Improvements to a Newton-Krylov Adjoint Algorithm for Aerodynamic Optimization

Similar documents
Aerodynamic Shape Optimization in Three-Dimensional Turbulent Flows Using a Newton-Krylov Approach

Application of Jetstream to a Suite of Aerodynamic Shape Optimization Problems. Karla Telidetzki

AIRFOIL SHAPE OPTIMIZATION USING EVOLUTIONARY ALGORITHMS

Studies of the Continuous and Discrete Adjoint Approaches to Viscous Automatic Aerodynamic Shape Optimization

Solution of 2D Euler Equations and Application to Airfoil Design

An Optimization Method Based On B-spline Shape Functions & the Knot Insertion Algorithm

Case C3.1: Turbulent Flow over a Multi-Element MDA Airfoil

Modeling External Compressible Flow

State of the art at DLR in solving aerodynamic shape optimization problems using the discrete viscous adjoint method

The Spalart Allmaras turbulence model

A High-Order Accurate Unstructured GMRES Solver for Poisson s Equation

Optimization with Gradient and Hessian Information Calculated Using Hyper-Dual Numbers

Multi-Element High-Lift Configuration Design Optimization Using Viscous Continuous Adjoint Method

An efficient method for predicting zero-lift or boundary-layer drag including aeroelastic effects for the design environment

Shape optimisation using breakthrough technologies

An Investigation of Directional-Coarsening And Line-Implicit Smoothing Applied to Agglomeration Multigrid

THE use of computer algorithms for aerodynamic shape

Introduction to ANSYS CFX

Adjoint-Based Sensitivity Analysis for Computational Fluid Dynamics

Constrained Aero-elastic Multi-Point Optimization Using the Coupled Adjoint Approach

Second Symposium on Hybrid RANS-LES Methods, 17/18 June 2007

Keywords: CFD, aerofoil, URANS modeling, flapping, reciprocating movement

THE EFFECTS OF THE PLANFORM SHAPE ON DRAG POLAR CURVES OF WINGS: FLUID-STRUCTURE INTERACTION ANALYSES RESULTS

Airfoil shape optimization using adjoint method and automatic differentiation. Praveen. C

Mesh Movement for a Discrete-Adjoint Newton-Krylov Algorithm for Aerodynamic Optimization

Case C2.2: Turbulent, Transonic Flow over an RAE 2822 Airfoil

Optimization of Laminar Wings for Pro-Green Aircrafts

Efficient Multi-point Aerodynamic Design Optimization Via Co-Kriging

Debojyoti Ghosh. Adviser: Dr. James Baeder Alfred Gessow Rotorcraft Center Department of Aerospace Engineering

Multipoint and Multi-Objective Aerodynamic Shape Optimization

Comparison of parallel preconditioners for a Newton-Krylov flow solver

Single- and Multi-Point Aerodynamic Shape Optimization Using A Parallel Newton-Krylov Approach

SELECTIVE ALGEBRAIC MULTIGRID IN FOAM-EXTEND

Grid. Apr 09, 1998 FLUENT 5.0 (2d, segregated, lam) Grid. Jul 31, 1998 FLUENT 5.0 (2d, segregated, lam)

LES Analysis on Shock-Vortex Ring Interaction

Interface and Boundary Schemes for High-Order Methods

AERODYNAMIC DESIGN OF FLYING WING WITH EMPHASIS ON HIGH WING LOADING

(c)2002 American Institute of Aeronautics & Astronautics or Published with Permission of Author(s) and/or Author(s)' Sponsoring Organization.

Adjoint Solver Workshop

OPTIMIZATIONS OF AIRFOIL AND WING USING GENETIC ALGORITHM

39th AIAA Aerospace Sciences Meeting and Exhibit January 8 11, 2001/Reno, NV

INVERSE METHODS FOR AERODYNAMIC DESIGN USING THE NAVIER-STOKES EQUATIONS

Comparisons of Compressible and Incompressible Solvers: Flat Plate Boundary Layer and NACA airfoils

Ail implicit finite volume nodal point scheme for the solution of two-dimensional compressible Navier-Stokes equations

NUMERICAL 3D TRANSONIC FLOW SIMULATION OVER A WING

High-order solutions of transitional flow over the SD7003 airfoil using compact finite-differencing and filtering

Geometry Parameterization Using Control Grids

Aerodynamic Inverse Design Framework using Discrete Adjoint Method

Multi-Mesh CFD. Chris Roy Chip Jackson (1 st year PhD student) Aerospace and Ocean Engineering Department Virginia Tech

Case C1.3: Flow Over the NACA 0012 Airfoil: Subsonic Inviscid, Transonic Inviscid, and Subsonic Laminar Flows

TAU mesh deformation. Thomas Gerhold

I. Introduction. Optimization Algorithm Components. Abstract for the 5 th OpenFOAM User Conference 2017, Wiesbaden - Germany

Three dimensional meshless point generation technique for complex geometry

Development of an Integrated Computational Simulation Method for Fluid Driven Structure Movement and Acoustics

EFFICIENT SOLUTION ALGORITHMS FOR HIGH-ACCURACY CENTRAL DIFFERENCE CFD SCHEMES

Introduction to CFX. Workshop 2. Transonic Flow Over a NACA 0012 Airfoil. WS2-1. ANSYS, Inc. Proprietary 2009 ANSYS, Inc. All rights reserved.

This is a repository copy of Nonconsistent mesh movement and sensitivity calculation on adjoint aerodynamic optimization.

Aerofoil Optimisation Using CST Parameterisation in SU2

The Continuous Adjoint Method for the Computation of First- and Higher-Order Sensitivities

Efficient Implicit Time-Marching Methods Using a Newton-Krylov Algorithm

Unstructured Mesh Solution Techniques using the NSU3D Solver

AERODYNAMIC DESIGN FOR WING-BODY BLENDED AND INLET

Computational Fluid Dynamics for Engineers

Express Introductory Training in ANSYS Fluent Workshop 04 Fluid Flow Around the NACA0012 Airfoil

Estimating Vertical Drag on Helicopter Fuselage during Hovering

High-Lift Aerodynamics: STAR-CCM+ Applied to AIAA HiLiftWS1 D. Snyder

A Scalable GPU-Based Compressible Fluid Flow Solver for Unstructured Grids

Modeling Unsteady Compressible Flow

Mesh Morphing and the Adjoint Solver in ANSYS R14.0. Simon Pereira Laz Foley

Digital-X. Towards Virtual Aircraft Design and Testing based on High-Fidelity Methods - Recent Developments at DLR -

Automatic Differentiation Adjoint of the Reynolds-Averaged Navier Stokes Equations with a Turbulence Model

Limiters for Unstructured Higher-Order Accurate Solutions of the Euler Equations

Lift Superposition and Aerodynamic Twist Optimization for Achieving Desired Lift Distributions

A linear solver based on algebraic multigrid and defect correction for the solution of adjoint RANS equations

AERODYNAMIC SHAPES DESIGN ON THE BASE OF DIRECT NEWTON TYPE OPTIMIZATION METHOD

41st AIAA Aerospace Sciences Meeting and Exhibit Jan 6 9, 2003/Reno, Nevada

Numerical Simulations of Fluid-Structure Interaction Problems using MpCCI

Fluent User Services Center

ADJOINT OPTIMIZATION OF 2D-AIRFOILS IN INCOMPRESSIBLE FLOWS

AN INVERSE DESIGN METHOD FOR ENGINE NACELLES AND WINGS

Airfoil Design Optimization Using Reduced Order Models Based on Proper Orthogonal Decomposition

Implicit/Multigrid Algorithms for Incompressible Turbulent Flows on Unstructured Grids

Comparison of B-spline Surface and Free-form. Deformation Geometry Control for Aerodynamic. Optimization

A projected Hessian matrix for full waveform inversion Yong Ma and Dave Hale, Center for Wave Phenomena, Colorado School of Mines

Analysis of the Adjoint Euler Equations as used for Gradient-based Aerodynamic Shape Optimization

INTERIOR POINT METHOD BASED CONTACT ALGORITHM FOR STRUCTURAL ANALYSIS OF ELECTRONIC DEVICE MODELS

ANSYS FLUENT. Airfoil Analysis and Tutorial

1.2 Numerical Solutions of Flow Problems

Progress and Future Prospect of CFD in Aerospace

Application of Wray-Agarwal Turbulence Model for Accurate Numerical Simulation of Flow Past a Three-Dimensional Wing-body

Grid Dependence Study of Transonic/Supersonic Flow Past NACA Air-foil using CFD Hemanth Kotaru, B.Tech (Civil Engineering)

Estimation of Flow Field & Drag for Aerofoil Wing

UNSTRUCTURED GRIDS ON NURBS SURFACES. The surface grid can be generated either in a parameter. surfaces. Generating grids in a parameter space is

Transition Flow and Aeroacoustic Analysis of NACA0018 Satish Kumar B, Fred Mendonç a, Ghuiyeon Kim, Hogeon Kim

Supersonic Wing Design Method Using an Inverse Problem for Practical Application

Driven Cavity Example

computational Fluid Dynamics - Prof. V. Esfahanian

THE DEVELOPMENT OF THE POTENTIAL AND ACADMIC PROGRAMMES OF WROCLAW UNIVERISTY OF TECH- NOLOGY ITERATIVE LINEAR SOLVERS

Morphing Wings: A Study Using Aerodynamic Shape Optimization

Faculty of Mechanical and Manufacturing Engineering, University Tun Hussein Onn Malaysia (UTHM), Parit Raja, Batu Pahat, Johor, Malaysia

Transcription:

Improvements to a Newton-Krylov Adjoint Algorithm for Aerodynamic Optimization David W. Zingg, Timothy M. Leung, Laslo Diosady, Anh H. Truong, and Samy Elias Institute for Aerospace Studies, University of Toronto 4925 Dufferin St., Toronto, Ontario M3H 5T6, Canada Marian Nemec NASA Ames Research Center, Moffett Field, CA 9435 We present and examine a number of improvements to a gradient-based algorithm for aerodynamic optimization. A Newton-Krylov algorithm is used to solve the compressible Navier-Stokes equations, the gradient is computed using the discrete-adjoint method with a preconditioned Krylov solver, and the optimum is found through a quasi-newton algorithm with a rank-two update formula. Constraints are imposed by penalizing the objective function. Improvements are made in three areas: 1) thickness constraints are generalized to permit the location of maximum thickness to be determined by the optimizer or alternatively to constrain the cross-sectional area; 2) new scalings of design variables and initial estimates of the inverse Hessian matrix in the quasi-newton method are investigated; 3) the algebraic grid perturbation algorithm is replaced by an algorithm based on a spring analogy. In each case, the effect of the improvements on the performance of the algorithm is presented. I. Introduction Numerical aerodynamic shape optimization can be of great use in the design process. It relieves the designer of the tedious task of searching the design space and thereby enables added focus on the specification of the design problem, i.e. the definition of the design space, the objectives, and the constraints. The adjoint method, whether applied in a discrete or a continuous manner, has proven to be an efficient approach to aerodynamic optimization when the number of design variables is large. 1, 2 Nemec and Zingg 3 presented an efficient Newton-Krylov algorithm for aerodynamic shape optimization based on the discrete adjoint method. The algorithm provides the basis for the two-dimensional aerodynamic optimization codes Optima2D, for single-element airfoils, and OptimaMB, for multi-element configurations. Their approach includes all terms needed to calculate accurate gradients, such as grid sensitivities and the linearization of the coupled turbulence model. As a result, the gradient is typically reduced by five orders of magnitude, ensuring convergence to at least a local minimum. While this can be more expensive than other approaches, a high degree of convergence is important when assessing design trade-offs. Alternatively, if quick turnaround is a high priority, one can run a fixed number of iterations, since the majority of the overall improvement in the objective function typically occurs in a fraction of the total number of iterations needed to converge fully (although the same cannot be said of the shape changes). The formulation of Nemec and Zingg has been applied to numerous design problems, including multipoint and multi-objective problems 4 and high-lift configurations. 5 The objective of this paper is to describe and demonstrate three improvements to Optima2D. The first provides more flexibility in the specification of geometric constraints. The second modifies the scaling of Professor, Senior Canada Research Chair in Computational Aerodynamics, 24 Guggenheim Fellow, Senior AIAA Member. Graduate Student Undergraduate Student NRC Research Associate, presently Research Scientist, ELORET Corp., Applications Branch, Advanced Supercomputing Division, AIAA Member. 1 of 11

the design variables and the initial estimate of the Hessian in the quasi-newton method in order to provide faster and more reliable convergence. The final improvement is a more robust procedure for perturbing the grid in response to shape changes. Examples are provided showing the benefits of each improvement. II. Algorithm Description The present algorithm parameterizes the geometry using B-splines. The B-spline control points and the angle of attack are the design variables. The compressible Navier-Stokes equations are solved with a Newton- Krylov method in which the linear system arising at each Newton iteration is solved using the generalized minimal residual method (GMRES) preconditioned with an incomplete lower-upper (ILU) factorization with limited fill. The Spalart-Allmaras turbulence model is used to compute the eddy viscosity. The gradient is calculated using the discrete adjoint method; solution of the adjoint equation is accomplished through the same preconditioned Krylov method. Geometric constraints are added to the objective function as penalty terms. A new set of design variables is computed using a quasi-newton optimizer in which an estimate of the inverse Hessian based on the BFGS (Broyden-Fanno-Goldfarb-Shannon) rank-two update formula is used to compute a search direction. 6 If the initial step does not produce sufficient progress toward the minimum, the step size is determined using a line search, which terminates when the strong Wolfe conditions are satisfied. 6 Each time a new shape is calculated, the initial grid is perturbed using a simple algebraic technique. For a complete description of the algorithm, see Nemec. 7 III. Algorithm Improvements A. Geometric Constraints In aerodynamic shape optimization, geometric constraints are typically introduced to represent structural and manufacturing requirements as well as practical considerations, such as fuel storage. In Optima2D, the designer can specify a minimum thickness at any location along the airfoil chord. Typically several such thickness constraints are introduced both for the purposes listed above and to prevent crossover during the optimization iterations. The latter are generally inactive once a converged shape is obtained. While this approach provides the designer with a great deal of control, it can sometimes be preferable to allow more flexibility to the optimizer. For example, rather than specifying a thickness to chord ratio of say 12% at 35% chord, it may be sufficient to require simply that the thickness to chord ratio be 12% at some unspecified chord location. Even more flexibility is provided if the geometric constraint is expressed in terms of a minimum cross-sectional area, rather than a minimum thickness. B. Scaling and the Initial Inverse Hessian Estimate The quasi-newton algorithm determines a search direction based on the gradient and an approximation to the inverse of the Hessian matrix. The first search direction is equivalent to the steepest-descent direction. After the first step, an initial estimate of the inverse Hessian is required. The BFGS secant update is then used to generate a sequence of matrices that approximate the inverse of the exact Hessian with increasing accuracy. The convergence of the quasi-newton method can be affected by the scaling of the design variables. 8, 9 Although the scaling does not affect the BFGS updates themselves, it influences the first step and therefore the initial estimate of the inverse Hessian, and this influence persists throughout the optimization. A poor initial inverse Hessian leads to poor inverse Hessian estimates in subsequent iterations, causing slow convergence and possibly line-search stall. Nemec and Zingg scaled each design variable by its initial value with a correction if the initial value is zero. 3 Hence the scaled design variables x are related to the unscaled design variables x by the following equation: x = L 1 x (1) where L is a diagonal matrix containing the initial values of the unscaled design variables. The ideal scaling is related to the square root of the diagonal entries of the Hessian matrix and is not directly related to the initial values of the design variables. However, the Hessian is of course unknown. 2 of 11

The initial estimate of the scaled inverse Hessian is currently given by: 6 with H = v T d v T v I (2) v = g 1 g d = x 1 x where the prime denotes scaled variables, and g is the gradient vector. An inappropriate scaling can lead to a poor initial estimate of the inverse Hessian from which the BFGS updates cannot recover. Consequently, it can be advantageous to use the unscaled estimate of the inverse Hessian. Whether the scaled or unscaled initial estimate is used, these estimates are simply scalar multiples of the identity matrix. Alternatively, we consider an initial estimate of the inverse Hessian which is a diagonal matrix with entries given by: A fix is needed if any of the entries has a value below a certain threshold. C. Grid Perturbation H ii = d i v i (3) The algebraic grid perturbation technique has proven to be effective for a range of optimization problems. For problems such as finding the optimum position of a flap, however, it can lead to grids of insufficient quality, and periodic regridding can be required. Regridding often involves human intervention, especially in three dimensions, and is thus undesirable. Poor quality grids can also result from the simple algebraic perturbation technique when the shape changes are substantial. For example, if the angle of the trailing-edge bisector changes too much, grid distortion can be excessive. In order to avoid regridding, we consider an alternative grid perturbation technique based on the lineal 1, 11 spring analogy. The purpose of a grid perturbation technique is to propagate the deformation of the boundary, or more specifically, the surface grid, into the interior of the grid while maintaining grid quality. In the lineal spring method, the edges of the elements act as springs whose stiffness is proportional to the edge length, i.e. [ k ij = 1/ (x j x i ) 2 + (y j y i ) 2] 1 2 (4) The original grid is in equilibrium. When the surface nodes move in response to a shape change, forces are generated due to the displacement. The force at node i is N i F i = k ij (δ j δ i ) (5) j=1 where N i is the number of neighbouring nodes of node i, and δ i is the displacement of node i from its position in the original grid. The interior nodes must thus move such that equilibrium is restored, i.e. F i =. This leads to a linear system of equations, which is solved using the conjugate gradient method. Since the grid cells near the airfoil surface tend to be small, they have high spring stiffness, and thus their deformation is minimal. A key aspect of this procedure in the present context is that the grid perturbation is always carried out relative to the original grid around the initial shape, not relative to the grid generated from the previous optimization iteration. Therefore, there is a unique grid and hence a unique solution for a given set of design variables. This is an important property for an optimization algorithm. IV. Results A. Geometric Constraints The geometric constraints are studied in the context of lift-to-drag ratio maximization at a Mach number of.75 and a Reynolds number of 2.88 million. Fully turbulent flow is assumed, i.e. transition is assumed to 3 of 11

Figure 1. B-spline control points; the shaded control points are used as design variables. -1.5 RAE 2822 Fixed Thickness Constraint -1 -.5 Cp.5 1.2.4.6.8 1 x/c Figure 2. Surface pressure coefficient distribution and airfoil shapes: RAE 2822 airfoil and optimum airfoil with fixed thickness constraint at x/c =.35. occur at the leading edge. The initial geometry is the RAE 2822 airfoil. Seventeen B-spline control points are used to parameterize the geometry. Three control points are frozen at the leading edge and four at the trailing edge. Hence there are eleven design variables including the angle of attack. Figure 1 shows the B-spline control points. The grid has a C topology with 289 nodes in the streamwise direction and 65 in the normal direction. For the first example, two thickness constraints are applied. A thickness to chord ratio of at least 11% is imposed at x/c =.35, and a.2% thickness to chord ratio is imposed at x/c =.99. The latter constraint is introduced to prevent crossover near the trailing edge. The pressure distributions and airfoil shapes are shown in Figure 2 for the initial (RAE 2822) and final airfoils. The shock wave has been eliminated on the upper surface. The resulting airfoil has lift and drag coefficients of.8195 and.147, respectively, at an angle of attack of 2.56 degrees, for a lift-to-drag ratio of 55.31. The same optimization problem is repeated with a minimum thickness requirement of 11%, but no particular value of x/c is specified. The resulting airfoil (labelled floating thickness constraint ) is compared with the previous airfoil in Figure 3. The location of maximum thickness has moved slightly aft. The resulting lift and drag coefficients are.8286 and.148, respectively, at an angle of attack of 2.72 degrees, giving a lift to drag ratio of 55.84. The optimizer has exploited the additional freedom to improve the lift to drag ratio by roughly 1%. The benefit is relatively small because the specified location of the maximum thickness is not far from the optimum location of maximum thickness. However, in other problems, for example in the 4 of 11

-1.5 Fixed Thickness Constraint Floating Thickness Constraint -1 -.5 Cp.5 1.2.4.6.8 1 x/c Figure 3. Surface pressure coefficient distribution and airfoil shapes: optimum airfoil with fixed thickness constraint at x/c =.35 and optimum airfoil with floating thickness constraint. design of natural-laminar-flow airfoils, the location of maximum thickness may be difficult to guess. Unless there is a specific reason to constrain the location of maximum thickness, such as the presence of a spar, it is preferable not to do so. Finally, the same optimization is repeated without any thickness constraints. The area is required to be equal to the cross-sectional area of the airfoil generated using the floating thickness constraint. This leads to lift and drag coefficients of.8368 and.141, respectively, at an angle of attack of 2.55 degrees, resulting in a lift to drag ratio of 59.22. This is roughly 6% higher than the lift to drag ratio of the airfoil designed using the floating thickness constraint. Furthermore, the resulting shape is smoother with less pinching near the leading edge and a more uniform load distribution, as shown in Figure 4. B. Scaling and the Initial Inverse Hessian Estimate Two optimization problems are used to examine the effect of the design-variable scaling and the initial inverse Hessian estimate. The first is a lift-constrained drag minimization with C l =.733. The second is a maximization of the lift-to-drag ratio. The following parameters are common to both optimization problems: M =.74, Re = 2.7 million 3 B-spline control points with two frozen at the leading edge and two frozen at the trailing edge; 27 design variables including the angle of attack thickness constraints: t/c.124 at x/c =.35, t/c.5 at x/c =.96, t/c.12 at x/c =.99 257 57 grid Figure 5 shows the converged airfoil and surface pressure coefficient distribution for the lift-constrained drag minimization, which has C l =.7237, C d =.1432. The corresponding values for the lift-to-drag ratio maximization are C l =.9354, C d =.1731, for a lift-to-drag ratio of 54. The following four design-variable scalings have been studied: 5 of 11

-1.5 Floating Thickness Constraint Area Constraint -1 -.5 Cp.5 1.2.4.6.8 1 x/c Figure 4. Surface pressure coefficient distribution and airfoil shapes: optimum airfoil with floating thickness constraint and optimum airfoil with area constraint. -1 Initial Final -.5 Cp.5 1.2.4.6.8 1 x/c Figure 5. Surface pressure coefficient distribution and airfoil shape for lift-constrained drag minimization. 6 of 11

.8.75 Objective Function.7.65.6.55.5 Hessian A Scaling Hessian A Scaling 1 HessianCScaling2.45.4.35 1 2 3 Number of Flow Solutions and Gradient Evaluations Figure 6. Convergence histories for the lift-constrained drag minimization problem. Scaling : each design variable is scaled by its initial value (with a fix if the initial value is zero) Scaling 1: the angle of attack design variable is scaled by its initial value; the B-spline control point design variables are scaled by an average value Scaling 2: each design variable is scaled by the square root of its initial value (with a fix if the initial value is zero) Scaling 3: no scaling is applied Scaling 2 is based on the observation that Scaling sometimes provides too much scaling, while Scaling 3 often provides too little. In addition, the following three initial inverse Hessian estimates have been examined: Hessian A: the original approach given by Eq. 2 Hessian B: similar to Hessian A, but calculated using unscaled variables Hessian C: the diagonal matrix given by Eq. 3 All possible combinations of scalings and initial inverse Hessian estimates have been applied to numerous optimization problems. The most promising combinations are Hessian A with Scaling 1 and Hessian C with Scaling 2. Convergence histories for these two combinations are compared with the baseline scheme, Hessian A with Scaling, in Figures 6 and 7. The lift-constrained drag minimization problem is shown in Figure 6; the lift-to-drag ratio maximization is shown in Figure 7. The results for the two problems are quite different. For the lift-constrained drag minimization, the Hessian C-Scaling 2 combination converges fastest, requiring 96 iterations to converge to within.5% of the final converged objective function value. The Hessian A-Scaling 1 combination requires 174 iterations to achieve the same degree of convergence. a To converge to within.1% of the final objective function value, a Note that an iteration involves one flow solution and one gradient evaluation. 7 of 11

L/D 55 54 53 52 51 5 49 48 47 46 45 44 43 42 41 Hessian A Scaling Hessian A Scaling 1 Hessian C Scaling 2 4 1 2 3 Number of Flow Solutions and Gradient Evaluations Figure 7. Convergence histories for the lift-to-drag ratio maximization problem. the two schemes require 132 and 293 iterations, respectively. The baseline scheme stalls without converging and is much slower than the other two combinations to achieve partial convergence. For the lift-to-drag ratio maximization, the combination of Hessian A and Scaling 1 converges fastest, requiring 97 and 137 iterations to achieve.5% and.1% convergence levels, respectively. The Hessian C-Scaling 2 combination requires 317 and 318 iterations, respectively, while the original Hessian A-Scaling combination requires 345 and 346. The savings associated with the new schemes relative to the baseline scheme are substantial. For the lift-constrained drag minimization, the Hessian C-Scaling 2 combination converges almost twice as fast as the Hessian A-Scaling 1 combination, and the Hessian A-Scaling combination does not converge. For the lift-to-drag ratio maximization, the combination of Hessian A with Scaling 1 is two to three times faster than both of the other options. Similar results are obtained for both objective functions with different numbers of design variables. Clearly, there is a significant benefit to choosing the best scaling and initial inverse Hessian estimate for a given optimization problem. The difficulty is that at this point we do not understand why these combinations behave differently for the two objective functions. Therefore, pending further study, we recommend the Hessian C-Scaling 2 combination for lift-constrained drag minimization problems and the Hessian A-Scaling 1 combination for lift-to-drag ratio maximization. C. Grid Perturbation The first example is an optimization of the flap position in a two-element geometry which was previously studied by Nemec. 7 The objective is to find the location of the flap that maximizes the lift coefficient without increasing the drag coefficient. The angle of attack and the flap angle are held fixed. Hence there are two design variables, the coordinates of the flap. The initial grid in the vicinity of the main airfoil trailing edge is shown in Figure 8. Figure 9 shows the new flap position and grid obtained using the algebraic grid perturbation technique. The grid is badly skewed in the region between the grid line emanating from the main airfoil trailing edge and the upper surface of the flap. The grid resulting from the spring-analogy technique is displayed in Figure 1. This grid is less skewed and more comparable to the initial grid. We are currently examining the quantitative implications of this apparent improvement in grid quality. Figure 11 shows another example of a grid generated using the spring-analogy technique with large shape 8 of 11

.5 Y -.5.9.92.94.96.98 1 1.2 1.4 1.6 X Figure 8. Original grid, showing the region between the trailing edge of the main airfoil and the flap..5 Y -.5.9.92.94.96.98 1 1.2 1.4 1.6 X Figure 9. Grid after flap movement using algebraic grid perturbation method..5 Y -.5.9.92.94.96.98 1 1.2 1.4 1.6 X Figure 1. Grid after flap movement using spring-analogy grid perturbation method. 9 of 11

.6.4.2 Y -.2 -.4 -.6 -.2.2.4.6.8 1 1.2 X Figure 11. Spring-analogy grid for the subsonic lift-to-drag ratio maximization problem. changes. The initial airfoil is the NACA 12. The objective is to maximize the lift-to-drag ratio at a Mach number of.25 and a Reynolds number of 2.88 million. The following three thickness constraints are imposed: a thickness to chord ratio of at least 4% at x/c =.5, 4% at x/c =.65, and 1.2% at x/c =.95. The final airfoil has a lift-to-drag ratio of 61 under these operating conditions. The grid perturbation technique based on the spring analogy leads to somewhat more reliable convergence without regridding than the algebraic technique for cases such as this with large shape changes. Figure 12 shows the grid produced using the spring-analogy technique for a lift-to-drag ratio maximization at transonic speed. The initial airfoil is again the NACA 12, so the shape changes are large, especially on the lower surface near the leading and trailing edges. The Mach number is.75, the Reynolds number, 2.88 million. The thickness to chord ratio is constrained to be at least 11% at x/c =.35, 3% at x/c =.2, 2% at x/c =.85, and.2% at x/c =.99. In this case, the optimization using the algebraic grid perturbation technique stalled after a small reduction in the gradient, while the spring analogy technique led to a gradient reduction of several orders of magnitude. These results must be considered preliminary. Although the spring-analogy technique shows promise, we have yet to generate conclusive evidence that it offers a quantifiable advantage over the algebraic technique. This investigation is still underway, and an approach based on a pseudo-elasticity model is also under study. 12 V. Conclusions We have presented improvements to a Newton-Krylov discrete-adjoint algorithm for aerodynamic shape optimization, including more flexible geometric constraints, a new design-variable scaling, a new initial estimate of the inverse Hessian, and an improved grid perturbation technique. Examples show that the area constraint leads to a more efficient, smoother airfoil, while the improvements to the optimizer lead to substantially faster and more reliable convergence. Acknowledgments The funding of the first author by the Natural Sciences and Engineering Research Council of Canada and the Canada Research Chairs program is gratefully acknowledged as is the funding of the last author through the National Research Council Research Associateship Award. 1 of 11

.6.4.2 Y -.2 -.4 -.6 -.2.2.4.6.8 1 1.2 X Figure 12. Spring-analogy grid for the transonic lift-to-drag ratio maximization problem. References 1 Pironneau, O., On Optimum Design in Fluid Mechanics, J. Fluid Mech., Vol. 64, No. 1, 1974, pp. 97-11. 2 Jameson, A., Aerodynamic Design Via Control Theory, NASA CR-181749, Nov. 1988. 3 Nemec, M., and Zingg, D.W., Newton-Krylov Algorithm for Aerodynamic Design Using the Navier-Stokes Equations, AIAA J., Vol. 4, No. 6, 22, pp. 1146-1154. 4 Nemec, M., Zingg, D.W., and Pulliam, T.H., Multipoint and Multi-Objective Aerodynamic Shape Optimization, AIAA J., Vol. 42, No. 6, 24, pp. 157-165. 5 Nemec, M., and Zingg, D.W., Optimization of High-Lift Configurations Using a Newton-Krylov Algorithm, AIAA Paper 23-3957, June 23. 6 Nocedal, J. and Wright, S.J., Numerical Optimization, Springer-Verlag, New York, 1999. 7 Nemec, M., Optimal Shape Design of Aerodynamic Configurations: A Newton-Krylov Approach, Ph.D. Thesis, University of Toronto Institute for Aerospace Studies, 22. 8 Gill, P.E., Murray, W., and Wright, M.H., Practical Optimization, Academic Press, Toronto, 1981. 9 Liu, P.-Y., and Zingg, D.W., Comparison of Optimization Algorithms Applied to Aerodynamic Design, AIAA Paper 24-454, Jan. 24. 1 Batina, J.T., Unsteady Euler Airfoil Solutions Using Unstructured Dynamic Meshes, AIAA J., Vol. 28, No. 8, 199, pp. 1381-1388. 11 Crumpton, P.I., and Giles, M.B., Implicit Time-Accurate Solutions on Unstructured Dynamic Grids, International Journal for Numerical Methods in Fluids, Vol. 25, 1997, pp. 1285-13. 12 Yang, Z., and Mavriplis, D.J., Unstructured Dynamic Meshes with Higher-order Time Integration Schemes for the Unsteady Navier-Stokes Equations, AIAA Paper 25-1222, Jan. 25. 11 of 11