+ All Categories
Home > Documents > Adjoint Optimization of wind Farm Layouts for Systems...

Adjoint Optimization of wind Farm Layouts for Systems...

Date post: 21-Mar-2018
Category:
Upload: doankiet
View: 217 times
Download: 4 times
Share this document with a friend
16
Adjoint Optimization of Wind Farm Layouts for Systems Engineering Analysis * Ryan N. King Department of Mechanical Engineering, University of Colorado, Boulder, CO 80309, USA. National Renewable Energy Laboratory, Golden, CO, 80401, USA. Peter E. Hamlington Department of Mechanical Engineering, University of Colorado, Boulder, CO 80309, USA. Peter Graf National Renewable Energy Laboratory, Golden, CO, 80401, USA. Katherine Dykes National Renewable Energy Laboratory, Golden, CO, 80401, USA. Wind is an important renewable energy resource, but in order for it to be competitive with other energy sources, wind farm layouts must be optimized to reduce the levelized cost of energy (LCOE). This requires an analysis that integrates high-fidelity flow models with physical turbine and cost models. We present a new flow model and turbine optimization tool, WindSE, that optimizes wind turbine locations and axial induction factors as part of the National Renewable Energy Laboratory’s open source wind energy systems engineering software tool WISDEM. WindSE enables gradient-based layout optimization by solving the discrete adjoint equations corresponding to the forward flow model. This provides a computationally efficient way to calculate gradients used in the optimization process. We discuss the development of WindSE and provide optimization results based on simulations of single and multiple inflow directions. In the case of strongly directional wind roses the optimal layouts take advantage of flow curvature and speedup effects that are not generated by standard industry linear flow models. When optimizing with respect to a uniform wind rose, local speedup effects are found to be less beneficial and the optimal layouts instead reduce wake losses by maximizing spacing. Sensitivity to wind direction bin sizes and weights is demonstrated with a full layout and axial induction factor optimization using a wind rose from an offshore wind farm. I. Introduction Optimizing wind turbine locations within a wind farm is a difficult problem due to the strong nonlinear coupling between turbine locations, power production, atmospheric boundary layer turbulence, and mechan- ical loads on turbine components. Existing wind farm optimization techniques used by industry often rely on heuristic guidelines and simplified linear flow models. Traditional approaches to optimizing turbine layouts have focused on maximizing total power output from linear flow models, but this does not take into account the costs of constructing, operating, and maintaining turbines within a farm. If not properly assessed, these * The submitted manuscript has been offered by employees of the Alliance for Sustainable Energy, LLC (Alliance), a contractor of the U.S. Government under Contract No. DE-AC36-08GO28308. Accordingly, the U.S. Government and Alliance retain a nonexclusive royalty-free license to publish or reproduce the published form of this contribution, or allow others to do so, for U.S. Government purposes.4 1 of 16 American Institute of Aeronautics and Astronautics Downloaded by UNIVERSITY OF COLORADO BOULDER on February 16, 2016 | http://arc.aiaa.org | DOI: 10.2514/6.2016-2199 34th Wind Energy Symposium 4-8 January 2016, San Diego, California, USA AIAA 2016-2199 Copyright © 2015 by Ryan N. King, Peter E. Hamlington, Peter Graf, and Katherine Dykes. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission. AIAA SciTech
Transcript

Adjoint Optimization of Wind Farm Layouts for

Systems Engineering Analysis∗

Ryan N. King

Department of Mechanical Engineering,

University of Colorado, Boulder, CO 80309, USA.

National Renewable Energy Laboratory, Golden, CO, 80401, USA.

Peter E. Hamlington

Department of Mechanical Engineering,

University of Colorado, Boulder, CO 80309, USA.

Peter Graf

National Renewable Energy Laboratory, Golden, CO, 80401, USA.

Katherine Dykes

National Renewable Energy Laboratory, Golden, CO, 80401, USA.

Wind is an important renewable energy resource, but in order for it to be competitivewith other energy sources, wind farm layouts must be optimized to reduce the levelized costof energy (LCOE). This requires an analysis that integrates high-fidelity flow models withphysical turbine and cost models. We present a new flow model and turbine optimizationtool, WindSE, that optimizes wind turbine locations and axial induction factors as part ofthe National Renewable Energy Laboratory’s open source wind energy systems engineeringsoftware tool WISDEM. WindSE enables gradient-based layout optimization by solvingthe discrete adjoint equations corresponding to the forward flow model. This provides acomputationally efficient way to calculate gradients used in the optimization process. Wediscuss the development of WindSE and provide optimization results based on simulationsof single and multiple inflow directions. In the case of strongly directional wind roses theoptimal layouts take advantage of flow curvature and speedup effects that are not generatedby standard industry linear flow models. When optimizing with respect to a uniform windrose, local speedup effects are found to be less beneficial and the optimal layouts insteadreduce wake losses by maximizing spacing. Sensitivity to wind direction bin sizes andweights is demonstrated with a full layout and axial induction factor optimization using awind rose from an offshore wind farm.

I. Introduction

Optimizing wind turbine locations within a wind farm is a difficult problem due to the strong nonlinearcoupling between turbine locations, power production, atmospheric boundary layer turbulence, and mechan-ical loads on turbine components. Existing wind farm optimization techniques used by industry often rely onheuristic guidelines and simplified linear flow models. Traditional approaches to optimizing turbine layoutshave focused on maximizing total power output from linear flow models, but this does not take into accountthe costs of constructing, operating, and maintaining turbines within a farm. If not properly assessed, these

∗The submitted manuscript has been offered by employees of the Alliance for Sustainable Energy, LLC (Alliance), a contractorof the U.S. Government under Contract No. DE-AC36-08GO28308. Accordingly, the U.S. Government and Alliance retain anonexclusive royalty-free license to publish or reproduce the published form of this contribution, or allow others to do so, forU.S. Government purposes.4

1 of 16

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F C

OL

OR

AD

O B

OU

LD

ER

on

Febr

uary

16,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

2199

34th Wind Energy Symposium

4-8 January 2016, San Diego, California, USA

AIAA 2016-2199

Copyright © 2015 by Ryan N. King, Peter E. Hamlington, Peter Graf, and Katherine Dykes. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission.

AIAA SciTech

costs can become prohibitive even for wind farms where power output is maximized, thereby limiting theeconomic competitiveness of wind energy compared to other renewable and non-renewable sources of elec-tricity. Recent studies1,2 have also used higher fidelity flow models to find optimal turbine spacing or controlsettings in an operational context, but a full layout optimization that considers multiple inflow directionshas not yet been attempted.

In order to develop a more rigorous approach to wind turbine layout optimization, a systems engineer-ing analysis that considers the Levelized Cost of Energy (LCOE) with high fidelity simulations is required.With an LCOE-based optimization routine, tradeoffs between power production, operations and mainte-nance (O&M) costs, and construction costs can be quantified, thus helping developers determine the mosteconomically competitive wind farm designs. A key challenge in using LCOE as an optimization metricis in estimating O&M costs. These costs are due in large part to damage-equivalent loads on mechanicalcomponents due to atmospheric turbulence. These loads are strongly coupled to unsteady coherent turbu-lence caused by wake interactions and atmospheric instability, and their accurate prediction places enormousdemands on the spatial and temporal resolution of flow field data used in the optimization procedure. Fur-thermore, real wind farms experience inflow from practically all directions but often have one or two primaryinflow directions. Capturing the correct distribution of wind directions in the flow modeling is crucial tooptimizing the power production and structural loads of a particular layout.

To address these challenges, we present WindSE, a wind farm layout optimizer that uses high fidelity flowmodels and gradient-based optimization enabled by adjoint techniques. The layout optimization software ispart of the National Renewable Energy Laboratory open source wind energy systems engineering analysispackage WISDEM.3 WISDEM offers a modular framework that allows users to interchange different modelsfor the physical turbine, capital costs, O&M costs, balance of station (BOS) costs, and the site flow model.This facilitates multi-fidelity LCOE-based analysis of wind farm designs that accounts for multi-directionalinflow.

The present paper is organized as follows: a background on wind farm layout optimization and adjointtechniques is provided in Section II, the specific wind farm optimization problem, constraints, and turbinerepresentation are described in Section III, details of the numerical implementation of WindSE are covered inSection IV, the analysis and optimization capabilities of WindSE are demonstrated on several test cases withincreasingly complex wind roses in Section V, and we discuss our results and plans for continued developmentin Section VI.

II. Wind Farm Design and Optimization Background

Utility scale wind farms in the U.S. typically involve tens to hundreds of turbines arranged in semi-regulararrays. The layout topology is generally an outcome of optimizing the power production subject to competinginfluences from site constraints. These constraints include the patchwork of viable building areas formed byleases and setbacks from environmental concerns or physical infrastructure, terrain and soil characteristicslike slope or vegetation, turbine manufacturer spacing requirements, and continuity requirements imposedby access roads and electrical connections. This results in a complex design problem with turbine layoutsvarying drastically between different geographic locations and exhibiting complex topologies.

The annual energy production (AEP) from a wind farm layout has traditionally been, and generallycontinues to be, assessed using reduced order engineering wind flow models. These models estimate therelative wind resource across a site based on the linearized Navier-Stokes equations and treat terrain fea-tures as perturbations in boundary conditions. The underlying governing equations are based on analyticalperturbation solutions to flow over a low hill introduced by Jackson and Hunt4 in 1975 and implementedin packages like MS3DJH/3R.5 This approach decomposes the terrain into sinusoidal hills and calculatesrelative speedup effects over each hill. The velocity deficit from a single turbine is then superimposed on thebackground wind resource at each turbine location. The PARK model developed by Jensen6 and the eddyviscosity model developed by Ainslie7 are the most common models used. These approaches decouple thewake calculation from the wind flow calculation, and further assume that the wake from a single turbine inisolation is appropriate for all turbines in a utility scale farm. Since these models are relatively inexpensiveto run, they are amenable to gradient-free optimization techniques, such as genetic algorithms, which canrequire orders of magnitude more iterations than gradient-based optimization techniques. However, the lin-ear superposition principles that underpin this approach result in poor accuracy in complex terrain, complexflow conditions, or farms with many rows of turbines.

2 of 16

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F C

OL

OR

AD

O B

OU

LD

ER

on

Febr

uary

16,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

2199

A. Past and Present Approaches to Wind Farm Optimization

Wind farm layout optimization has been performed to date primarily using engineering flow models withanalytical wake deficits. These models are inexpensive to run but lack the fidelity to accurately captureturbulent flow effects. Their relatively cheap computational cost allows gradient-free methods to be used inthe optimization process. Studies utilizing a gradient-free approach have looked at layouts that minimizewake loss,8 maximize net present value,9 and minimize noise propagation.10 Other optimization approachesinclude particle swarm optimizations that determine turbine layouts and rotor diameters,11 extended patternsearches for multimodal layout optimization,12 and even game theoretic methods.13 Fleming et al.14 usedfinite difference gradients with engineering wake models in a recent optimization study. An exhaustivereview of wind farm optimization efforts has been compiled by Herbet-Acero et al.15 The results of theseoptimizations are heavily influenced by the analytical wake models, and consequently we expect significantdifferences in the optimal layout when higher fidelity flow models are used.

Recently, higher fidelity flow models have been used in a limited range of wind farm optimization ap-plications. The Technical University of Denmark (DTU) has developed TOPFARM16 which employs animproved wake model, the dynamic wake meandering model, as well as a parabolic Navier-Stokes solver.TOPFARM employs a hybrid optimization approach that combines sequential linear programming (SLP)with gradient-free genetic algorithms. The genetic algorithm is used to find the neighborhood of the globaloptimum, then the gradient-based SLP algorithm takes over. The genetic algorithm step still penalizes largedesign spaces, limiting TOPFARM to relatively small layouts.

Large eddy simulations (LES) are also increasingly being applied to wind farm modeling,17–19 and in somecases for layout optimization. Meneveau and collaborators have extensively studied the effects of spacing andalignment in infinite wind farms with turbines in regular gridded turbine arrays.20 Their studies are basedon pseudospectral LES that captures many of the most important atmospheric boundary layer turbulenceeffects and complex wake interactions, but consider only the spacing between predefined structured rowsin their optimization. Goit and Meyers1 have used similar LES methods with adjoint techniques to studyoptimal wind farm control by adjusting rotor properties in an operational context. This benefits from highfidelity fluid dynamics and leverages the power of adjoint optimization, but once again the optimization hasbeen simplified by using fixed layouts. Funke et al.21 studied adjoint optimization of ocean turbine layouts,but they used the shallow water equations making their results incompatible with atmospheric boundarylayer flow fields.

The present approach is to use adjoint optimization techniques to enable gradient-based optimization ofwind turbine layouts in realistic atmospheric boundary layer flows. Laminar flow, Reynolds Averaged Navier-Stokes (RANS), and LES solvers are being developed that can be used interchangeably in the WISDEMframework. This allows for efficient, high fidelity layout optimization capabilities that do not currently exist.Furthermore, we emphasize optimization in the presence of multi-directional inflow angles that better reflectreal wind farm atmospheric conditions.

B. Adjoint Optimization Enables High Fidelity Flow Models

It is well known that engineering models based on simplifications of the governing equations are inaccurate inwind farms with complex terrain, non-neutral atmospheric stability, and multiple turbine rows. High fidelityflow models that account for terrain, stability, and turbulence are necessary to improve the accuracy of AEPand loading predictions. These models incur greater computational costs than existing engineering models,which makes gradient-free optimization approaches intractable. Instead, gradient-based optimizations arenecessary to reduce the number of iterations and expensive function evaluations. A conventional finitedifference approach to calculating the necessary gradients is also prohibitively expensive because it requiresa function evaluation for each design variable. For utility scale wind farms with several hundred positionvariables this negates many of the advantages of gradient-based optimization.

Fortunately, the necessary gradients can be obtained relatively inexpensively using adjoint optimizationtechniques. An adjoint approach allows one to calculate gradients at a cost that scales with the number ofobjective functions rather than the number of design variables. For wind farm optimization with a singleLCOE objective function this means that gradients can be found by solving the adjoint or dual partialdifferential equation (PDE) at roughly the same cost as a single forward model function evaluation.

Adjoint techniques have a rich history in fluid dynamics. They were have been greatly successful inairfoil shape optimization, and recently have been used to evaluate uncertainty and errors with the intent

3 of 16

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F C

OL

OR

AD

O B

OU

LD

ER

on

Febr

uary

16,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

2199

of improving meshes. Adjoint techniques for PDE constrained optimization were popularized by Lions,22

later expanded to flow control by Pironneau,23 and finally adapted to fluid dynamics by Jameson,24–27

who applied control theory to airfoil shape optimization. Giles and Pierce28–31 put the adjoint approachon a solid mathematical foundation by investigating the behavior of boundary conditions, quasi-1D Eulerequations, shocks, and various formulations of the continuous adjoint. Reviews of implementation of discreteand continuous adjoints in optimization practices have been published by Jameson,32 Giles and Pierce,33

and Giannakoglou and Papadimitriou.34

Adjoint techniques have seen uses in many other fields and applications, including geophysical flows suchas mantle convection,35 numerical weather prediction36 and data assimilation.37 Other applications includemesh optimization,38,39 error and uncertainty analysis,40–43 and large scale multi-disciplinary analysis andoptimization.44,45 Investigation of unsteady adjoints of chaotic systems by Wang and Blonigan46–48 shows astrong sensitivity to initial conditions, numerical accuracy, and important dynamical properties. This makesfinding adjoint gradients of chaotic PDEs, like the LES equations, very challenging.

C. Background on Adjoint Optimization Techniques

Adjoint operators, adjoint equations, and adjoint variables arise naturally from consideration of dynamicalsystems whose behavior can be described by differential equations. All that is required is an ordinary orpartial differential equation governing the evolution of a dynamical system and a function of the outputthat measures a quantity of interest called an objective function. The forward operator is the differentialequation that takes independent control variables as inputs and advances the state variables forward in time.The adjoint operator maps perturbations in the input variables to gradients of the objective function. Theseadjoint variables are, in a sense, optimal open-loop control inputs for the given objective function.

The adjoint operator is defined using an inner product 〈·, ·〉 by the bilinear identity 〈Au, v〉 = 〈u,Bv〉,where B is adjoint to A. This holds if A is a continuous differential operator, in which case B is found throughintegration by parts, or if A is a matrix, in which case B = AT , the conjugate transpose of A. This bilinearidentity reveals that the adjoint operator is implicitly defined by the forward model and can achieve thesame scalar output as A but with a different order and size of matrix-vector multiplications. Fundamentally,it is this change in the order and dimension of matrix-vector multiplications that lends adjoint methods theircomputational power. In the following, we will show how utilizing this identity while calculating gradientsof an objective function achieves dramatic computational savings. Due to changes in the size and order ofmatrix-vector products inherent in the adjoint process, the computational cost of finding gradients scaleswith the dimension of the cost function, which is typically just a scalar, instead of with the dimension of thedesign variable space.

Consider a dynamical system with governing equations that can be expressed in residual form

F(u,m) ≡ 0 (1)

where F is a vector-valued differential equation, u ∈ Rn is the system state vector, and m ∈ Rm is avector of input parameters. Additionally, consider a scalar objective functional J(u,m) ∈ R that measuresa quantity of interest. Many engineering problems can be formulated as optimization problems that seekoptimal control parameters to minimize J :

minimize J (u,m) (2)

subject to F (u,m) = 0

h (m) = 0

g (m) ≤ 0

where h and g are equality and inequality constraints on the control parameter m, such as upper and lowerbounds on a control input. In common engineering problems, F is a PDE that is expensive to evaluate andthe dimension of the control space Rm is high. For example, m could contain the initial conditions for anumerical weather prediction model with F being the forward operator of the weather model, or m mightdescribe the shape of an airfoil that is being optimized for lift subject to the Euler equations in F . Efficientlysolving this PDE-constrained optimization problem requires algorithms that scale well to high dimensionsand minimize the number of evaluations of F . Gradient-based optimization algorithms can approach secondorder convergence to local minima and can minimize the evaluations of F , but require the gradient of the

4 of 16

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F C

OL

OR

AD

O B

OU

LD

ER

on

Febr

uary

16,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

2199

objective functional with respect to all of the control parameters. This gradient, dJ/dm, is given by thechain rule as

dJ

dm=

∂J

∂m+∂J

∂u

∂u

∂m. (3)

Since J (u,m) is a user defined function, ∂J/∂m and ∂J/∂u are straightforward to compute. However,taking ∂/∂m (·) to be a row vector, ∂u/∂m ∈ Rn×m, which makes ∂u/∂m very difficult to compute for highdimensional control and state spaces. A finite difference approach to calculate this gradient would requirem evaluations of F , which is intractable for many contemporary engineering problems.

In the adjoint approach, ∂u/∂m ∈ Rn×m is eliminated from Eq. (3) by taking the derivative of the PDEconstraint F (u,m) = 0:

∂F∂m

+∂F∂u

∂u

∂m= 0 . (4)

Solving for ∂u/∂m and substituting into Eq. (3) yields:

dJ

dm=

∂J

∂m− ∂J

∂u

[∂F∂u

]−1

︸ ︷︷ ︸ΨT

∂F∂m

(5)

where Ψ is the adjoint variable and maps source perturbations in F (u,m) = 0 to sensitivities of J . It isgoverned by: [

∂F∂u

]TΨ =

[∂J

∂u

]T. (6)

With a forward solution of F and the adjoint solution Ψ, the derivative of the objective function can bereadily calculated as

dJ

dm=

∂J

∂m−ΨT ∂F

∂m. (7)

The resulting adjoint gradients are typically more accurate than finite difference gradients, and can becalculated at a fixed cost that is independent of the dimension of m. This enables efficient gradient-basedoptimization for systems governed by computationally demanding PDEs.

III. Wind Farm Optimization Problem

Wind farm layout optimization is approached as a PDE-constrained optimization problem using theadjoint gradient theory developed in the previous section. The PDE constraint F describes fluid flow withina wind farm, the turbine locations are the control vector m, and turbine positioning rules constrain thepossible locations.

A. Optimization Problem Definition

The long term vision for the WISDEM software is to implement an objective functional that calculates LCOEbased on energy production, O&M cots, and turbine capital costs. The WindSE model is responsible for theenergy production portion, and for the purposes of model testing we focus solely on an energy productionoptimization. We seek to maximize the steady state power output of N turbines experiencing K different

5 of 16

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F C

OL

OR

AD

O B

OU

LD

ER

on

Febr

uary

16,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

2199

inflow directions by posing the following optimization problem:

minm=[x,y,a]

J = −K∑

m=1

N∑n=1

αk1

2ρACp,n‖u (xn, yn) ‖3 (8)

subject to

uj∂ui∂xj

= −1

ρ

∂p

∂xi+ ν

∂2ui∂x2j

+1

ρ

N∑n=1

fn,AD (9)

∂ui∂xi

= 0 (10)

fn,AD =1

2ρACt,n‖u (xn, yn) ‖2 (11)

Cp,n = 4an (1− an)2

(12)

Ct,n = 4an (1− an) (13)

Lx < xn < Ux (14)

Ly < yn < Uy (15)

where the subscript n indexes the individual turbines, the subscript k indexes the different flow directions, αk

is a weight factor for a given inflow direction, A is the turbine rotor area, Cp is the turbine power coefficient,Ct is the turbine thrust coefficient, u is the simulation flow field, and (xn, yn) is the position of turbine n.The body force from each turbine fn,AD is calculated based on the velocity at the center of the rotor and theforce is distributed across the rotor plane. The turbines are constrained to remain within a rectangular boxdescribed by the upper and lower bounds in the x and y directions given by Lx, Ly, Ux, Uy. For our initialstudy we focus on optimizing J by finding optimal turbine positions and axial induction factors representedby the control vector m = [x,y,a] for n = 1 . . . N .

B. Governing Equations

WindSE is being developed to have laminar, RANS, and LES flow solver capabilities. Our initial resultsare for low Reynolds number steady state laminar flow conditions. The governing equations in this caseare simply the incompressible steady state Navier Stokes equations with an O (100) Reynolds number anda body force imparted by the turbine actuator disk. This is given by Eqs. (9) and (10) with ui the laminarvelocity and ν the kinematic viscosity. The viscosity in the implemented form of the equations can be readilyaugmented with an eddy viscosity to incorporate a RANS turbulent flow solver.

C. Turbine Representation

The turbines are represented as non-rotating actuator disks using standard actuator disk theory. When notoptimizing the axial induction factor, we assume a constant axial induction factor a = 1

3 and calculate thebody force from turbine n according to Eq. (11). The turbine power production is calculated based off therotor velocity according the formulation in Eq. (8). The turbine body force is smoothed over the rotor diskin the streamwise and spanwise directions using a Gaussian distribution.

D. Initial and Boundary Conditions

Dirichlet boundary conditions are used to prescribe the velocity and wind direction on inflow boundaries.The pressure is set to 0 on outflow boundaries. A steady state solution is found for each of the K inflowdirections. Different inflow directions are achieved by applying a rotation to the turbine coordinates x andy. The rotation is incorporated into the adjoint gradient calculation and allows the same computationaldomain to be used for the entire wind rose.

6 of 16

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F C

OL

OR

AD

O B

OU

LD

ER

on

Febr

uary

16,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

2199

IV. Solution Implementation

A. WindSE Numerical Solver

The WindSE flow solver is implemented in a software package called FEniCS,49 which automates the solutionof PDEs using the finite element method. FEniCS is written in Python, and can be easily integratedwith other systems engineering tools like WISDEM. FEniCS comes with a user-friendly problem solvingenvironment with a mathematical syntax for writing the variational form of PDE problems and meshingthe finite element, and automated compilers for generating finite element basis functions and assemblingthe variational forms. The Python code is just-in-time (JIT) compiled to C++ for computational speed.FEniCS can interface to common HPC libraries such as PETSc and Trilinos for numerical linear algebra,ParMETIS and SCOTCH for domain decomposition, and MPI and OpenMP for parallelization. FEniCShas been well tested and validated on a number of computational problems in solid and fluid dynamics,eigenvalue problems, and coupled PDEs.

The variational form of finite element problems lends itself well to gradient based optimization problems.The software package dolfin-adjoint50 performs a high level automated differentiation and can derive bothdiscrete adjoint and tangent linear models of a forward problem implemented in FEniCS. The discreteadjoint and tangent linear models are important in the gradient based optimization algorithms used in dataassimilation, optimal control, and error estimation. The automatic differentiation routines in dolfin-adjointtreat each forward model as a series of equation solves and therefore derives adjoint and tangent linear modelsat a higher level of abstraction than traditional algorithmic differentiation, which treats forward models asa series of elementary instructions. This higher level of abstraction gives dolfin-adjoint greater flexibilityand automation across a wide range of PDE applications. Moreover, dolfin-adjoint can be implemented onunsteady and nonlinear PDEs, and can also be run in parallel. It can directly interface to the optimizationalgorithms in SciPy and also contains routines for checking the correctness of adjoint gradients.

We solve Eqs. (9) and (10) in a coupled fashion using a Taylor-Hood mixed finite element space. Thisinvolves first order Lagrange elements for the pressure field and second order Lagrange elements for thevelocity field and has favorable built-in stability properties. The equations are solved using a nonlinearNewton solver with Jacobians calculated using the automatic differentiation tools available in FEniCS.

B. Gradient Based Optimization Process

We use gradient-based algorithms implemented in Python’s SciPy module to find local minima using the ad-joint gradients provided by dolfin-adjoint. Specifically, the BFGS and SLSQP algorithms are used dependingon the type of constraint. Both algorithms allow for bounds on turbine locations, but SLSQP is required forimplementing constraints on spacing between turbines.

Since the optimization algorithms are gradient-based, they find local rather than global minima. Thelayout optimization problem can have many local minima, particularly with few turbines and few inflowdirections. We take a statistical approach to finding the global optimum and sample many different randomstarting positions. We note that with many inflow directions, the optimization problem actually becomesmore convex and fewer optimization iterations are required to find minima.

The optimizations were run on the NREL high performance computer Peregrine and were parallelizedusing MPI. The discrete adjoint calculation is automatically parallelized by dolfin-adjoint if the forwardmodel is run in parallel, drastically simplifying code development. The optimization typically required 50 –75 iterations to converge to an optimum for a single flow direction. There are more function evaluations thanoptimization iterations because the line search method used to determine step size at each iteration oftenrequires several function evaluations. In multi-direction optimizations the forward runs are more expensivebecause many inflow directions must be evaluated, the optimization is more convex and typically convergesin less than 30 iterations. Figure 1 shows a schematic of the multi-direction optimization workflow. Thelayout and axial induction factors are optimized over all inflow directions simultaneously in a multileveloptimization process rather than in a sequential optimization process.

7 of 16

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F C

OL

OR

AD

O B

OU

LD

ER

on

Febr

uary

16,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

2199

Figure 1: Flow chart of multi-directional optimization process.

V. Optimization Results For Single and Multiple Inflow Directions

We present results for turbine layout and axial induction factor optimization with single and multipleinflow directions. We examine different spacing constraints and find that the optimizer takes advantage offlow curvature effects to capture local speedups in the single inflow direction case. The turbines are generallyarranged in a duct that concentrates the flow on downstream turbines. As more directions are considered,the optimal layouts tend to form radially-oriented ducts with even weighting of inflow directions. Weightinga particular direction heavily recovers the single inflow direction solution. We also note that optimal layoutsare typically symmetric, providing a useful heuristic to identify whether a solution is caught in a localminimum.

A. Single Inflow Direction Optimization

In the first set of simulations, constant inflow direction and velocity are examined. The turbines are initializedto a regular grid and fixed axial induction factors are used. The optimal layouts and resulting flow fields areshown in Figure 2a and Figure 2b for 16 and 32 turbines, without spacing constraints. The flow is left toright with slip boundary conditions on the top and bottom walls. The turbines are constrained to lie within arectangular site constraint, as indicated by the black rectangle. Because the adjoint gradient technique findsgradients at a cost that is independent of the number of design variables, the 16 and 32 turbine optimizationsinvolve the same computational expense.

We also explore the effects of turbine spacing constraints by imposing 1 and 1.5 rotor diameter spacing(RD) constraints. This site and minimum turbine spacing is smaller than is typical on real world windfarms, but the results still demonstrate many interesting insights not found in linear flow model optimizationresults. Figures 3a and 3b show the optimization results for the same simulations as in the unconstrainedcase. Although the turbines cannot pack as closely together, they still form a duct shape.

The optimized layouts exhibit several noteworthy characteristics. First, the layouts appear to form ductsthat accelerate the incoming flow and channel that flow toward downstream turbine rows. Turbines arealso placed along the spanwise site constraint walls to keep the flow from diverting around the downstreamrows. A similar pattern was observed in the OpenTidalFarm results.21 Another interesting feature is thatthe downstream turbines are not pushed all of the way to the downstream site constraint. This is a differentbehavior from that typically observed in linear flow optimization results. In particular, engineering models

8 of 16

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F C

OL

OR

AD

O B

OU

LD

ER

on

Febr

uary

16,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

2199

(a) Unconstrained optimized layout for 16 turbines. (b) Unconstrained optimized layout for 32 turbines.

Figure 2: Optimization results for a single flow direction left to right. Top row shows optimized turbinelocations and flow streamlines, with the black rectangle indicating the site constraint. Bottom row showsthe steady state velocity field with red indicating high speeds and blue indicating low speeds.

(a) Optimized layout with 1.5 RD spacing. (b) Optimized layout with 1 RD spacing.

Figure 3: Optimization results for a single flow direction left to right with 32 turbines and different turbinespacing constraints. Top row shows optimized turbine locations and flow streamlines, with the black rectangleindicating the site constraint. Bottom row shows the steady state velocity field with red indicating highspeeds and blue indicating low speeds.

rely on superimposing analytical velocity deficits in turbine wakes, so the conventional wisdom is to spaceturbine rows as far apart as possible in the streamwise direction to give the wake a chance to recover. Incontrast, the WindSE simulations are able to capture the speedup effect from streamline curvature introduced

9 of 16

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F C

OL

OR

AD

O B

OU

LD

ER

on

Febr

uary

16,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

2199

by the upwind turbines that concentrates the incoming flow. The WindSE layout appears to move the mainturbine rows upstream such that the flow in the ‘duct’ is prevented from diverting laterally. If the turbineswere placed as far downstream as possible, then the flow would have the space to expand laterally and thespeedup effect would be lost.

B. Multiple Inflow Direction Optimization

In order to begin the study of multiple inflow directions, two sequential forward runs are performed withthe inflow direction changing by 180 degrees and the lateral slip boundary conditions remaining constant.Both forward simulations are performed, the power outputs are summed into the total objective function,and dolfin-adjoint derives the adjoint equations based on the combined record of equation solves in bothsimulations. The resulting optimal layout is symmetric about both the flow centerlines and a 90 degreeangle bisecting the two inflow directions, as seen in Figure 4. The layout is optimized for 32 turbineswithout spacing constraints and the site constraints are outlined with a black rectangle. Figure 5 shows thecorresponding flow fields when the wind direction is from the left (Figure 5a) and from the right (Figure 5b).

Adding complexity, we next simulate flow from perpendicular directions. First, two simulations with a90 degree change in inflow angle are considered. As seen in Figures 6a and 6b, two triangular ducts form tocreate a local speedup effect when the flow is from the north and then from the west. As the 90 degree winddirection change is subdivided into a series of smaller direction changes, the optimal layout has fewer distinctducts but is still symmetric about the center of rotation. The magnitude of the direction changes betweensequential simulations, called bins, has a strong impact on the resulting optimal layout. Depending on thesize of the site constraints, the optimal layout tends to a crescent when the wind direction bins are sufficientlysmall. With so many evenly weighted different inflow directions it is more difficult to take advantage of aduct-induced speedup and instead the optimizer simply minimizes wake losses.

(a) Equal weighting of inflow directions. (b) Westward flow is weighted 10× eastward flow.

Figure 4: Optimal layouts for 32 turbines without spacing constraints and flow from the west and east.

(a) The flow field for the opposite directions opti-mized layout when the flow is left to right.

(b) The flow field for the opposite directions opti-mized layout when the flow is right to left.

Figure 5: Flow fields generated by optimizing the layout over two equally weighted but 180 degrees differentinflow directions.

10 of 16

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F C

OL

OR

AD

O B

OU

LD

ER

on

Febr

uary

16,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

2199

(a) The flow field for the perpendicular directionsoptimized layout when the flow is left to right.

(b) The flow field for the perpendicular directionsoptimized layout when the flow is top to bottom.

Figure 6: Flow fields generated by optimizing the layout over two equally weighted but 90 degrees differentinflow directions.

(a) π/2 bins. (b) π/4 bins. (c) π/6 bins.

(d) π/8 bins. (e) π/10 bins.

Figure 7: Optimal layouts with inflow ranging from the west to the north with varying bin direction sizes.Site constraints are marked with the black square.

11 of 16

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F C

OL

OR

AD

O B

OU

LD

ER

on

Febr

uary

16,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

2199

C. Full Wind Rose and Axial Induction Factor Optimization

Finally, we study the optimization of turbine locations and axial induction factors using a full real world windrose. We use the same wind rose from the Princess Amalia wind farm as the study by Fleming et al.14 Thedirection weights αk used in the objective function are based on the 8 m/s wind rose. We optimize layoutswith 6 and 12 turbines constrained to stay within a square site boundary in the center of the computationaldomain. The wind rose, optimal layouts, and resulting flow fields for a west wind are shown in Figures 8and 9.

As the wind rose is refined into smaller wind direction bins the optimal layout changes to better takeadvantage of a few select directions that are statistically more likely. No spacing constraints were enforced inthe 12 turbine optimization, and as a result WindSE placed multiple turbines close together in the northwestcorner of the site. The optimal axial induction factors are also slightly lower than 1/3 which is the optimalvalue for a turbine in isolation. This lower axial induction factor has the added benefit of reducing loads onthe turbines.

(a) Wind rose with 6 bins. (b) Wind rose with 12 bins. (c) Wind rose with 24 bins.

(d) Optimal layout for 6 bins. (e) Optimal layout for 12 bins. (f) Optimal layout for 24 bins.

(g) Flow field with a west wind. (h) Flow field with a west wind. (i) Flow field with a west wind.

Figure 8: Layout optimization and flow field results for 6 turbines with the Princess Amalia wind rose.

12 of 16

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F C

OL

OR

AD

O B

OU

LD

ER

on

Febr

uary

16,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

2199

(a) Wind rose with 6 bins. (b) Wind rose with 12 bins. (c) Wind rose with 24 bins.

(d) Optimal layout for 6 bins. (e) Optimal layout for 12 bins. (f) Optimal layout for 24 bins.

(g) Flow field with a west wind. (h) Flow field with a west wind. (i) Flow field with a west wind.

Figure 9: Layout optimization and flow field results for 12 turbines with the Princess Amalia wind rose.Without spacing constraints, WindSE places multiple turbines in the northwest corner of the site.

13 of 16

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F C

OL

OR

AD

O B

OU

LD

ER

on

Febr

uary

16,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

2199

VI. Summary and Conclusions

We have introduced a new wind farm flow model and corresponding adjoint model for wind turbine layoutand axial induction factor optimization. This enables efficient gradient-based optimization of wind turbinelayouts for farm-wide systems engineering analysis. We find discrete adjoints of our objective functionalusing the Python packages FEniCS and dolfin-adjoint, and use gradient-based optimization algorithms tofind optimal layouts. The adjoint approach finds gradients at a cost that is independent of the number ofdesign variables and facilitates the use of high fidelity flow models rather than simplified engineering models.

Our initial results for a single inflow direction demonstrate that the optimizer is able to take advantageof flow curvature and speedup effects when a dominant wind direction is present. These curvature effectsare not represented by the analytical wake deficits used in common industry flow models. As a result, ouroptimal layouts reflect very different strategies than are common in the wind industry. Instead of spacingparallel turbine rows as far apart as possible in the streamwise direction, WindSE channels the flow towardthe center of the farm and creates a local speedup. Turbines are then arranged to prevent the flow fromdiverting around downwind rows by moving them closer to the upwind speedup zone.

In multiple direction studies, the duct strategy is less effective for uniformly distributed wind roses. Theoptimal layout is tends to be symmetric about the center of rotation for optimization studies with just afew inflow directions. However, refining the wind rose binning process does affect the optimization results.Unevenly weighting the inflow directions allows the flow to recover a directional duct shape, and also reflectsreal world wind roses which often exhibit a preferred direction. As more evenly weighted direction bins areintroduced, the directionality of a duct is less optimal and the optimizer focuses on a layout topology thatminimizes wake losses. Finally, optimal axial induction factors tend to be below the optimal value for asingle turbine in isolation, suggesting that turbine-level controls may be useful for power optimization andalso benefit from a reduction in loads.

WindSE extends the wind farm systems engineering analysis in the NREL WISDEM toolset to includehigh fidelity layout and controls optimization. Ongoing work to continue development of WindSE is focusedon adding greater complexity and accuracy to the flow solver and extending the simulations to three dimen-sions. Additional work is focused on improving the fidelity of the turbine representation to allow for morecontrol optimization.

References

1Jay P. Goit and Johan Meyers. Optimal control of energy extraction in wind-farm boundary layers. 768:5–50.2Richard J.A.M. Stevens, Dennice F. Gayme, and Charles Meneveau. Effects of turbine spacing on the power output of

extended wind-farms. Wind Energy, pages n/a–n/a, 2015.3Katherine Dykes, Andrew Ning, Ryan King, Peter Graf, George Scott, and Paul S. Veers. Sensitivity Analysis of Wind

Plant Performance to Key Turbine Design Parameters: A Systems Engineering Approach. American Institute of Aeronauticsand Astronautics, January 2014.

4P. S. Jackson and J. C. R. Hunt. Turbulent wind flow over a low hill. Quarterly Journal of the Royal MeteorologicalSociety, 101(430):929–955, 1975.

5J. L. Walmsley, P. A. Taylor, and T. Keith. A simple model of neutrally stratified boundary-layer flow over complexterrain with surface roughness modulations (MS3djh/3r). Boundary-Layer Meteorology, 36(1):157–186, July 1986.

6N O Jensen. A note on wind generator interaction. RisM2411 Ris National Laboratory Roskilde, (RIS-M-2411):1–16,1983.

7J.F. Ainslie. Calculating the flowfield in the wake of wind turbines. Journal of Wind Engineering and IndustrialAerodynamics, 27(13):213–224, January 1988.

8Andrew Kusiak and Zhe Song. Design of wind farm layout for maximum wind energy capture. 35(3):685–694.9Javier Serrano Gonzlez, Angel G. Gonzalez Rodriguez, Jos Castro Mora, Jess Riquelme Santos, and Manuel Burgos

Payan. Optimization of wind farm turbines layout using an evolutive algorithm. 35(8):1671–1681.10Wing Yin Kwong, Peter Y. Zhang, David Romero, Joaquin Moran, Michael Morgenroth, and Cristina Amon. Wind

farm layout optimization considering energy generation and noise propagation. In ASME 2012 International Design Engineer-ing Technical Conferences and Computers and Information in Engineering Conference, pages 323–332. American Society ofMechanical Engineers.

11Souma Chowdhury, Jie Zhang, Achille Messac, and Luciano Castillo. Unrestricted wind farm layout optimization(UWFLO): Investigating key factors influencing the maximum power generation. 38(1):16–30.

12Bryony L. Du Pont and Jonathan Cagan. An extended pattern search approach to wind farm layout optimization.134(8):081002.

13J.R. Marden, S.D. Ruben, and L.Y. Pao. A model-free approach to wind farm control using game theoretic methods.21(4):1207–1214.

14 of 16

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F C

OL

OR

AD

O B

OU

LD

ER

on

Febr

uary

16,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

2199

14Paul A. Fleming, Andrew Ning, Pieter M. O. Gebraad, and Katherine Dykes. Wind plant system engineering throughoptimization of layout and yaw control: Wind plant system engineering. Wind Energy, 19(2):329–344, February 2016.

15Jos Herbert-Acero, Oliver Probst, Pierre-Elouan Rthor, Gunner Larsen, and Krystel Castillo-Villar. A review of method-ological approaches for the design and optimization of wind farms. 7(11):6930–7016.

16Gunner Chr Larsen, Helge Aagaard Madsen, Niels Troldborg, Torben J. Larsen, Pierre-Elouan Rthor, Peter Fuglsang,Sren Ott, Jakob Mann, Thomas Buhl, Morten Nielsen, and others. TOPFARM-next generation design tool for optimisation ofwind farm topology and operation.

17Marc Calaf, Charles Meneveau, and Johan Meyers. Large eddy simulation study of fully developed wind-turbine arrayboundary layers. Physics of Fluids, 22(1):015110–015110–16, January 2010.

18Fernando Port-Agel, Yu-Ting Wu, Hao Lu, and Robert J. Conzemius. Large-eddy simulation of atmospheric boundarylayer flow through wind turbines and wind farms. Journal of Wind Engineering and Industrial Aerodynamics, 99(4):154–168,April 2011.

19Matthew J. Churchfield, Sang Lee, John Michalakes, and Patrick J. Moriarty. A numerical study of the effects ofatmospheric and wake turbulence on wind turbine dynamics. Journal of Turbulence, page N14, 2012.

20Johan Meyers and Charles Meneveau. Optimal turbine spacing in fully developed wind farm boundary layers. WindEnergy, 15(2):305–317, 2012.

21S.W. Funke, P.E. Farrell, and M.D. Piggott. Tidal turbine array optimisation using the adjoint approach. RenewableEnergy, 63:658–673, March 2014.

22J.-L Lions. Optimal control of systems governed by partial differential equations. Springer-Verlag, Berlin; New York,1971.

23Olivier Pironneau. On optimum design in fluid mechanics. Journal of Fluid Mechanics, 64(01):97–110, 1974.24Antony Jameson. Aerodynamic design via control theory. Journal of Scientific Computing, 3(3):233–260, 1988.25Antony Jameson. Optimum aerodynamic design using CFD and control theory. American Institute of Aeronautics and

Astronautics, June 1995.26Antony Jameson, Juan J. Alonso, James J. Reuther, Luigi Martinelli, and John C. Vassberg. Aerodynamic shape

optimization techniques based on control theory. AIAA, 1998.27Antony Jameson, L. Martinelli, and N. A. Pierce. Optimum aerodynamic design using the NavierStokes equations.

Theoretical and Computational Fluid Dynamics, 10(1-4):213–237, 1998.28Michael B. Giles and Niles A. Pierce. Adjoint equations in CFD: duality, boundary conditions and solution behaviour.

AIAA paper, 97:1850, 1997.29Michael B. Giles and Niles A. Pierce. On the properties of solutions of the adjoint Euler equations. Numerical Methods

for Fluid Dynamics VI. ICFD, pages 1–16, 1998.30Michael B. Giles and Niles A. Pierce. Analytic adjoint solutions for the quasi-one-dimensional Euler equations. Journal

of Fluid Mechanics, 426(2001):327–345, 2001.31M. B. Giles. Discrete Adjoint Approximations with Shocks. In Thomas Y. Hou and Eitan Tadmor, editors, Hyperbolic

Problems: Theory, Numerics, Applications, pages 185–194. Springer Berlin Heidelberg, January 2003.32Antony Jameson. Aerodynamic shape optimization using the adjoint method. Lectures at the Von Karman Institute,

Brussels, 2003.33Michael B. Giles and Niles A. Pierce. An introduction to the adjoint approach to design. Flow, turbulence and combustion,

65(3-4):393–415, 2000.34Kyriakos C. Giannakoglou and Dimitrios I. Papadimitriou. Adjoint methods for shape optimization. In Optimization

and computational fluid dynamics, pages 79–108. Springer, 2008.35Alik Ismail-Zadeh, Gerald Schubert, Igor Tsepelev, and Alexander Korotkii. Inverse problem of thermal convection:

numerical approach and application to mantle plume restoration. Physics of the Earth and Planetary Interiors, 145(1-4):99–114, July 2004.

36Philippe Courtier, John Derber, Ron Errico, Jean-Francois Louis, and Tomislava Vukievi. Important literature on theuse of adjoint, variational methods and the Kalman filter in meteorology. Tellus A, 45(5):342–357, 1993.

37Ronald M. Errico. What Is an Adjoint Model? Bulletin of the American Meteorological Society, 78(11):2577–2591,November 1997.

38Krzysztof J. Fidkowski and David L. Darmofal. Review of Output-Based Error Estimation and Mesh Adaptation inComputational Fluid Dynamics. AIAA Journal, 49(4):673–694, April 2011.

39Michael B. Giles. On adjoint equations for error analysis and optimal grid adaptation in CFD. Frontiers of computationalfluid dynamics, pages 155–169, 1998.

40Karthikeyan Duraisamy, J. J. Alonso, Francisco Palacios, and Praveen Chandrashekar. Error Estimation for High SpeedFlows Using Continuous and Discrete Adjoints. In 48th AIAA Aerospace Sciences Meeting, pages 2010–128, 2010.

41Karthikeyan Duraisamy and Juan Alonso. Adjoint Based Techniques for Uncertainty Quantification in Turbulent Flowswith Combustion. American Institute of Aeronautics and Astronautics, June 2012.

42J. E. Hicken and J. J. Alonso. PDE-constrained optimization with error estimation and control. Journal of ComputationalPhysics.

43Karthik Mani and Dimitri J. Mavriplis. Error estimation and adaptation for functional outputs in time-dependent flowproblems. Journal of Computational Physics, 229(2):415–440, January 2010.

44JRRA Martins and John T. Hwang. Review and Unification of Methods for Computing Derivatives of MultidisciplinarySystems. In Proceedings of the 53rd AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Confer-ence, 2012.

45Joaquim R. R. A. Martins and Andrew B. Lambe. Multidisciplinary Design Optimization: A Survey of Architectures.AIAA Journal, 51(9):2049–2075, September 2013.

15 of 16

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F C

OL

OR

AD

O B

OU

LD

ER

on

Febr

uary

16,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

2199

46P. J. Blonigan, R. Chen, Q. Wang, and J. Larsson. Towards adjoint sensitivity analysis of statistics in turbulent flowsimulation. In Proceedings of the Summer Program, page 229, 2012.

47Qiqi Wang. Forward and adjoint sensitivity computation of chaotic dynamical systems. Journal of ComputationalPhysics, 2012.

48Qiqi Wang. Progress and Challenges in Sensitivity Computation of Chaotic Fluid Flow Simulations. American Instituteof Aeronautics and Astronautics, January 2013.

49Anders Logg, Kent-Andre Mardal, and Garth Wells, editors. Automated Solution of Differential Equations by the FiniteElement Method, volume 84 of Lecture Notes in Computational Science and Engineering. Springer Berlin Heidelberg, Berlin,Heidelberg, 2012. FEniCS.

50P. E. Farrell, D. A. Ham, S. W. Funke, and M. E. Rognes. Automated Derivation of the Adjoint of High-Level TransientFinite Element Programs. SIAM Journal on Scientific Computing, 35(4):C369–C393, January 2013.

16 of 16

American Institute of Aeronautics and Astronautics

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F C

OL

OR

AD

O B

OU

LD

ER

on

Febr

uary

16,

201

6 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/6.2

016-

2199


Recommended