1

nbn

b1

sin(nθ)

sin(θ)

= CL ,α[1 − sin(θ) − (1 − RT )| cos(θ)|]2b1 RA(1 + RT )

(53)

PHILLIPS 135

Fig. 13 Washout contribution to the induced-drag factor for taperedwings with optimum washout distribution.

To evaluate the induced drag for this wing we make use of the Fourierexpansions,

1 = 2

π

[2 sin(θ) +

∞∑m = 1

2

2m + 1sin[(2m + 1)θ ]

](54)

| cos(θ)| = 2

π

[sin(θ) +

∞∑m = 1

2m + 1 − (−1)m

2m(m + 1)sin[(2m + 1)θ ]

]

(55)

and the trigonometric identities,

sin[(2m + 1)θ ]

sin(θ)= sin[(2m − 1)θ ]

sin(θ)+ 2 cos(2mθ) (56)

2 cos(2mθ) sin[(2n + 1)θ ] = sin[(2m + 2n + 1)θ]

− sin[(2m − 2n − 1)θ ] (57)

Using Eqs. (54) through (57) in Eqs. (52) and (53), the Fouriercoefficients, bn , can be related to the Fourier coefficients, an . Theresult is

bn =

a1 − (1 − a1)CL ,α

2(1 + RT )RA, for n = 1

an

[1 + CL ,α

2(1 + RT )RA

], for n �= 1

(58)

Applying Eq. (58) together with Eqs. (26) and (29) to Eqs. (30)and (31), we find that a wing with linear taper and the washoutdistribution specified by Eq. (51) results in

κDL = πCL ,α

(1 + RT )CL ,α

κD (59)

κD� =(

πCL ,α

2(1 + RT )CL ,α

)2

κD (60)

Using Eqs. (59) and (60) in Eq. (25), the induced drag for a wingwith linear taper and the optimum washout distribution defined byEq. (51) can be written as

CDi = C2L

π RA+ κD

π RA

[CL − πCL ,α�

2(1 + RT )

]2

(61)

From Eq. (61), we see that any wing with linear taper and the corre-sponding optimum washout distribution specified by Eq. (51) pro-duces the same minimum induced drag as an elliptic wing with no

twist, provided that the total washout is related to the lift coefficientaccording to

�opt = 2(1 + RT )CL

πCL ,α

(62)

A wing of arbitrary planform with an optimum washout distri-bution produces exactly the same induced drag as an elliptic wingwith no washout, only while operating at the design lift coefficient.Whereas an elliptic wing produces minimum possible induced dragat all angles of attack, a washout-optimized wing of any other plan-form will produce minimum possible induced drag only at the angleof attack resulting in the design lift coefficient. However, a washout-optimized wing of nonelliptic planform very nearly duplicates theperformance of an elliptic wing over a wide range of lift coefficients.This is demonstrated in Fig. 14 by comparing the induced drag gen-erated by washout-optimized rectangular and tapered wings withthat produced by an untwisted elliptic wing. Each of the three wingsused for this figure has an aspect ratio of 8.0 and the rectangular andtapered wings are both optimized for a design lift coefficient of 0.4.The rectangular wing requires an elliptic washout distribution with4.64 deg of total washout at the wingtips. The tapered wing has ataper ratio of 0.4 and uses the corresponding washout distributionshown in Fig. 10, which has 3.25 deg of total washout at the wingtipsand maximum washin of 0.81 deg located at approximately 60% of

Fig. 14 Comparison of the induced drag produced by three wings ofaspect ratio 8.0; a rectangular wing and a tapered wing (RT = 0.4), bothof which are washout-optimized for a design lift coefficient of 0.4, andan elliptic wing without washout.

Fig. 15 Comparison of the induced drag produced by three wings ofaspect ratio 4.0; a rectangular wing and a tapered wing (RT = 0.4), bothof which are washout-optimized for a design lift coefficient of 0.4, andan elliptic wing without washout.

136 PHILLIPS

Fig. 16 Comparison of the induced drag produced by three wings ofaspect ratio 20.0; a rectangular wing and a tapered wing (RT = 0.4), bothof which are washout-optimized for a design lift coefficient of 0.4, andan elliptic wing without washout.

the semispan. Similar results for wings of aspect ratio 4.0 and 20.0are shown in Figs. 15 and 16.

ConclusionsWhen the traditional form of the lifting-line solution is used for

wings with washout, the Fourier coefficients in the infinite series alldepend on angle of attack. Because these coefficients must be reeval-uated for each operating point, this form of the solution is cumber-some for evaluating traditional wing properties. A more practicalform of the lifting-line solution for finite twisted wings has beenpresented here. This analytical solution is based on a change ofvariables that allows the infinite series to be divided into two parts,one series that depends only on planform and another series thatdepends on both planform and washout. The Fourier coefficients inboth of these infinite series depend only on wing geometry and canbe evaluated independent of the angle of attack.

The present form of the lifting-line solution for twisted wingsshows that a wing of any planform shape can be optimized for anygiven design lift coefficient to produce exactly the same induceddrag as an elliptic wing with no geometric or aerodynamic twist.The lifting-line theory predicts that minimum possible induced dragoccurs whenever the product of local chord length and local aero-dynamic angle of attack varies elliptically with the spanwise coor-dinate. The traditional solution to satisfying this requirement is theclassic elliptic wing, which has a chord length that varies ellipti-cally with the spanwise coordinate and constant aerodynamic angleof attack. However, this is only one of many possibilities for winggeometry that will satisfy the requirement for minimum possible in-duced drag. Another obvious solution is a wing with constant chordhaving a washout distribution that produces an elliptic variation inaerodynamic angle of attack. By using the more general washoutdistribution described in Eq. (50), together with the optimum totalwashout specified by Eq. (45), a wing of any planform can be madeto satisfy the requirement for minimum possible induced drag.

Minimizing induced drag using both chord length and washoutvariation provides greater flexibility than using chord-length varia-tion alone. For example, the stall characteristics of a wing depend onboth the planform shape and the washout distribution. By using theoptimum washout distribution, minimum induced drag can be main-tained at the design lift coefficient for any planform shape. Thus, anadditional degree-of-freedom is provided that can be used to helpcontrol the stall characteristics or any other measure of performancethat is affected by wing planform and/or washout distribution.

From the results presented here, it can be seen that a wing ofany planform shape having no geometric or aerodynamic twist isa special case of a washout-optimized wing. Except for the caseof an elliptic planform, a wing with no washout is optimized for a

design lift coefficient of zero. With the wing operating at any otherlift coefficient, washout could be used to reduce the induced drag atlow Mach numbers.

References1Prandtl, L., “Tragflugel Theorie,” Nachricten von der Gesellschaft

der Wissenschaften zu Gottingen, Ges-chaeftliche Mitteilungen, Klasse,Germany, 1918, pp. 451–477.

2Prandtl, L., “Applications of Modern Hydrodynamics to Aeronautics,”NACA TR-116, June 1921.

3Glauert, H., “The Monoplane Aerofoil,” The Elements of Aerofoiland Airscrew Theory, Cambridge Univ. Press, Cambridge, UK, 1926,pp. 137–155.

4Multhopp, H., “Die Berechnung der Auftriebs Verteilung vonTragflugeln,” Luftfahrtforschung, Vol. 15, No. 14, 1938, pp. 153–169.

5Rasmussen, M. L., and Smith, D. E., “Lifting-Line Theory for ArbitrarilyShaped Wings,” Journal of Aircraft, Vol. 36, No. 2, 1999, pp. 340–348.

6Lotz, I., “Berechnung der Auftriebsverteilung beliebig geformterFlugel,” Zeitschrift fur Flugtechnik und Motorluftschiffahrt, Vol. 22, No. 7,1931, pp. 189–195.

7Karamcheti, K., “Elements of Finite Wing Theory,” Ideal-Fluid Aerody-namics, Wiley, New York, 1966, pp. 535–567.

8Anderson, J. D., “Incompressible Flow over Finite Wings: Prandtl’sClassical Lifting-Line Theory,” Fundamentals of Aerodynamics, 3rd ed.,McGraw–Hill, New York, 2001, pp. 360–387.

9Bertin, J. J., “Incompressible Flow About Wings of Finite Span,” Aero-dynamics for Engineers, 4th ed., Prentice–Hall, Upper Saddle River, NJ,2002, pp. 230–302.

10Katz, J., and Plotkin, A., “Finite Wing: The Lifting-Line Model,” Low-Speed Aerodynamics, 2nd ed., Cambridge Univ. Press, Cambridge, UK,2001, pp. 167–183.

11Kuethe, A. M., and Chow, C. Y., “The Finite Wing,” Foundations ofAerodynamics, 5th ed., Wiley, New York, 1998, pp. 169–219.

12McCormick, B. W., “The Lifting-Line Model,” Aerodynamics, Aero-nautics, and Flight Mechanics, 2nd ed., Wiley, New York, 1995,pp. 112–119.

13Glauert, H., and Gates, S. B., “The Characteristics of a Tapered andTwisted Wing with Sweep-Back,” Reports and Memoranda 1226, Aeronau-tical Research Council, London, Aug. 1929.

14Amstutz, E., “Calculation of Tapered Monoplane Wings,” NACA TM-578, Feb. 1930.

15Anderson, R. F., “Charts for Determining the Pitching Moment of Ta-pered Wings with Sweepback and Twist,” NACA TN-483, Dec. 1933.

16Anderson, R. F., “Determination of the Characteristics of TaperedWings,” NACA TR-572, May 1937.

17Datwiler, G., “Calculations of the Effect of Wing Twist on the Air ForcesActing on a Monoplane Wing,” NACA TN-520, March 1935.

18Cohen, D., “A Method of Determining the Camber and Twist of a Sur-face to Support a Given Distribution of Lift,” NACA TN-855, Aug. 1942.

19Falkner, V. M., “The Calculation of Aerodynamic Loading on Sur-faces of Any Shape,” Reports and Memoranda 1910, Aeronautical ResearchCouncil, London, Aug. 1943.

20DeYoung, J., and Harper, C. W., “Theoretical Symmetric Span Loadingat Subsonic Speeds for Wings Having Arbitrary Plan Form,” NACA TR-921,Dec. 1948.

21Stevens, V. I., “Theoretical Basic Span Loading Characteristics ofWings with Arbitrary Sweep, Aspect Ratio and Taper Ratio,” NACA TN-1772, Dec. 1948.

22Munk, M. M., “On the Distribution of Lift Along the Span of an Airfoilwith Displaced Ailerons,” NACA TN-195, June 1924.

23Munk, M. M., “A New Relation Between the Induced Yawing Momentand the Rolling Moment of an Airfoil in Straight Motion,” NACA TR-197,June 1925.

24Glauert, H., “Theoretical Relationships for an Airfoil with Hinged Flap,”Reports and Memoranda 1095, Aeronautical Research Council, London,July 1927.

25Hartshorn, A. S., “Theoretical Relationship for a Wing with UnbalancedAilerons,” Reports and Memoranda 1259, Aeronautical Research Council,London, Oct. 1929.

26Pearson, H. A., “Theoretical Span Loading and Moments of TaperedWings Produced by Aileron Deflection,” NACA TN-589, Jan. 1937.

27Pearson, H. A., “Span Load Distribution for Tapered Wings with Partial-Span Flaps,” NACA TR-585, Nov. 1937.

28Pearson, H. A., and Jones, R. T., “Theoretical Stability and ControlCharacteristics of Wings with Various Amounts of Taper and Twist,” NACATR-635, April 1937.

29Filotas, L. T., “Solution of the Lifting Line Equation for Twisted EllipticWings,” Journal of Aircraft, Vol. 8, No. 10, 1971, pp. 835–836.