Optimum Shape Design for Unsteady Flows with Time...

14
Optimum Shape Design for Unsteady Flows with Time-Accurate Continuous and Discrete Adjoint Methods Siva K. Nadarajah McGill University, Montreal, Quebec H3A 2S6, Canada and Antony Jameson Stanford University, Stanford, California 94305 DOI: 10.2514/1.24332 This paper presents an adjoint method for the optimal control of unsteady ows. The goal is to develop the continuous and discrete unsteady adjoint equations and their corresponding boundary conditions for the time- accurate method. First, this paper presents the complete formulation of the time-dependent optimal design problem. Second, we present the time-accurate unsteady continuous and discrete adjoint equations. Third, we present results that demonstrate the application of the theory to a two-dimensional oscillating airfoil. The results are compared with a multipoint approach to illustrate the added benet of performing full unsteady optimization. Nomenclature A = ux Jacobian matrix B = boundary b = boundary velocity component c = chord length D = articial dissipation ux E = internal energy F = Euler numerical ux vector f = Euler ux vector G = gradient I = cost function i, j = cell indices M = Mach number N = outward-facing normal p = pressure R = residual S = shape function S = face areas of the computational cell s = arc length T = time period t = time t f = nal time u = velocity (physical domain) V = cell volume w = state vector x = coordinates (physical domain) = angle of attack = adjustable constant for articial dissipation scheme = step size = coordinates (computational domain) = density = Lagrange multiplier ! r = reduced frequency I. Introduction I N THE past decade, aerodynamic shape optimization has been the focus of attention due largely to advanced algorithms that have allowed researchers to calculate gradients cheaply and efciently. The majority of work in aerodynamic shape optimization has been focused on the design of aerospace vehicles in a steady ow environment. Investigators have applied these advanced design algorithms, particularly the adjoint method, to numerous problems, ranging from the design of two-dimensional airfoils to full aircraft congurations to decrease drag, increase range, and reduce sonic boom [111]. These problems were tackled using many different numerical schemes on both structured and unstructured grids [1215]. Unlike xed-wing aircraft, helicopter rotors and turbomachinery blades operate in unsteady ows and are constantly subjected to unsteady loads. Therefore, optimal control techniques for unsteady ows are needed to improve the performance of helicopter rotors and turbomachinery and to alleviate the unsteady effects that contribute to utter, buffeting, poor gust and acoustic response, and dynamic stall. Recently, the design of blade proles using unsteady techniques was attempted. Diverse methods were employed in the design of rotorcraft and turbomachinery blades. The following are a selected number of papers on this topic. Ghayour and Baysal [16] solved the unsteady transonic small-disturbance equation and its continuous adjoint equation to perform an inverse design at Mach 0.6. Aerodynamic shape optimization of rotor airfoils in an unsteady viscous ow was attempted by Yee et al. [17] using a response- surface methodology. Here the authors used an upwind-biased- factorized implicit numerical scheme to solve the Reynolds- averaged NavierStokes equations with a BaldwinLomax turbulence model. A response-surface methodology was then employed to optimize the rotor blade. The objective function was a sum of the L=D at three different azimuth angles and was later redened to include unsteady aerodynamic effects. Florea and Hall [18] modeled a cascade of turbomachinery blades using the steady and time-linearized Euler equations. Gradients for aeroelastic and aeroacoustic objective functions were then computed using the discrete adjoint approach. Both the ow and adjoint equations were solved using a nite volume LaxWendroff scheme. The gradients were then used to improve the aeroelastic stability and acoustic response of the airfoil. Presented as Paper 5436 at the 9th AIAA/ISSMO Symposium on Multidisciplinary Analysis and Optimization Conference, Atlanta, GA, 46 September 2002; received 29 March 2006; revision received 29 November 2006; accepted for publication 18 January 2007. Copyright © 2007 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved. 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 0001-1452/ 07 $10.00 in correspondence with the CCC. Assistant Professor, Computational Fluid Dynamics Lab, Department of Mechanical Engineering, 688 Sherbrooke Street West, Room 711; [email protected]. Thomas V. Jones Professor of Engineering, Department of Aeronautics Astronautics, Durand Building, 496 Lomita Mall. Fellow AIAA. AIAA JOURNAL Vol. 45, No. 7, July 2007 1478

Transcript of Optimum Shape Design for Unsteady Flows with Time...

Page 1: Optimum Shape Design for Unsteady Flows with Time …aero-comlab.stanford.edu/Papers/AIAA-24332-153.pdfAerodynamic shape optimization of rotor airfoils in an unsteady viscous flow

Optimum Shape Design for Unsteady Flows with Time-AccurateContinuous and Discrete Adjoint Methods

Siva K. Nadarajah∗

McGill University, Montreal, Quebec H3A 2S6, Canada

and

Antony Jameson†

Stanford University, Stanford, California 94305

DOI: 10.2514/1.24332

This paper presents an adjoint method for the optimal control of unsteady flows. The goal is to develop the

continuous and discrete unsteady adjoint equations and their corresponding boundary conditions for the time-

accuratemethod. First, this paper presents the complete formulation of the time-dependent optimal design problem.

Second, we present the time-accurate unsteady continuous and discrete adjoint equations. Third, we present results

that demonstrate the application of the theory to a two-dimensional oscillating airfoil. The results are comparedwith

a multipoint approach to illustrate the added benefit of performing full unsteady optimization.

Nomenclature

A = flux Jacobian matrixB = boundaryb = boundary velocity component�c = chord lengthD = artificial dissipation fluxE = internal energyF = Euler numerical flux vectorf = Euler flux vectorG = gradientI = cost functioni, j = cell indicesM = Mach numberN = outward-facing normalp = pressureR = residualS = shape functionS = face areas of the computational cells = arc lengthT = time periodt = timetf = final timeu = velocity (physical domain)V = cell volumew = state vectorx = coordinates (physical domain)� = angle of attack� = adjustable constant for artificial dissipation scheme� = step size� = coordinates (computational domain)� = density = Lagrange multiplier

!r = reduced frequency

I. Introduction

I NTHEpast decade, aerodynamic shape optimization has been thefocus of attention due largely to advanced algorithms that have

allowed researchers to calculate gradients cheaply and efficiently.The majority of work in aerodynamic shape optimization has beenfocused on the design of aerospace vehicles in a steady flowenvironment. Investigators have applied these advanced designalgorithms, particularly the adjoint method, to numerous problems,ranging from the design of two-dimensional airfoils to full aircraftconfigurations to decrease drag, increase range, and reduce sonicboom [1–11]. These problems were tackled using many differentnumerical schemes on both structured and unstructured grids [12–15].

Unlike fixed-wing aircraft, helicopter rotors and turbomachineryblades operate in unsteady flows and are constantly subjected tounsteady loads. Therefore, optimal control techniques for unsteadyflows are needed to improve the performance of helicopter rotors andturbomachinery and to alleviate the unsteady effects that contributeto flutter, buffeting, poor gust and acoustic response, and dynamicstall.

Recently, the design of blade profiles using unsteady techniqueswas attempted. Diverse methods were employed in the design ofrotorcraft and turbomachinery blades. The following are a selectednumber of papers on this topic. Ghayour and Baysal [16] solved theunsteady transonic small-disturbance equation and its continuousadjoint equation to perform an inverse design at Mach 0.6.Aerodynamic shape optimization of rotor airfoils in an unsteadyviscous flow was attempted by Yee et al. [17] using a response-surface methodology. Here the authors used an upwind-biased-factorized implicit numerical scheme to solve the Reynolds-averaged Navier–Stokes equations with a Baldwin–Lomaxturbulence model. A response-surface methodology was thenemployed to optimize the rotor blade. The objective function was asum of the L=D at three different azimuth angles and was laterredefined to include unsteady aerodynamic effects. Florea and Hall[18] modeled a cascade of turbomachinery blades using the steadyand time-linearized Euler equations. Gradients for aeroelastic andaeroacoustic objective functions were then computed using thediscrete adjoint approach. Both the flow and adjoint equations weresolved using a finite volume Lax–Wendroff scheme. The gradientswere then used to improve the aeroelastic stability and acousticresponse of the airfoil.

Presented as Paper 5436 at the 9th AIAA/ISSMO Symposium onMultidisciplinary Analysis and Optimization Conference, Atlanta, GA, 4–6September 2002; received 29 March 2006; revision received 29 November2006; accepted for publication 18 January 2007. Copyright © 2007 by theAmerican Institute of Aeronautics and Astronautics, Inc. All rights reserved.Copies of this paper may be made for personal or internal use, on conditionthat the copier pay the $10.00 per-copy fee to theCopyright Clearance Center,Inc., 222RosewoodDrive,Danvers,MA01923; include the code 0001-1452/07 $10.00 in correspondence with the CCC.

∗Assistant Professor, Computational Fluid Dynamics Lab, Department ofMechanical Engineering, 688 Sherbrooke Street West, Room 711;[email protected].

†Thomas V. Jones Professor of Engineering, Department of AeronauticsAstronautics, Durand Building, 496 Lomita Mall. Fellow AIAA.

AIAA JOURNALVol. 45, No. 7, July 2007

1478

Page 2: Optimum Shape Design for Unsteady Flows with Time …aero-comlab.stanford.edu/Papers/AIAA-24332-153.pdfAerodynamic shape optimization of rotor airfoils in an unsteady viscous flow

Traditionally, a multipoint design approach is one possibletechnique for the optimization of blade profiles in an unsteady flowenvironment. This approach only requires a small extension of asteady flow design code to redesign a blade or airfoil profile formultiple flow conditions. Because the steady flow equations are usedto design the blades, uncertainties still prevail surrounding theperformance of these blades in an unsteady flow environment.

In this paper, we develop a framework in which to perform twomajor tasks: first, to perform sensitivity analysis in a nonlinearunsteady flow environment; second, to further modify the shape ofthe object to achieve the objective of the design using a full unsteadyoptimization method based on control theory. Optimal control oftime-dependent trajectories is generally complicated by the need tosolve the adjoint equation in reverse time from a final boundarycondition using information from the trajectory solution, which inturn depends on the control derived from the adjoint solution. In thiswork, we extend the adjoint method to unsteady periodic flows usinga time-accurate approach. The time-accurate unsteady adjointequations are based on Jameson’s [19] cell-centered multigrid-driven fully implicit scheme with upwind-biased, blended, first- andthird-order artificial-dissipation fluxes.

The goal of this research is to develop both the time-accuratecontinuous and discrete adjoint equations and use them in theredesign of the RAE 2822 airfoil undergoing a pitching oscillation toachieve lower time-averaged drag while keeping the time-averagedlift constant. This technique is compared with the multipoint andsteady adjoint approaches to gauge the effectiveness of the method.

II. Governing Equations

The Euler equations for a rigidly translating control volume �,defined by boundary @� with an outward-facing normal N, can bewritten in integral form as

d

dt

ZZ�

w dx1 dx2 �I@�

fiNi ds� 0 (1)

The state vectorw and a component of the inviscidflux vector,fi, canbe written as

w�

8>><>>:

��u1�u2�E

9>>=>>;; fi �

8>><>>:

��ui � bi��u1�ui � bi� � �1ip�u2�ui � bi� � �2ip�E�ui � bi� � pui

9>>=>>;

In these equations, x1 and x2 are the Cartesian coordinates; ui and biare the Cartesian velocity components of the fluid and boundary,respectively; and E is the total energy. The results presented in thispaper are based on transonic flow calculations in which the ideal gasequation is applicable. Consequently, the pressure can be expressedas

p� �� � 1���E � 1

2�uiui�

A. Numerical Discretization of Governing Equations

A finite-volume methodology is used to discretize the integralform of the conservation laws. When using a discretization on abody-conforming structured mesh, it is useful to consider atransformation to the computational coordinates �1 and �2, definedby the metrics

Kij ��@xi@�j

�; J� det�K�; K�1ij �

�@�i@xj

The simulations contained in this research are restricted to rigidmeshtranslation. Consequently, the volumetric integral from Eq. (1) canbe approximated as the product of the cell volume and the temporalderivative of the solution at the cell center. The Euler equations canthen be written in computational space as

V@w

@t� @Fi@�i� 0 (2)

where the convective flux is now defined with respect to thecomputational cell faces by Fi � Sijfj, and the quantity Sij � JK�1ijrepresents the projection of the �i cell face along the xj axis. WhenEq. (2) is formulated for each computational cell, a system of first-order ordinary differential equations is obtained. Equation (2) canthen be written for each computational cell in semidiscrete form as

V@w

@t� R�w�i;j � 0 (3)

where

R�w�ij �@F1

@�1� @F2

@�2� fi�1

2;j� fi�12;j � fi;j�1

2� fi;j�12 (4)

The � 12notation indicates that the quantity is calculated at the flux

faces. To stabilize the scheme, a blended mix of first- and third-orderfluxes first introduced by Jameson et al. [20] is added to theconvective flux at each cell face and can be defined as

di�12;j� �2

i�12;j�wi�1;j � wi;j�

� �4i�1

2;j�wi�2;j � 3wi�1;j � 3wi;j � wi�1;j� (5)

Thefirst term inEq. (5) is thefirst-order diffusion term,where �2i�1

2;jis

proportional to the normalized second derivative of pressure. Thisterm is dominant in the vicinity of shocks and serves to damposcillations and overshoots associated with this discontinuity in thesolution field. The �4

i�12;jcoefficient scales themagnitude of the third-

order dissipative flux. The coefficient is scaled so that it becomes thedominant term away from the shock, eliminating the odd–evendecoupling associated with central-difference schemes.

The time-derivative term can be approximated by a kth-orderimplicit backward-difference formula (BDF) such as

d

dt� 1

�t

Xkq�1

1

q����q (6)

where �� � wn�1i;j � wni;j. A second-order expansion of Eq. (6) will

result in the following equation for the semidiscrete form (3) of thegoverning equations:

Vi;j

�3

2�twn�1i;j �

2

�twni;j �

1

2�twn�1i;j

�� R

�wn�1i;j

�� 0 (7)

Equation (7) represents a set of highly nonlinear, coupled, ordinarydifferential equations and can be solved at each time step using theexplicit multistage modified Runge–Kutta scheme. We define a newmodified residual R�wi;j� as

R�wi;j� � Vi;j�

3

2�twn�1i;j �

2

�twni;j �

1

2�twn�1i;j

�� R

�wn�1i;j

(8)

where the modified residual R�wi;j� is the sum of the steady-stateresidual and a source term that originates from the second-orderdiscretization of the time derivative. The modified residual is thenmarched to pseudosteady state in a fictitious time t, as follows:

dwi;jdt� R�wi;j� � 0

The procedure advances the solution forward in time from thet� n�t to t� �n� 1��t. The final time-accurate solution will becomposed of a sequence of pseudotime steady-state solutions.

NADARAJAH AND JAMESON 1479

Page 3: Optimum Shape Design for Unsteady Flows with Time …aero-comlab.stanford.edu/Papers/AIAA-24332-153.pdfAerodynamic shape optimization of rotor airfoils in an unsteady viscous flow

III. General Formulation of the Time-DependentOptimal Design Problem

The aerodynamic properties that define the cost function arefunctions of the flowfield variables, w, and the physical location ofthe boundary, which may be represented by the function S. We thenintroduce the cost function:

I � 1

T

Ztf

0

L�w;S� dt�M�w�tf�� (9)

The cost function is a sum of a time-averaged functionL�w;S� and afunctionM that is a function of the solutionw�t� at the final time. Achange in S then results in a change

�I � 1

T

Ztf

0

�@LT

@w�w� @L

T

@S�S�dt� @M

T

@w�w�tf� (10)

in the cost function. Using control theory, the governing equations ofthe flowfield are now introduced as a constraint in such away that thefinal expression for the gradient does not require reevaluation of theflowfield. To achieve this, �w must be eliminated from Eq. (10).From Eq. (3), a variation of the semidiscrete form of the governingequations can be written as

V@

@t�w�

�@R

@w

��w�

�@R

@S

��S � 0

Next, introduce a Lagrange multiplier to the time-dependent flowequation and integrate it over time to yield

1

T

Ztf

0

T�V@

@t�w�

�@R

@w

��w�

�@R

@S

��S�� 0 (11)

Subtract Eq. (11) from the variation of the cost function to arrive atthe following equation:

�I � 1

T

Ztf

0

�@LT

@w�w� @L

T

@S�S�dt� @M

T

@w�w�tf�

� 1

T

Ztf

0

T�V@

@t�w�

�@R

@w

��w�

�@R

@S

��S�dt (12)

Next, collect the �w and �S terms and integrate

1

T

Ztf

0

TV@

@t�w dt

by parts, to yield

�I� 1

T

Ztf

0

�@LT

@w�V @

T

@t� T

�@R

@w

���wdt

��@MT

@w�VT T�tf�

��w�tf��

1

T

Ztf

0

�@LT

@S� T

�@R

@S

���S dt

Choose to satisfy the adjoint equation

V@

@t��@R

@w

�T

�� @L@w

with the terminal boundary condition

V

T �tf� �

@M@w

Then the variation of the cost function reduces to

�I � GT�S

where

G T � 1

T

Ztf

0

�@LT

@S� T @R

@S

�dt

Optimal control of time-dependent trajectories is generallycomplicated by the need to solve the adjoint equation in reverse timefrom a final boundary condition using data from the trajectorysolution, which in turn depends on the control derived from theadjoint solution. The sensitivities are determined by the solution ofthe adjoint equation in reverse time from the terminal boundarycondition and the time-dependent solution of the flow equation.These sensitivities are then used to get a direction of improvementand steps are taken until convergence is achieved.

IV. Time-Accurate Continuous Adjoint Equation

To formulate the time-accurate continuous adjoint equation, wefirst introduce the cost function. In this work, the cost function ischosen to be the time-averaged drag coefficient and is defined as

I � 1

T

Ztf

0

Cd dt�1

T

Ztf

0

Ca cos �� Cn sin� dt

� 112�p1M

21 �cT

Ztf

0

ZBp

�@x2@�

cos � � @x1@�

sin�

�d� dt

where Ca and Cn are the axial and normal force coefficients,respectively. The design problem is now treated as a control problemin which the control function is the airfoil shape, which is chosen tominimize I subject to the constraints defined by theflow equations. Avariation in the shape causes a variation �p in the pressure and,consequently, a variation in the cost function:

�I � 112�p1M

21 �cT

Ztf

0

ZB�p

�@x2@�

cos� � @x1@�

sin�

� p��

�@x2@�

�cos� � �

�@x1@�

�sin�

�d� dt (13)

Because p depends on w through the equation of state, then thevariation �p is determined from the variation �w. From Eq. (2), thevariation of the governing equation can be written as

V@

@t�w� @

@�i�Fi � 0

where �Fi � Ci�w� �Sijfj. The Euler Jacobian matrix in thecomputational domain Ci is defined as Ci � SijAj, where Ai �@fi=@w is the Euler Jacobian matrix. Multiplying by a costate vector , also known as a Lagrange multiplier, and integrating over thespace and time produces

1

T

Ztf

0

ZD T�V@

@t�w� @

@�i�Ci�w� �Sijfj�

�dD dt� 0

If we confine the simulations to rigid mesh translation, then we canseparate the equation into two terms and switch the order of thedomain and time integrals for the first term, to yield

ZD

1

T

Ztf

0

TV@

@t�w dt dD

� 1

T

Ztf

0

ZD T

@

@�i�Ci�w� �Sijfj� dD dt� 0

If is differentiable, then the two terms in the preceding equation canbe integrated by parts, to give

ZD

�1

T�V T�w�tf0 �

1

T

Ztf

0

V@ T

@t�w dt

�dD

� 1

T

Ztf

0

�ZBni

T�Ci�w� �Sijfj� dB

�ZD

@ T

@�i�Ci�w� �Sijfj� dD

�dt� 0

The next procedure is to rearrange the terms in the equation such that

1480 NADARAJAH AND JAMESON

Page 4: Optimum Shape Design for Unsteady Flows with Time …aero-comlab.stanford.edu/Papers/AIAA-24332-153.pdfAerodynamic shape optimization of rotor airfoils in an unsteady viscous flow

integrands that are multiplied by the variation of the state vector, �w,are grouped together and terms that are multiplied by the variation ofthe metric terms are separated into a different integral. Thisprocedure is crucial to isolate the integral that will produce the time-accurate continuous adjoint equation:

ZD

1

T�V T�tf��w�tf� �V T�0��w�0��dD�

1

T

Ztf

0

ZDV@ T

@t�w

� @ T

@�iCi�wdDdt� 1

T

Ztf

0

ZBni

T�Fi dB

� 1

T

Ztf

0

�ZBni

T�Sijfj dB�ZD

@ T

@�i�Sijfj dD

�dt� 0

Because the left-hand expression equals zero, it may be subtractedfrom the variation of the cost function (13) to give

�I � 112�p1M

21 �cT

Ztf

0

ZBW

�p

�@x2@�

cos� � @x1@�

sin�

� p��

�@x2@�

�cos� � �

�@x1@�

�sin�

�d� dt

�ZD

1

T�V T�tf��w�tf� � V T�0��w�0�� dD

� 1

T

Ztf

0

ZD

�V@ T

@t� @

T

@�iCi

��w dD dt

� 1

T

Ztf

0

ZBni

T�Fi dB �1

T

Ztf

0

�ZBni

T�Sijfj dB

�ZD

@ T

@�i�Sijfj dD

�dt (14)

Because is an arbitrary differentiable function, it may be chosen insuch away that �I no longer depends explicitly on the variation of thestate vector, �w. The gradient of the cost function can then beevaluated directly from the metric variations without having torecompute the variation �w resulting from the perturbation of eachdesign variable. The variation �w can then be eliminated by solvingfor the Lagrange multiplier by setting the transpose of theintegrand of the second integral in the third line of Eq. (14) to zero toproduce a differential adjoint system governing :

V@

@t� CTi

@

@�i� 0 in D (15)

The convective flux of the time-accurate continuous adjointequation (15) is discretized using a second-order central spatialdiscretization. The temporal discretization is based on a second-orderbackward-difference formula. Artificial dissipation based on theJameson et al. [20] scheme, similar to that employed for the flowsolver, is added to stabilize the scheme. Refer to Nadarajah [21] for amore detailed overview of the numerical discretization. Apseudotime derivative is added and the time-accurate adjointequation is marched to a periodic steady-state solution.

At the outer boundary, incoming characteristics for correspondto outgoing characteristics for �w. Consequently, we can chooseboundary conditions for such that

ni TCi�w� 0

If the coordinate transformation is such that �S is negligible in the farfield, then the only remaining boundary term is

�ZBW

T�F2 d�1

Thus, by letting satisfy the boundary condition, then

jnj �1

12�p1M

21 �c

�@x2@�

cos� � @x1@�

sin�

�on BW (16)

where nj are the components of the surface normal. Because theinitial condition for the Lagrange multipliers are set to zero, thenV T�0��w�0� � 0. If the problem is periodic in nature and the costfunction used for this problem is not dependent upon tf, thenV T�tf��w�tf� � 0. Equation (14) finally reduces to the following:

�I� 112�p1M

21 �cT

Ztf

0

ZBW

p

��

�@x2@�

�cos�� �

�@x1@�

�sin�

�d�dt

� 1

T

Ztf

0

�ZBni

T�Sijfj dB�ZD

@ T

@�i�Sijfj dD

�dt

The preceding equation is then used to solve for the gradient, whichcan then provide a direction of improvement to reduce the objectivefunction.

V. Time-Accurate Discrete Adjoint Equation

As in the case of steady flow, the time-accurate discrete adjointequation is obtained by applying control theory directly to the set oftime-accurate discrete field equations. The resulting equationdepends on the type of scheme used to solve the flow equations. Toformulate the discrete adjoint equation, we first take a variation ofEq. (8) with respect to w and S (only terms that are multiplied to �ware shown):

�Rn�1i;j �w� � Vi;j

�3

2�t�wn�1i;j �

2

�t�wni;j �

1

2�t�wn�1i;j

� �Rn�1i;j �w� (17)

Next, multiply the preceding equation by the transpose of theLagrange multiplier and sum over the domain and time, to yield

1

T

Xtft�0

X�

Ti;j�Ri;j � �

1

T� Tn�1i;j �R

n�1i;j

� Tn�2i;j �Rn�2i;j � T

n�3i;j �R

n�3i;j � � (18)

Substitute Eq. (17) into the n� 1, n� 2, and n� 3 terms of themodified residual in the preceding equation, to yield

1

T

Xtft�0

X�

Ti;j�Ri;j � �

Tn�1i;j

TVi;j

�3

2�t�wn�1i;j �

2

�t�wni;j

� 1

2�t�wn�1i;j � �Rn�1i;j

�� T

n�2i;j

TVi;j

�3

2�t�wn�2i;j �

2

�t�wn�1i;j

� 1

2�t�wni;j � �Rn�2i;j

�� T

n�3i;j

TVi;j

�3

2�t�wn�3i;j �

2

�t�wn�2i;j

� 1

2�t�wn�1i;j � �Rn�3i;j

��

Keeping only the n� 1 terms, the preceding equation reduces to

1

T

Xtft�0

X�

Ti;j�Ri;j � �

Vi;jT

�3

2�t T

n�1i;j �

2

�t T

n�2i;j

� 1

2�t T

n�3i;j

��wn�1i;j �

1

T T

n�1i;j �Rn�1i;j � (19)

Next, we introduce the discrete cost function for the dragminimization problem as

Ic �1

T

Xtft�0

Cd�t�1

T

Xtft�0�Ca cos�� Cn sin���t

� 112�p1M

21 �cT

Xtft�0

XUTEi�LTE

pi;W��x2�i cos� ��x1�i sin���t

where LTE is the lower trailing edge, UTE is the upper trailing edge,

NADARAJAH AND JAMESON 1481

Page 5: Optimum Shape Design for Unsteady Flows with Time …aero-comlab.stanford.edu/Papers/AIAA-24332-153.pdfAerodynamic shape optimization of rotor airfoils in an unsteady viscous flow

and pi;W is the wall pressure. In this research, the wall pressure isdefined as

pi;W �1

2�pi;2 � pi;1�

wherepi;2 andpi;1 are the values of the pressure in the cell above andbelow the wall. A variation in the cost function will result in avariation �p in the pressure and variations �x2� and �x1� in themetrics. The variation of the cost function for drag minimization canbe written as

�Ic �1

12�p1M

21 �cT

�Xtft�0

�XUTEi�LTE

1

2��x2�i cos� ��x1�i sin��

@p

@w��wi;2 � �wi;1�

�XUTEi�LTE

�1

2�pi;2 � pi;1� � p1

��cos����x2�i�

� sin����x1�i����t (20)

The time-dependent discrete Euler equations can now be introducedinto �I as a constraint, to produce

�I � �Ic �1

T

Xtft�0

X�

Ti;j�Ri;j�w� (21)

Substitute Eqs. (19) and (20) into the preceding expression, whichcan then be rearranged into two main categories: first, terms that aremultiplied by the variation of the state vector, �w; second, terms thatare multiplied by the variation of the shape function, �S. The time-accurate discrete adjoint equation can now be defined as

@ n�1i;j

@�� Vi;j

�3

2�t T

n�1i;j �

2

�t T

n�2i;j �

1

2�t T

n�3i;j

� Tn�1i;j �wRn�1i;j � 0 (22)

The first term, @ n�1i;j =@�, is added tomarch the solution in a fictitious

time � to a periodic steady-state solution. The second term representsa second-order forward-difference formula for the temporaldiscretization of the time-accurate adjoint equation. The termdemonstrates that to acquire the solution of the adjoint equation at thetime step n� 1 requires the solution at n� 2 and n� 3. Therefore,the adjoint equation is integrated backward in time from a terminalboundary condition. The last term requires the variation of theresidual as a function of the state vector. Here, Rn�1i;j denotes the

discrete representation of the residual which is the sumof the discreteconvective and dissipativefluxes in each cell. A full differentiation ofthe equation would involve differentiating every term that is afunction ofw. A collection of all terms that are multiplied to �wwillproduce the discrete adjoint convective and dissipative fluxes. Referto Nadarajah [21] for a complete derivation of the discrete adjointfluxes.

To acquire the discrete adjoint boundary condition, Eq. (22) can beexpanded at cell i; 2 to create the following:

@ n�1i;2

@�� Vi;2

�3

2�t T

n�1i;2 �

2

�t T

n�2i;2 �

1

2�t T

n�3i;2

� 1

2

hAT

n�1

i�12;2

� n�1i;2 � n�1i�1;2

�� ATn�1

i�12;2

� n�1i�1;2 � n�1i;2

� BTn�1i;52

� n�1i;3 � n�1i;2

���

i� 0 (23)

whereA andB are the Euler Jacobians in the � and directions, and�is the source term for drag minimization,

���x2�i n�12i;2��x1�i

n�13i;2���x2�i! cos�

��x1�i! sin�� @p@w

�t

All of the terms in Eq. (23), except for the source term, scale as thesquare of�xi. Therefore, as the mesh width is reduced, the terms inthe source term must approach zero as the solution reaches a steadystate. One then recovers the continuous adjoint boundary condition,as stated in Eq. (16). Equation (21) then simplifies to the followingform:

�I � � 1

T

��1

2�pi;2 � pi;1� � p1

�����x2�i�! cos�

� ���x1�i�! sin���t� Tn�1i;j �fR

n�1i;j

��

The preceding equation represents the total gradient obtained usingthe time-accurate discrete adjoint approach to reduce the total drag ofa pitching airfoil, and �fR

n�1i;j represents the variation of the residual

as a function of the shape function.

VI. Design Process

The objective of this work is to change the shape of the airfoil tominimize its time-averaged coefficient of drag. Given the derivationprovided in previous sections, the adjoint boundary condition caneasily be modified to admit other figures of merit. The shape of theairfoil is constrained such that themaximum thickness-to-chord ratioremains constant between the initial and final designs. In addition,the mean angle of attack �o is allowed to vary to ensure that the time-averaged coefficient of lift remains constant between designs. Theairfoil undergoes a forced pitching oscillation about the quarter-chord. The angle of incidence is given by

��t� � �o � �m sin�!t�

where �o � 0 deg, and the maximum angle of attack is�m � 1:01 deg. One period of oscillation is defined from t� 0 tot� 2. To compute the entire unsteady flow solution for eachperiod, ��t� is divided into N discrete points or time steps. Theindividual stepswithin each iteration of the design process for the fullunsteady design and multipoint approaches are outlined next.

A. Full Unsteady Design

The following steps detail the procedures for the full unsteadycontinuous and discrete adjoint-based design optimization problem:

1. Unsteady Flow Calculation

A multigrid scheme is used to drive the unsteady residual at eachtime step to a negligible value. The duration of the physical timehistory (typically quantified in the number of oscillatory periods)depends on the physics of the flowfield and the accuracyrequirements of the calculation. Generally, it requires five periodsbefore a limit cycle is achieved. Here, a period refers to one fulloscillation. During the last period, the flow solution at each time stepis saved in memory. Fifteen multigrid cycles were used for each timestep. If 24 time steps are used for each cycle and five cycles are usedto achieve the limit cycle, then a total of 1800 multigrid cycles arerequired to obtain the unsteady solution. To maintain the time-averaged lift coefficient, the mean angle of attack �o is perturbed.However,�o is onlymodified every three periods, because it requiresat least three periods for the global coefficients such as time-averagedlift and drag to converge to an accuracy level of 1E-3. A total of 15periods are needed instead of five to achieve the desired time-averaged lift coefficient. This multiplies the total cost by three times.

2. Unsteady Adjoint Calculation

The unsteady adjoint equation, either the discrete or continuousversion, requires integration in reverse time. The same numerical

1482 NADARAJAH AND JAMESON

Page 6: Optimum Shape Design for Unsteady Flows with Time …aero-comlab.stanford.edu/Papers/AIAA-24332-153.pdfAerodynamic shape optimization of rotor airfoils in an unsteady viscous flow

scheme employed to solve the unsteady flow is used here, as wellwith minor adjustments in the code to allow integration in reversetime.Only three periods are needed before the limit cycle is achieved.Fifteenmultigrid cycles are used at each time step,which translates toa total of 1080 cycles to achieve a limit cycle for the adjoint equation.

3. Gradient Evaluation

An integral over the last period of the adjoint solution is used toform the gradient. This gradient is then smoothed using an implicitsmoothing technique. This ensures that each new shape in theoptimization sequence remains smooth and acts as a preconditioner,which allows the use of much larger steps. The smoothing leads to alarge reduction in the number of design iterations needed forconvergence. Refer to Nadarajah and Jameson [22] for a morecomprehensive overview of the gradient smoothing technique. Anassessment of alternative search methods for a model problem isgiven by Jameson and Vassberg [23].

4. Airfoil-Shape Modification

The airfoil shape is then modified in the direction of improvementusing a steepest-descent method.

Let F represent the design variable and G the gradient. Animprovement can then be made with a shape change

�F ���G

where � is determined by trial and error to ensure a fast rate ofconvergence while guaranteeing that the objective functionconvergence to an error level of at least 1E-6. The step-size valueis constant for all cases presented in this work.

5. Grid Modification

The internal grid is modified based on perturbations on the surfaceof the airfoil. The method modifies the grid points along each gridindex line projecting from the surface. The arc length between thesurface point and the far-field point along the grid line is firstcomputed, then the grid point at each location along the grid line isattenuated proportional to the ratio of its arc length distance from thesurface point and the total arc length between the surface and the farfield.

6. Repeat Design Process

The entire design process is repeated until the objective functionconverges. The problems in this work typically required between 9and 25 design cycles. Each design cycle required 1800 multigridcycles to compute the flow solution and 1080 cycles for the adjointsolution.

B. Multipoint Design

In the multipoint design approach, the unsteady flow and adjointsolvers are replacedwith a steady flow and adjoint solver at each timestep. The gradient is an average of the gradients from each time step,defined as

�g� 1

N

XNi�1

!tigi

where gi is the gradient and !ti is the weight at each design point. Inthis work, the weights are chosen to be unity. If the time-averaged liftis not constrained, then only one period is required for both the flow

and adjoint solvers, and the total computational cost is 720 multigridcycles.

Table 1 illustrates a cost comparison between the unsteady andmultipoint design approaches. Here, themiddle two columns containthe total number of multigrid cycles used to compute the Euler andadjoint equations. The numbers in the last column signify the ratio ofcost of the unsteady design method with respect to the multipointapproach. Using the full unsteady design approach requires fourtimes the computational cost as doing the multipoint approach.However, if the time-averaged lift is constrained, an additional fourperiods are required for the multipoint approach to converge thetime-averaged lift coefficient and an additional ten are needed for theunsteady case. This brings the total number of multigrid cyclesrequired by the flow solver per design cycle to 1800 for themultipoint case and 5400 for the unsteady. The cost of the adjointsolver remains the same, because the mean angle of attack is onlymodified during the flow-solver stage. The unsteady approach is thenthree times the computational cost of the multipoint approach.

The difference in cost between one steady Runge–Kutta iterationand one unsteady Runge–Kutta iteration was not factored into thecomputing cost, because the difference isminimal, requiring only theaddition of the time derivatives of the flow variables for the implicittime-stepping algorithm.

VII. Results

The following subsections present results of the time-averageddragminimization problem for a two-dimensional airfoil undergoinga periodic pitching motion. The first subsection contains a code-validation study. The second subsection is dedicated to a temporal-resolution study. The third subsection demonstrates the redesign ofthe RAE 2822 airfoil to reduce the time-averaged drag coefficientwhile maintaining the time-averaged lift. The last subsectionpresents a comparison between the multipoint and unsteady designapproaches.

A. Code Validation

An Euler solution is computed on a 193 � 33 grid and the liftcoefficient versus angle of attack is compared with the experimentalNACA 64A010 CT6 [24] data. A spatial-resolution study [21] wasconducted for grids of various types and sizes to determine theminimum number of grid points needed to establish an accuratehysteresis loop. It was deemed that a 193 � 33 grid, as shown inFig. 1, was adequate for inviscid solutions.

Fig. 1 NACA 64A010 193 � 33 mesh.

Table 1 Comparison of computational cost between the design

approaches

Method Euler (multigrid cycles) Adjoint (multigrid cycles) Cost

Multipoint 360 360 1Full unsteady 1800 1080 4

NADARAJAH AND JAMESON 1483

Page 7: Optimum Shape Design for Unsteady Flows with Time …aero-comlab.stanford.edu/Papers/AIAA-24332-153.pdfAerodynamic shape optimization of rotor airfoils in an unsteady viscous flow

The computations are performed at a freestream Mach number of0.78, a mean angle of attack of �o � 0 deg, a maximum angle ofattack of �m � 1:01 deg, and at a reduced frequency of !r � 0:202.Five periods of computation are required in order for the periodicflow to be established and to allow the time-averaged lift and dragcoefficients to converge to an error level of 1E-3. Figure 2 illustratesthe hysteresis loop and the results reproduce the experimental resultswith sufficient accuracy. In Fig. 2, the lift coefficient versus angle ofattack loop moves in a counterclockwise direction. The nonlinearbehavior due to the movement of the shock causes an amplitudereduction of the lift coefficient and a phase lag.

The primary goal of this work is to investigate the benefits ofaerodynamic shape optimization of airfoils undergoing unsteadymotion via an adjoint-based unsteady optimization approach. Thetraditional approach is to perform amultipoint design using solutionsfrom various phases of the unsteady motion. The followingsubsections will establish that under specific conditions, there is anadvantage to an unsteady optimization approach. First, a temporal-resolution study will be examined. Second, an optimization of apitching airfoil to reduce drag will be demonstrated. Finally, athorough investigation of the advantage of an unsteady optimizationapproach over a multipoint will be presented.

The chosen test case is that of a pitching RAE 2822 airfoil intransonic flow at a Mach number of 0.78, a mean angle of attack of�o � 0 deg, and a reduced frequency of !r � 0:20 on a 193 � 33grid. Figure 3 illustrates the convergence history for the unsteadyflow solver and the unsteady continuous and discrete adjoint

equations. The continuous and discrete unsteady adjoint equationshave the same convergence rate, and the residuals reach machinezero within 100 multigrid W-cycles.

For unsteady transonic flows, as the airfoil oscillates at a smallangle of attack, the shock wave moves back and forth about a meanlocation and is closely sinusoidal and lags the airfoil motion. This lagis evident in the lift hysteresis loop, in which the maximum lift doesnot occur at the maximum angle of attack. The nonlinear behavior ofinviscid unsteady transonic flows is primarily due to the movementof the shock. According to McCroskey [25], however, the shockwave motion is greatest at moderate reduced frequencies. He furtherclarifies that for small-amplitude oscillations for reduced frequenciesapproximately below 0.5, large shock motion is seen. This motiongradually reduces as the reduced frequency increases. It is alsoexpected that at very low reduced frequencies, the flowcharacteristics are very similar to that of steady-state computations.Lift hysteresis loops were computed for several different reducedfrequencies at the same Mach number and are compared with thesteady lift curve in Fig. 4. As expected, at a reduced frequency of0.01, the lift-curve slope is similar to the steady solution and the slopereduces as the reduced frequency increases.

B. Temporal Resolution

To quantify the required number of time steps per period, atemporal convergence study was performed. The code was run with5, 11, 23, 47, 95, and 191 time steps per period. At each time step, themodified residual R�wi;j� from Eq. (8) was driven to machine zero.

−1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 1−0.1

−0.08

−0.06

−0.04

−0.02

0

0.02

0.04

0.06

0.08

0.1

Angle of Attack

Lift

Coe

ffici

ent

Time Accurate, 23 Time Steps

Experimental

Fig. 2 Comparison of lift hysteresis with NACA 64A010 CT6 [24]experimental data at M1 � 0:78 and !r � 0:202.

0 20 40 60 80 100 120 140 160 180 200

10−14

10−12

10−10

10−8

10−6

10−4

10−2

100

Multigrid Cycles

Log(

Err

or)

Unsteady Flow SolverUnsteady Continuous AdjointUnsteady Discrete Adjoint

Fig. 3 Convergence history of the unsteady flow solver and unsteady

continuous and discrete adjoint equations; 193 � 33 mesh; RAE 2822

airfoil atM1 � 0:78.

−1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 1

0.3

0.35

0.4

0.45

0.5

0.55

0.6

0.65

0.7

0.75

Angle of Attack

Lift

Coe

ffici

ent,

c l

Steady

ωr = 0.01

ωr = 0.05

ωr = 0.20

ωr = 0.45

Fig. 4 Lift hysteresis at various reduced frequencies for the RAE 2822

atM1 � 0:78.

5 10 15 20 25 30 35 40 45 5010

−6

10−5

10−4

10−3

Number of Time Steps Per Period

cl e

rror

Fig. 5 Clerroras a function of the number of time steps per period using a

logarithmic vertical scale.

1484 NADARAJAH AND JAMESON

Page 8: Optimum Shape Design for Unsteady Flows with Time …aero-comlab.stanford.edu/Papers/AIAA-24332-153.pdfAerodynamic shape optimization of rotor airfoils in an unsteady viscous flow

To guarantee temporal convergence, both the time-averaged lift anddrag coefficients were monitored. Generally, 60 time periods wereneeded to eliminate errors due to initial transients and to ensure amachine-zero convergence of the time-averaged lift and drag.Figure 5 illustrates theClerror as a function of the temporal resolution.The case with 191 time steps per period was used as the controlsolution, and each point on the figure represents the differencebetween the time-averaged lift coefficient of the control solution andthe current solution. Figure 5 illustrates that the error decay rate isproportional to �t2. Table 2 shows that 23 time steps per period issufficient to obtain a lift coefficient accurate to four decimal places.

To perform optimum shape design for unsteady flows, it isimportant to access the required temporal resolution to obtainaccurate sensitivity of the objective function to the design variable. Inthis work, the objective function is the time-averaged dragcoefficient, and the design variables are the points on the surface ofthe airfoil. Figure 6 illustrates the L2-norm of the gradient error as afunction of temporal resolution for both the continuous and discreteadjoint approaches, using the gradient computed with 191 time stepsas the control solution. The gradients of the two adjoint approachesconverge at the same rate. To ensure accurate gradients, the residualsat each time step for both the flow and adjoint solvers were driven tomachine zero, and 60 time periods were used for both solvers toensure elimination of transient errors. The decay rate of the L2-normof the gradient error is proportional to �t2. The equivalentconvergence decay rates between the gradient (Fig. 6) and time-averaged lift (Fig. 5) confirm the proper implementation of both thecontinuous and discrete adjoint equations. Figure 7 shows thegradient at each point on the surface of the airfoil. The figureillustrates that the gradient is almost identical to engineeringaccuracy for a vast range of points, except in the region of the shockwave between grid points 80 and 100. A closer inspection of the datareveals that these points converge as the temporal resolution isincreased. This is further illustrated in Table 3, which shows theconvergence of the L2-norm of the gradient error. The gradient errorfor the case with 23 time steps per period is approximately 1e-04,providing an equivalent level of accuracy to that observed for thetime-averaged lift. Thus, 23 time steps per period is deemedsufficient for this particular test case.

Figure 8 shows a comparison of the discrete and continuousadjoint gradients for the case with 23 time steps per period. A closer

examination of the discrete and continuous adjoint solutions revealsthat the excellent gradient comparison is primarily due to thecomparable adjoint solutions between the two approaches. Acomparison of the second adjoint variable along the wall isdemonstrated in Fig. 9. In Fig. 10, both the second and third adjointvariables are shown. Values are plotted for points along a grid lineextending from the airfoil surface to the far-field boundary. The twomethods differ in the manner that the adjoint convective anddissipative fluxes are discretized, as well as in the enforcement of theboundary condition. In the case of the continuous adjoint approach,the boundary condition is computed for the cells below thewall alongj� 1; however, in the case of the discrete adjoint method, theboundary condition appears as source terms added to the adjointconvective and dissipative fluxes along the cells above the wall(j� 2). Therefore, the discrete adjoint approach does not requirevalues of adjoint variables along j� 1; however, they are set to beequal to the values in j� 2. The effect of this difference in theenforcement of the boundary condition can be observed in a close-upview of the adjoint values close to the boundary, as illustrated inFig. 10. Despite the difference in the boundary conditions and theslight difference in the adjoint solution, the two approaches producevirtually indistinguishable gradients, as demonstrated in Fig. 8.

C. Time-Averaged Drag Minimization of RAE 2822 Airfoil

with Fixed Time-Averaged Lift Coefficient

Figure 11 illustrates the initial and final geometry for the RAE2822 airfoil. The solid line represents the initial airfoil geometry andthe dashed line illustrates the redesigned airfoil. A distinctive featureof the new airfoil is in the drastic reduction of the upper-surfacecurvature. A reduced curvature leads to a weaker shock and thus alower wave drag, but it also leads to a reduction in airfoil camber,resulting in a loss in lift.

To maintain the time-averaged lift coefficient Cl, the mean angleof attack �o is perturbed to a new value. The impact of this decisionresulted in a need to compute more periods to allow theCl andCd toconverge. In this work, �o was perturbed every three periods. Thisallowed the Cl to converge to a new value before the angle of attack

Table 2 Time-averaged coefficient of lift as a function of temporal

resolution

Time stepsper period

5 11 23 47 95 191

Cl 0.5084 0.5086 0.5087 0.5087 0.5087 0.5087

5 10 15 20 25 30 35 40 45 5010

−5

10−4

10−3

10−2

Number of Samples Per Period

||err

or|| 2

DiscreteContinuous

Fig. 6 Gradient errorL2-normas a function of the number of time steps

per period using a logarithmic vertical scale.

20 40 60 80 100 120

−0.01

−0.005

0

0.005

0.01

Grid Point

Gra

dien

t

5 SPP

11 SPP

23 SPP

47 SPP

95 SPP

Fig. 7 Comparison of discrete unsteady adjoint gradients for various

temporal resolutions.

Table 3 L2-norm of the gradient error as a function of temporal

resolution

Time steps per period L2-norm

5 1.8686e-0311 6.1302e-0423 1.4652e-0447 2.8524e-05

NADARAJAH AND JAMESON 1485

Page 9: Optimum Shape Design for Unsteady Flows with Time …aero-comlab.stanford.edu/Papers/AIAA-24332-153.pdfAerodynamic shape optimization of rotor airfoils in an unsteady viscous flow

was perturbed any further. A total of 15 periods were used for eachdesign cycle. Figure 12 illustrates the convergence rate ofCd, whichis reduced by 43% from 132 drag counts to 75 drag counts within 21design cycles. In Fig. 12, the change in the objective function or�I,where I is the objective function, is also shown. During the first 16design cycles, �I converges linearly, as expected. Linearconvergence is characteristic of a steepest-descent-type method.As the final airfoil profile is realized, the convergence �I increasesrapidly. The code is automatically stopped as soon as a change of 1E-6 is detected. This level of change corresponds to a change to thesixth decimal place of the drag coefficient and this is sufficient forengineering accuracy.

Figure 13 illustrates the initial and final pressure contours at a 180-deg phase. The sonic line represented by a dashed line is overplottedon each figure. It is clearly visible that the strong shock on the uppersurface of the initial geometry was considerably reduced.Figures 14a–14d illustrate the upper and lower surface instantaneouspressure coefficients for the initial and final design. In Fig. 14a, acomparison of the initial instantaneous pressure distribution versusthe final at a 0-deg phase shows an almost complete reduction of thewave drag. The strong shock on the suction side of the airfoil isweakened at all other phases of the oscillation.

D. Multipoint Versus Unsteady Optimization

A multipoint design approach was often the method of choice foroptimization of airfoils in an unsteady flow environment, due to its

20 40 60 80 100 120

−0.01

−0.005

0

0.005

0.01

Grid Point

Gra

dien

t

Discrete

Continuous

Fig. 8 Comparison of discrete and continuous unsteady adjoint

gradients for 23 time steps per period.

10 20 30 40 50 60

−0.3

−0.25

−0.2

−0.15

−0.1

−0.05

Grid Point

Sec

ond

Adj

oint

Var

iabl

e

DiscreteContinuous

Fig. 9 Comparison of the second adjoint variable at the wall for thediscrete and continuous unsteady adjoint methods for 23 time steps per

period.

−0.03 −0.02 −0.01 0 0.010

5

10

15

20

25

30

35

Second Adjoint Variable

Grid

Num

ber

in th

e j D

irect

ion

−0.01 0 0.01 0.020

5

10

15

20

25

30

35

Third Adjoint Variable

Discrete

Continuous

0.01 0.0120

1

2

3

4

5

6−0.03 −0.026

0

1

2

3

4

Fig. 10 Comparison of the second and third adjoint variables for the

discrete and continuous unsteady adjoint methods for 23 time steps per

period.

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

−0.06

−0.04

−0.02

0

0.02

0.04

0.06

Initial

Final

Fig. 11 Initial and final geometries for the RAE 2822 airfoil at

M1 � 0:78, !r � 0:20, and �o � 0deg.

0.008

0.009

0.010

0.011

0.012

0.013

0.014

Tim

e−A

vera

ged

Dra

g

0 2 4 6 8 10 12 14 16 18 20 221E−6

1E−5

1E−4

1E−3

∆ I

Time−Averaged Drag∆ I

Fig. 12 Convergence of the time-averaged drag coefficient for the RAE

2822 airfoil atM1 � 0:78, !r � 0:20, and �o � 0deg.

1486 NADARAJAH AND JAMESON

Page 10: Optimum Shape Design for Unsteady Flows with Time …aero-comlab.stanford.edu/Papers/AIAA-24332-153.pdfAerodynamic shape optimization of rotor airfoils in an unsteady viscous flow

lower computational and memory cost. In this subsection of thepaper, we make the argument that even if a multipoint designapproach is cheaper, it cannot replace a full unsteady optimizationtechnique. Through the use of final airfoil profiles, convergence

histories, and gradient comparisons, the following results show thatthere are benefits to unsteady optimization.

To compare the unsteady optimization approach to that of themultipoint technique, unsteady design cases are computed at reduced

Fig. 13 Pressure contour plot for the RAE 2822 airfoil at grid of 192 � 32, M1 � 0:78, !r � 0:20, and fixed Cl � 0:51.

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

−2

−1.8

−1.6

−1.4

−1.2

−1

−0.8

Pre

ssur

e C

oeffi

cien

t

X Coordinate

a) Phase = 0 deg

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

−2

−1.8

−1.6

−1.4

−1.2

−1

−0.8

Pre

ssur

e C

oeffi

cien

t

X Coordinate

b) Phase = 90 deg

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

−2

−1.8

−1.6

−1.4

−1.2

−1

−0.8

Pre

ssur

e C

oeffi

cien

t

X Coordinate

c) Phase = 180 deg

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

−2

−1.8

−1.6

−1.4

−1.2

−1

−0.8

Pre

ssur

e C

oeffi

cien

t

X Coordinate

d) Phase = 270 deg

Fig. 14 Initial and final pressure coefficients at various phases for the RAE 2822 airfoil atM1 � 0:78,!r � 0:20, and �o � 0deg (◇ is initial pressure

and □ is final pressure).

NADARAJAH AND JAMESON 1487

Page 11: Optimum Shape Design for Unsteady Flows with Time …aero-comlab.stanford.edu/Papers/AIAA-24332-153.pdfAerodynamic shape optimization of rotor airfoils in an unsteady viscous flow

frequencies ranging from 0.01 to 0.45 at a Mach number ofM1 � 0:78. The time-averaged lift coefficient for all reducedfrequencies is fixed atCl � 0:51. Lift hysteresis loops for the variousreduced frequencies are compared with the steady lift curve in Fig. 4.

Figure 15 illustrates a comparison of final airfoil geometriesbetween airfoils designed using the full unsteady optimizationapproach at various reduced frequencies and the multipointtechnique. The airfoils are designed at a Mach number ofM1 � 0:78, a mean angle of attack of �o � 0 deg, and an angle ofattack deviation of �1:01 deg. We show in Fig. 15 that the airfoildesigned using the multipoint approach, using five time steps perperiod, designated as mp 5 is almost identical to the one designedusing a full unsteady optimization approach at a reduced frequencyof !r � 0:01. Figure 16 demonstrates a close-up view of the finalupper surface between the 10 and 75% chord location for variousdesign approaches. Thefinal airfoil based on themultipoint approachusing various number of time steps per period virtually producesidentical profiles. As expected and seen in the previous figure, theunsteady design at a reduced frequency of 0.01 is very similar to thatof the multipoint approaches. As the reduced frequency increases,the final airfoil profile departs from that produced by the simplemultipoint approach.However, the differences are very small, exceptin areas on the upper surface, for which a greater reduction in thecurvature is seen for higher reduced frequencies.

Figure 17 further supports this fact, with almost identicalconvergence histories of the objective function (time-averaged drag)between the various multipoint approaches and the unsteady designat !r � 0:01. Table 4 contains a comparison of the initial and finaltime-averaged drag counts for the various design approachesperformed at various reduced frequencies. The table furtherillustrates that for the four cases (mp 5, mp 11, mp 23, and!r � 0:01), the initial time-averaged drag is approximately 147 dragcounts and the final drag is 111 counts. As the reduced frequencyincreases, the initial drag reduces by a small amount due to nonlineareffects primarily due to the movement of the shock; however, thefinal drag count reduces substantially. Therefore, the three unsteadycases showcased in Table 4 and Figs. 15 and 17 represent the lower,middle, and upper limits of the reduced-frequency range. Figure 18demonstrates the change in�I, where I is the objective function and�I converges linearly until the final airfoil profile is realized and theconvergence accelerates. All design cases are automatically haltedonce �I reaches 1E-6. As illustrated in both Figs. 16 and 17, theconvergence of the unsteady case at !r � 0:01 is very similar to themultipoint cases. As the reduced frequency increases, the number ofdesign cycles required to obtain the final design increases. Weconjecture that this is due to the level of unsteadiness of the problem.

In Figs. 19 and 20,we demonstrate the comparison of the gradientsbetween the adjoint-based multipoint and unsteady approaches. Thegradients are plotted in a clockwise direction from the lower trailingedge to the upper trailing edge. Figure 19 illustrates that the gradientsbetween the multipoint approach using 23 time steps per period arevery similar to the unsteady design approach for various reducedfrequencies for a majority of the grid points, except for grid pointsbetween 80 and 100, as shown in Fig. 20. This range of grid pointscoincides with the shock footprint and produces the dominantgradients in this design problem. Thefigure illustrates once again thatthe gradients produced by the unsteady test case at a reducedfrequency of!r � 0:01 are similar to that produced by themultipointapproach. At higher reduced frequencies, the gradients differ greatlyin this range of grid points and this is largely due to the increase in thenonlinear behavior of the unsteady transonic flow.

In summary, Figs. 15–20 demonstrate that at very low reducedfrequencies, the design convergence histories, final airfoil profiles,

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

−0.06

−0.04

−0.02

0

0.02

0.04

ωr = .010

ωr = .050

ωr = .202

ωr = .450

mp 5

Fig. 15 Comparison of final airfoil geometries between variousreduced frequencies and the multipoint approach at M1 � 0:78,��� 0deg, and fixed Cl � 0:51.

0.2 0.3 0.4 0.5 0.6 0.70.04

0.042

0.044

0.046

0.048

0.05

0.052

0.054

0.056

X Coordinate

Y C

oord

inat

e

mp 5

mp 11

mp 23

ωr = .010

ωr = .050

ωr = .202

ωr = .450

Fig. 16 Close-up view of the final airfoil geometry at M1 � 0:78,��� 0deg, and fixed Cl � 0:51.

2 4 6 8 10 12 14 16 18 20 22

0.008

0.009

0.01

0.011

0.012

0.013

0.014

Design Cycles

Tim

e−Av

erag

ed D

rag

Coe

ffici

ent,

C d

mp 5

mp 11

mp 23

ωc = 0.010

ωc = 0.050

ωc = 0.202

ωc = 0.450

Fig. 17 Convergence history of the time-averaged drag coefficient for

various design approaches for the RAE 2822 airfoil at M1 � 0:78,�o � 0deg, and fixed Cl � 0:51.

Table 4 Initial and final time-averaged drag for various design

approaches

Case Initial Cd Final Cd Reduction

Multipoint, mp 5 0.0147 0.0111 24.5%Multipoint, mp 11 0.0147 0.0111 24.5%Multipoint, mp 23 0.0147 0.0112 23.8%Full unsteady, !r � 0:01 0.0146 0.0112 23.3%Full unsteady, !r � 0:05 0.0138 0.0092 33.3%Full unsteady, !r � 0:20 0.0132 0.0075 43.2%Full unsteady, !r � 0:45 0.0132 0.0072 45.5%

1488 NADARAJAH AND JAMESON

Page 12: Optimum Shape Design for Unsteady Flows with Time …aero-comlab.stanford.edu/Papers/AIAA-24332-153.pdfAerodynamic shape optimization of rotor airfoils in an unsteady viscous flow

and gradients are very similar to that produced by amultipoint designapproach. At moderate reduced frequencies between 0.05 and 0.45,however, these trends begin to deviate. The greater the nonlinearbehavior of the unsteady transonic flow, the larger the differencefrom the multipoint approach. Figure 21 further supports thisobservation by illustrating the additional number of time periodsrequired for the higher reduced frequency cases to obtain aconverged time-averaged drag coefficient to machine accuracy. Theadditional periods are required to diminish the transient solutions dueto the increase in the nonlinear behavior, to arrive at a periodicsteady-state solution. The figure demonstrates that only four periodsare required to obtain a converged drag coefficient at a reducedfrequency of 0.01; however, 60 periods were required for the 0.20case and 100 for the 0.45 case.

Ultimately, to compare the behavior of thefinal airfoil designed bythe two approaches, the airfoil designed using the multipointapproach was computed using the time-accurate flow solver for thethree moderate reduced frequencies. Table 5 lists the final time-averaged drag coefficients computed at aMach number of 0.78 and ata fixed lift coefficient of 0.51. At !r � 0:05, the difference betweenthe unsteady and multipoint approaches is 0.2%. At a reducedfrequency of 0.20 and 0.45, the difference grows to 6.4 and 8.3%,resulting in lower drag coefficients for the airfoils designed using theunsteady approach. Figure 22 illustrates a comparison of the pressuredistribution at a reduced frequency of 0.45 between the airfoilsdesigned using the multipoint and unsteady approaches. It is clearlyseen that the 8.3% improvement in the drag coefficient for theunsteady technique is due to theweaker shock on the upper surface ofthe airfoil.

The multipoint approach certainly provides an inexpensivealternative to the unsteady design technique and has produced anoptimized airfoil with a lower time-averaged drag coefficient over arange of reduced frequencies. However, if an airfoil, turbine, or rotorblade is designed for a specific range of reduced frequencies, theadjoint-based unsteady optimization technique may provide anadded benefit.

VII. Conclusions

This paper presents a complete formulation of the continuous anddiscrete unsteady inviscid adjoint approaches to automatic

0 2 4 6 8 10 12 14 16 18 20 2210

−6

10−5

10−4

10−3

Design Cycles

∆ I

mp 5

mp 11

mp 23

ωc = 0.010

ωc = 0.050

ωc = 0.202

ωc = 0.450

Fig. 18 Convergence of �I for various design approaches for the RAE2822 airfoil at M1 � 0:78, �o � 0deg, and fixed Cl � 0:51.

20 40 60 80 100 120

−0.01

−0.005

0

0.005

0.01

Grid Point

Gra

dien

t

mp 23

ωr = 0.010

ωr = 0.050

ωr = 0.202

ωr = 0.450

Fig. 19 Comparison between unsteady discrete adjoint gradients for

various reduced frequencies and the multipoint approach for the RAE

2822 airfoil at M1 � 0:78, �o � 0deg, and fixed Cl � 0:51.

80 82 84 86 88 90 92 94 96 98 100

−0.01

−0.005

0

0.005

0.01

Grid Point

Gra

dien

t

mp 5

mp 11

mp 23

ωr = 0.010

ωr = 0.050

ωr = 0.202

ωr = 0.450

Fig. 20 Close-up view of the comparison between unsteady discreteadjoint gradients for various reduced frequencies and the multipoint

approach for the RAE 2822 airfoil atM1 � 0:78, �o � 0deg, and fixed

Cl � 0:51.

0 10 20 30 40 50 60 70 80 90 100

10−12

10−10

10−8

10−6

10−4

10−2

100

Number of Time Periods

Err

or

ωr = 0.010

ωr = 0.050

ωr = 0.202

ωr = 0.450

Fig. 21 Convergence of time-averaged lift coefficient for various

reduced frequencies.

Table 5 Comparison ofCd between themultipoint andunsteady designapproaches

Case Unsteady Multipoint Improvement

!r � 0:05 0.00925 0.00927 0.2%!r � 0:20 0.00750 0.00801 6.4%!r � 0:45 0.00719 0.00784 8.3%

NADARAJAH AND JAMESON 1489

Page 13: Optimum Shape Design for Unsteady Flows with Time …aero-comlab.stanford.edu/Papers/AIAA-24332-153.pdfAerodynamic shape optimization of rotor airfoils in an unsteady viscous flow

aerodynamic design. A 46% reduction in the time-averaged dragcoefficient was achieved for the RAE 2822 airfoil at a reducedfrequency of 0.45 while maintaining the time-averaged liftcoefficient. A comparison between final airfoil profiles, convergencehistories of the objective function, and gradients demonstrate that atlow reduced frequencies, there is no added benefit of performingaerodynamic shape optimization for unsteady flows via an adjoint-based unsteady optimization technique; however, at moderatereduced frequencies, an unsteady optimization technique producedfinal airfoil profileswith time-averaged drag coefficients between 6.4to 8.3% improvement over the multipoint approach. The frameworkwas established to extend this method to viscous dominated flows inwhich secondary flow effects are present.

Acknowledgments

This research benefited greatly from the generous support of theU.S. Air Force Office of Scientific Research (AFOSR) under grantnumber AF F49620-98-1-022 and the Department of Energy undercontract number LLNLB341491 as part of the Accelerated StrategicComputing Initiative (ASCI) program.

References

[1] Jameson, A., “Computational Aerodynamics for Aircraft Design,”Science, Vol. 245, No. 4916, July 1989, pp. 361–371.

[2] Reuther, J., Cliff, S., Hicks, R., and van Dam, C. P., “Practical Design

Optimization of Wing/Body Configurations Using the EulerEquations,” AIAA Paper 92-2633, 1992.

[3] Gallman, J., Reuther, J., Pfeiffer, N., Forrest, W., and Bernstorf, D.,“Business Jet Wing Design Using Aerodynamic Shape Optimization,”34th Aerospace Sciences Meeting and Exhibit, Reno, NV, AIAAPaper 96-0554, 1996.

[4] Reuther, J., Alonso, J. J., Martins, J. R. R. A., and Smith, S. C., “ACoupled Aero-Structural Optimization Method for Complete AircraftConfigurations,” 37th Aerospace Sciences Meeting and Exhibit, Reno,NV, AIAA Paper 99-0187, 1999.

[5] Kasidit, L., and Jameson, A., “Case Studies in Aero-Structural WingPlanform and Section Optimization,” 22nd AIAA AppliedAerodynamics Conference, Providence, RI, AIAA Paper 2004-5372,2004.

[6] Nadarajah, S., Jameson, A., and Alonso, J. J., “An Adjoint Method forthe Calculation of Remote Sensitivities in Supersonic Flow,” 40thAerospace Sciences Meeting and Exhibit, Reno, NV, AIAAPaper 2002-0261, 2002.

[7] Mavriplis, D., “Multigrid Solution of the Discrete Adjoint forOptimization Problems on Unstructured Meshes,” AIAA Journal,Vol. 44, Jan. 2006, pp. 42–50.

[8] Soto, O., Lohner, R., and Yang, C., “An Adjoint-Based DesignMethodology for CFD Problems,” International Journal of NumericalMethods for Heat and Fluid Flow, Vol. 14, No. 6, 2004, pp. 734–759.

[9] Giles, M. B., Duta, M., Muller, J., and Pierce, N., “AlgorithmDevelopments for Discrete Adjoint Methods,” AIAA Journal, Vol. 15,No. 5, 2003, pp. 1131–1145.

[10] Nemec, M., Zingg, D., and Pulliam, T. H., “Multipoint and Multi-Objective Aerodynamic Shape Optimization,” AIAA Journal, Vol. 42,No. 6, 2004, pp. 1057–1065.

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

−2

−1.8

−1.6

−1.4

−1.2

−1

−0.8

Pre

ssur

e C

oeffi

cien

t

X Coordinate

a) Phase = 0 deg

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

−2

−1.8

−1.6

−1.4

−1.2

−1

−0.8

Pre

ssur

e C

oeffi

cien

t

X Coordinate

b) Phase = 90 deg

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

−2

−1.8

−1.6

−1.4

−1.2

−1

−0.8

Pre

ssur

e C

oeffi

cien

t

X Coordinate

c) Phase = 180 deg

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

−2

−1.8

−1.6

−1.4

−1.2

−1

−0.8

Pre

ssur

e C

oeffi

cien

t

X Coordinate

d) Phase = 270 deg

Fig. 22 Comparison of pressure distributions between airfoils designed via multipoint and unsteady approaches at various phases for the RAE 2822airfoil atM1 � 0:78, !r � 0:45, and �o � 0deg (◇ is unsteady and□ is multipoint).

1490 NADARAJAH AND JAMESON

Page 14: Optimum Shape Design for Unsteady Flows with Time …aero-comlab.stanford.edu/Papers/AIAA-24332-153.pdfAerodynamic shape optimization of rotor airfoils in an unsteady viscous flow

[11] Mohammadi, B., and Pironneau, O., “Shape Optimization in FluidMechanics,” Annual Review of Fluid Mechanics, Vol. 36, 2005,pp. 255–279.

[12] Kim, H., andNakahashi, K., “UnstructuredAdjointMethod for Navier-Stokes Equations,” JSME International Journal, Series B (Fluids and

Thermal Engineering), Vol. 48, No. 2, May 2005, pp. 202–207.[13] Nielsen, E., Lu, J., Park, M. A., and Darmofal, D., “An Implicit, Exact

Dual Adjoint Solution Method for Turbulent Flows on UnstructuredGrids,” Computers and Fluids, Vol. 33, No. 9, 2004, pp. 1131–1155.

[14] Elliot, J., and Peraire, J., “Aerodynamic Design Using UnstructuredMeshes,” AIAA Paper 96-1941, 1996.

[15] Anderson, W. K., and Venkatakrishnan, V., “Aerodynamic DesignOptimization on Unstructured Grids with a Continuous AdjointFormulation,” Computers and Fluids, Vol. 28, Nos. 4–5, 1999,pp. 443–480.

[16] Ghayour, K., and Baysal, O., “Unsteady Aerodynamics and ShapeOptimization Using Modified Transonic Small Disturbance Equation,”37th Aerospace Sciences Meeting and Exhibit, Reno, NV, AIAAPaper 99-0654, 1999.

[17] Yee, K., Kim, Y., and Lee, D., “Aerodynamic Shape Optimization ofRotor Airfoils Undergoing Unsteady Motion,” 37th AerospaceSciences Meeting and Exhibit, Reno, NV, AIAA Paper 99-3107, 1999.

[18] Florea, R., and Hall, K. C., “Sensitivity Analysis of Unsteady InviscidFlow Through Turbomachinery Cascades,” 38th Aerospace SciencesMeeting and Exhibit, Reno, NV, AIAA Paper 2000-0130, 2000.

[19] Jameson, A., “Time Dependent Calculations Using Multigrid, with

Applications to Unsteady Flows Past Airfoils and Wings,” AIAA 10thComputational Fluid Dynamics Conference, Honolulu, HA, AIAAPaper 91-1596, 1991.

[20] Jameson, A., Schmidt, W., and Turkel, E., “Numerical Solutions of theEuler Equations by Finite Volume Methods with Runge-Kutta TimeStepping Schemes,” AIAA Paper 81-1259, 1981.

[21] Nadarajah, S., “The Discrete Adjoint Approach to Aerodynamic ShapeOptimization,” Ph.D. Dissertation, Department of Aeronautics andAstronautics, Stanford Univ., Stanford, CA, Jan. 2003.

[22] Nadarajah, S., and Jameson, A., “Optimal Control of Unsteady FlowsUsing a Time Accurate Method,” 9th AIAA/ISSMO Symposium onMultidisciplinary Analysis andOptimization Conference, Atlanta, GA,AIAA Paper 2002-5436, 2002.

[23] Jameson, A., and Vassberg, J. C., “Studies of Alternative NumericalOptimization Methods Applied to the Brachistochrone Problem,”Computational Fluid Dynamics Journal, Vol. 9, No. 3, 2000, pp. 281–296.

[24] Davis, S. S., “NACA 64A010 (NASA Ames model) OscillatoryPitching,” Compendium of Unsteady Aerodynamic Measurements,AGARD, Rept. R-702, Neuilly sur-Seine, France, Aug. 1982.

[25] McCroskey, W. J., “Unsteady Airfoils,” Annual Review of Fluid

Mechanics, Vol. 14, 1982, pp. 285–311.

N. AlexandrovAssociate Editor

NADARAJAH AND JAMESON 1491