multi-fidelity sparse polynomial chaos surrogate models ...multi-fidelity sparse polynomial chaos...

16
Multi-Fidelity Sparse Polynomial Chaos Surrogate Models for Flutter Database Generation Markus P. Rumpfkeil * University of Dayton, Ohio, 45469, USA Dean E. Bryson and Philip S. Beran Air Force Research Laboratory, Wright-Patterson Air Force Base, Ohio, 45433, USA In this paper, multi-fidelity sparse polynomial chaos expansion (MFSPCE) models of critical flutter dynamic pressures as a function of Mach number, angle of attack, and thick- ness to chord ratio are constructed in lieu of solely using computationally expensive high- fidelity engineering analyses. Compressed sensing is used to reconstruct the sparse rep- resentation, and an all-at-once approach is employed to generate multi-fidelity polynomial chaos expansions with hybrid additive-multiplicative bridge functions of arbitrary order. Once an accurate MFSPCE is constructed, it can be used for evaluating a large number of designs for design space exploration or for uncertainty quantification. To demonstrate that accurate MFSPCE models can be obtained at lower computational cost than high-fidelity ones, analytic test functions are used. As a more complex application example, the well known AGARD 445.6 dynamic aeroelastic test case model is employed. The high and low fidelity levels considered are Euler and panel solutions, respectively, all combined with a modal structural solver. I. Motivation and Background Aeroelastic flutter can be a dangerous phenomenon encountered in flexible structures subjected to aero- dynamic forces such as aircraft, buildings, and bridges. Flutter occurs as a result of interactions between aerodynamics, stiffness, and inertial forces on a structure caused by positive feedback between the body’s deflection and the force exerted by the fluid flow. The implicit assumption for aircraft flutter is that the most critical case at any given Mach number will occur at sea level conditions since the dynamic pressure is highest there. However, Bendiksen 1, 2 presents a counterexample involving a generic swept wing representative for a transport aircraft (called G¨ ottingen or G-wing) in which transonic limit cycle flutter occurs at altitude rather than at sea level. Often, before detecting such counter-intuitive behaviors, several design stages have already been completed, or – in a worst-case scenario – only flight testing will reveal them leading to massive cost and schedule overruns. Unfortunately, transonic aeroelastic experiments are extremely expensive and there are only a few avail- able in the public domain for validation purposes. One set are the tests performed in the Langley transonic dynamics tunnel in the early 1960s known as the AGARD 445.6 dynamic aeroelastic test cases. 3 A series of subsonic and transonic flutter data were obtained on different wing models in both air and freon-12. One of these models was denoted “weakened 3” and was made of symmetric NACA 65A004 airfoils with a sweep angle of 45 at the quarter chord line, a semi-span of 0.762 m, and a taper ratio of 0.66. The uncambered model was rigidly mounted on the tunnel wall at zero angle of attack, thus eliminating any static aeroelastic deformation. In this paper, the experiments on the “weakened 3” model conducted in air are simulated, and the design space is extended by changing the angle of attack as well as the thickness to chord ratio. Computational flutter simulations tend to be very expensive as well since they require unsteady coupled aeroelastic analyses. In order to save computational time while trying to maintain high-fidelity information * Associate Professor, Hans von Ohain Endowed Chair, Dept. of Mechanical and Aerospace Eng., Associate Fellow AIAA, [email protected] Research Aerospace Engineer, Senior Member AIAA Principal Research Aerospace Engineer, Associate Fellow AIAA Cleared for Public Release, Distribution Unlimited: 88ABW-2018-5426 1

Upload: others

Post on 23-Oct-2020

5 views

Category:

Documents


0 download

TRANSCRIPT

  • Multi-Fidelity Sparse Polynomial Chaos Surrogate

    Models for Flutter Database Generation

    Markus P. Rumpfkeil ∗

    University of Dayton, Ohio, 45469, USA

    Dean E. Bryson†and Philip S. Beran ‡

    Air Force Research Laboratory, Wright-Patterson Air Force Base, Ohio, 45433, USA

    In this paper, multi-fidelity sparse polynomial chaos expansion (MFSPCE) models ofcritical flutter dynamic pressures as a function of Mach number, angle of attack, and thick-ness to chord ratio are constructed in lieu of solely using computationally expensive high-fidelity engineering analyses. Compressed sensing is used to reconstruct the sparse rep-resentation, and an all-at-once approach is employed to generate multi-fidelity polynomialchaos expansions with hybrid additive-multiplicative bridge functions of arbitrary order.Once an accurate MFSPCE is constructed, it can be used for evaluating a large number ofdesigns for design space exploration or for uncertainty quantification. To demonstrate thataccurate MFSPCE models can be obtained at lower computational cost than high-fidelityones, analytic test functions are used. As a more complex application example, the wellknown AGARD 445.6 dynamic aeroelastic test case model is employed. The high and lowfidelity levels considered are Euler and panel solutions, respectively, all combined with amodal structural solver.

    I. Motivation and Background

    Aeroelastic flutter can be a dangerous phenomenon encountered in flexible structures subjected to aero-dynamic forces such as aircraft, buildings, and bridges. Flutter occurs as a result of interactions betweenaerodynamics, stiffness, and inertial forces on a structure caused by positive feedback between the body’sdeflection and the force exerted by the fluid flow. The implicit assumption for aircraft flutter is that the mostcritical case at any given Mach number will occur at sea level conditions since the dynamic pressure is highestthere. However, Bendiksen1,2 presents a counterexample involving a generic swept wing representative fora transport aircraft (called Göttingen or G-wing) in which transonic limit cycle flutter occurs at altituderather than at sea level. Often, before detecting such counter-intuitive behaviors, several design stages havealready been completed, or – in a worst-case scenario – only flight testing will reveal them leading to massivecost and schedule overruns.

    Unfortunately, transonic aeroelastic experiments are extremely expensive and there are only a few avail-able in the public domain for validation purposes. One set are the tests performed in the Langley transonicdynamics tunnel in the early 1960s known as the AGARD 445.6 dynamic aeroelastic test cases.3 A series ofsubsonic and transonic flutter data were obtained on different wing models in both air and freon-12. One ofthese models was denoted “weakened 3” and was made of symmetric NACA 65A004 airfoils with a sweepangle of 45◦ at the quarter chord line, a semi-span of 0.762m, and a taper ratio of 0.66. The uncamberedmodel was rigidly mounted on the tunnel wall at zero angle of attack, thus eliminating any static aeroelasticdeformation. In this paper, the experiments on the “weakened 3” model conducted in air are simulated, andthe design space is extended by changing the angle of attack as well as the thickness to chord ratio.

    Computational flutter simulations tend to be very expensive as well since they require unsteady coupledaeroelastic analyses. In order to save computational time while trying to maintain high-fidelity information

    ∗Associate Professor, Hans von Ohain Endowed Chair, Dept. of Mechanical and Aerospace Eng., Associate Fellow AIAA,[email protected]†Research Aerospace Engineer, Senior Member AIAA‡Principal Research Aerospace Engineer, Associate Fellow AIAACleared for Public Release, Distribution Unlimited: 88ABW-2018-5426

    1

  • the construction of highly accurate global surrogate models such as polynomial chaos expansions (PCEs)4–6

    is very attractive. Gradient enhanced PCE4–8 models have been showing great promise as well; however, dueto the unavailability of gradient information in some of the employed analysis codes, they are not employedhere. PCEs also support the usage of multi-fidelity training data.8–12 The general idea is to combine trendsfrom inexpensive, low-fidelity (LF) data (e.g., coarser meshes, less sophisticated models) with a fitting ofthe high-fidelity (HF) data (e.g., finer meshes, better models, experimental data). Thus, the trend of theunknown function underlying the intensively sampled LF data needs to be mapped to the usually lessintensively sampled HF data, for example, via the popular correction-based method.13 The correction iscalled bridge function, scaling function, or calibration, and can be multiplicative,14 additive15,16 or hybridmultiplicative/additive.8–10,12

    Lastly, significant efforts have been made to develop efficient methodologies to reduce the curse of dimen-sionality. The compressed sensing theory17–20 has shown great potential for PCEs depending on the sparsityof the underlying solution or the decay rate of the PCE coefficients. For instance, it has been shown thatusing L1-minimization in the PCE method is very effective when the number of non-zero coefficients in thePCE are few.20 Compressed sensing is a developing research topic in the field of signal processing, aimingat reconstructing the sparse signals using many fewer samples than are required by the Shannon-Nyquistsampling principle.19

    In summary, the goal of this work is to perform aeroelastic flutter simulations of an AGARD 445.6like geometry at multiple fidelity levels, and then to fuse this multi-fidelity information to give an accuraterepresentation of the underlying design space at a small overall computational cost. Section II outlines theconstruction process of multi-fidelity sparse polynomial chaos expansion (MFSPCE) models, and Section IIIshows applications of MFSPCEs to analytic test functions. Section IV describes the computational multi-fidelity aeroelastic flutter analyses in more detail, while Section V shows results of employing MFSPCEs tothe generation of flutter databases. Finally, Section VI concludes this paper.

    II. All-at-Once Multi-fidelity Sparse Polynomial Chaos Expansions

    Usually, multi-fidelity PCE implementations take a multi-step approach, separating the estimation ofcorrective functions from the fitting of the model of interest.21–23 Typically, corrections are of the additiveor multiplicative forms, and one must choose a priori what form the correction should take or estimate ablend of both forms. The all-at-once approach developed by Bryson and Rumpfkeil,8,12 which is employedhere, differs in that only a single fitting is performed, simultaneously determining additive and multiplicativecorrections to the low-fidelity data to best approximate the high-fidelity function in a least-squares sense. Theall-at-once approach can also be augmented with the addition of gradient and Hessian information for bothhigh- and low-fidelity data if it is available. A brief review of the approach is given next for completeness.

    In a traditional PCE, a high-fidelity function is approximated by the series

    fh (x) =

    P∑k=0

    βkΨk (x) + �h (x) . (1)

    where �h (x) is the error relative to the true function and Ψk (x) are the selected multidimensional basesformed by

    Ψk (x) =

    D∏i=1

    Ψaki (xi) . (2)

    The multi-index aki indicates the order of the one-dimensional polynomial in the ith direction for the kth

    term in the total expansion.24 The number of terms, P + 1, in the expansion is a combination of the designspace dimensionality, D, and the chosen polynomial order, p,

    P + 1 =

    (D + p

    p

    ). (3)

    The coefficients βk may be calculated non-intrusively by the point collocation method (as is done here) orby the Galerkin spectral projection method.21

    For the inclusion of multifidelity data in the model fitting, it is assumed that the high-fidelity function is

    2

  • matched by multiplicative and additive corrections to the low-fidelity model,

    fh (x) = fl (x) + α (x) fl (x) + δ (x) . (4)

    where the corrective terms themselves can be represented as PCEs as well,

    α (x) =

    Q∑k=0

    αkΨk (x) + �α (x) , (5)

    and

    δ (x) =

    R∑k=0

    δkΨk (x) + �δ (x) . (6)

    As indicated by the limits of summation (Q and R) the corrections need not be of the same polynomialorder. The treatment of the high- and low-fidelity data in equation (4) during the determination of themodel coefficients depends on whether one or both function fidelities are evaluated at a particular trainingpoint. If only the LF function is evaluated, the PCE estimate of the high-fidelity value is introduced intoequation (4), which yields

    fl (x) =

    P∑k=0

    βkΨk (x)− fl (x)Q∑k=0

    αkΨk (x)−R∑k=0

    δkΨk (x) + �hl (x) , (7)

    where �hl (x) represents the combined total error introduced by all three PCEs. If both fidelities are evaluated,the true function values participate in finding the best calibration between the two data sets,

    fl (x)− fh (x) = −fl (x)Q∑k=0

    αkΨk (x)−R∑k=0

    δkΨk (x) + �l (x) , (8)

    where �l (x) denotes the combined error of the additive and multiplicative corrections. The two fl (x) termscould equivalently be lumped together into a single term multiplied by α (x) + 1, but they are kept separatehere for consistency with standard hybrid approaches in the literature. Because a PCE of the LF data itselfis not sought, the values taken on by the function at the training points may be included in both sides of themathematical system without introducing inconsistencies. This fact enables the additive and multiplicativecorrections to be determined simultaneously in a single regression system. In total, there are P +Q+R+ 3unknown coefficients to be approximated (βk, αk, and δk). Equations (7) and (8) have the coefficients αk andδk in common, and equations (1) and (7) share the coefficients βk. Following Boopathy and Rumpfkeil

    7 aswell as Li et al.,25 gradient and Hessian information may be included by taking derivatives of the multifidelityPCE system given by equations (1), (7), and (8).

    Estimated values of the PCE coefficients (β̂k, α̂k, and δ̂k) are determined by singular value decomposition,which minimizes the errors relative to the training data in a least-squares sense and is relatively robust toill-conditioned systems. The new approximate functions using the estimated coefficients are denoted β̂ (x),

    α̂ (x), and δ̂ (x). To concisely express the regression matrix, one can define the block matrices

    ΦP (x) = [Ψ0 (x) . . .ΨP (x)] fl (x) ΦP (x) = [fl (x) Ψ0 (x) . . . fl (x) ΨP (x)] (9)

    Here, only the system of equations with function values alone is considered and the entire regression system

    3

  • Aβ̂ = b can be written as

    ΦP(x(1)h

    )0 0

    ...

    ΦP(x(Nh)h

    )0 0

    ΦP(x(1)l

    )−fl

    (x(1)l

    )ΦQ

    (x(1)l

    )−ΦR

    (x(1)l

    )...

    ΦP(x(Nl)l

    )−fl

    (x(Nl)l

    )ΦQ

    (x(Nl)l

    )−ΦR

    (x(Nl)l

    )0 −fl

    (x(1)hl

    )ΦQ

    (x(1)hl

    )−ΦR

    (x(1)hl

    )...

    0 −fl(x(Nhl)hl

    )ΦQ

    (x(Nhl)hl

    )−ΦR

    (x(Nhl)hl

    )

    β̂0...

    β̂P

    α̂0...

    α̂Q

    δ̂0...

    δ̂R

    =

    fh

    (x(1)h

    )...

    fh

    (x(Nh)h

    )fl

    (x(1)l

    )...

    fl

    (x(Nl)l

    )fl

    (x(1)hl

    )− fh

    (x(1)hl

    )...

    fl

    (x(Nhl)hl

    )− fh

    (x(Nhl)hl

    )

    .

    (10)The first Nh rows of the system correspond to equation (1) evaluated at the training points with onlyhigh-fidelity data. The next Nl rows correspond to equation (7) evaluated at the training points were thelow-fidelity function alone is calculated. The last Nhl rows correspond to equation (8) evaluated at trainingpoints where both the high- and low-fidelity functions are calculated.

    II.A. Training Point Selection

    Using a regression procedure to solve for the PCE coefficients via non-intrusive point collocation, the selectionof training points is decoupled from the underlying basis functions. However, thoughtful selection may helpto improve the convergence of the surrogate model. Training points and basis functions of the PCE may beselected to correspond to the underlying probability distributions of the input parameters. Here, the designparameters are assumed to be uniformly distributed, corresponding to Legendre basis polynomials. Thetraining points are selected to be Gauss-Patterson knots and each grid level is a subset of the next higherlevel grid. This nested property is desired to enable reusing data in an adaptive process. Structured grids maybe built upon tensor products of one-dimensional abscissae, but the number of points grows exponentiallywith the number of dimensions and grid level. To ameliorate the cost of evaluating the responses on sucha mesh, Smolyak sparse grids with a slow-exponential growth rule are generated using Burkardt’s sparsegrid mixed growth anisotropic rules library26 and the numbers of resulting training points for several designspace dimensions and grid levels are tabulated in Table 1. In the multi-fidelity cases, a low-fidelity grid (ofwhich the high-fidelity grid is fully a subset) is specified to be at least one grid level higher.

    Table 1. Number of points for several grid levels and dimensions of slow-growth, sparse, Gauss-Patterson grids.

    Grid Dimension, D

    Level 2 3 4 5 6 10 20 40

    0 1 1 1 1 1 1 1 1

    1 5 7 9 11 13 21 41 81

    2 9 19 33 51 73 201 801 3201

    3 17 39 81 151 257 1201 10001 82401

    4 33 87 193 391 737 5281 90561 1557121

    5 33 135 385 903 1889 19105 641409 huge

    6 65 207 641 1743 4161 60225 huge huge

    7 97 399 1217 3343 8481 169185 huge huge

    II.B. Compressed Sensing

    Compressed sensing (CS), also known as compressive sampling, is a growing subject in the fields of imageor signal processing, applied mathematics, computer science and statistics.17–20 CS offers a framework to

    4

  • accurately, or even exactly, recover a sparse signal from a set of incomplete observations. The method canefficiently reconstruct a sparse signal such that the number of required measurements are much lower thanthe cardinality of the signal. In other words, CS aims at selecting a small number of basis polynomials withgreat impact on the model response.27

    There exists a wide variety of methods for sparse recovery of a signal β̂ from a set of incomplete mea-surements. Knowing (or hoping) that the original signal is sparse, and having obtained the measurement

    vector b, one can attempt to recover β̂ by solving an optimization problem of the form

    minβ̂

    0.5||Aβ̂ − b||22 + λ||β̂||1 + 0.5λ2||β̂||22 (11)

    A greedy solution can be found to equation (11) using a least angle regression (LARS) algorithm. TheLARS algorithm is similar to an orthogonal matching pursuit (OMP). LARS maintains an active set ofcolumns and builds this set by adding the column with the largest correlation with the residual to thecurrent residual. However, unlike OMP, LARS solves a penalized least squares problem, thus taking a stepalong an equiangular direction at each step, that is, a direction having equal angles with the vectors in theactive set. LARS and OMP do not allow a column (PCE basis) to leave the active set. However if thisrestriction is removed from LARS the resulting algorithm LASSO (least absolute shrinkage and selectionoperator) can provably solve equation (11).

    The elastic net algorithm was developed to overcome some of the limitations of the LASSO formu-lation. Specifically, if the Ntot × Ncoef Vandermonde matrix is over-determined (Ncoef < Ntot) whereNcoef = P +Q+R+ 3 and Ntot = Nh +Nl +Nhl, the LASSO selects at most Ncoef variables before itsaturates, due to the nature of the convex optimization problem. Also, if there is a group of variables amongwhich the pairwise correlations are very high, then the LASSO tends to select only one arbitrary one. If thereare high correlations between predictors, it has been empirically observed that the prediction performanceof the LASSO is dominated by ridge regression. However, it is hard to estimate the λ2 penalty in practice,and the aforementioned issues typically do not arise very often, thus the LASSO algorithm is used in thiswork. In particular, the library mlpack28 is employed and LASSO is activated if one sets λ > 0 and λ2 = 0in equation (11).

    It must be stressed that selecting a proper value for λ is very important for an accurate computation ofthe sparse PCE coefficients. If λ is too large, then the reconstructed signal may not be accurate enough,while very small values of λ may result in over-fitting or a less sparse solution. Following Salehi et al.,27 eithera true root-mean-square error (RMSE) or the leave-one-out cross-validation error29 is used to determine an

    optimal λ̂. In order to find it, the LASSO is solved for multiple values of λ, and the associated error iscalculated for each case. Then, λ̂ is simply set to the value of λ for which the error is the smallest.

    III. Analytic Test Functions

    To test MFSPCE framework described in the previous section, two analytic functions are employed herein several different dimensions over the domain −2 ≤ xi ≤ 2 for i = 1, . . . , D. The first case is the well-knownRosenbrock function,

    fh (x) =

    D−1∑i=1

    (1− xi)2 + 100(xi+1 − x2i

    )2. (12)

    Globally it is easy to approximate the overall quartic shape of this function, but locally it is difficult to capturethe gently sloped valleys leading to a global minimum. The second function is a sum of exponential-sineproducts, which is not a polynomial and thus cannot be fully represented by PCEs,

    gh (x) =1

    D

    D∑i=1

    exp (−0.1xi) sin(π

    2xi

    ). (13)

    The low-fidelity version of the Rosenbrock function is a transformation of the HF function value by linear

    5

  • additive and multiplicative factors,

    fl (x) =fh (x)− 4.0−

    ∑Di=1 0.5xi

    10.0 +∑Di=1 0.25xi

    . (14)

    Thus, linear hybrid corrections to the LF data will fully recover the HF function. For the sum of exponential-sine products a low-fidelity expression is taken to be

    gl (x) =1

    D

    D∑i=1

    sin(π

    2xi

    ), (15)

    which mimics the global behavior of the HF function but lacks some detail.

    III.A. Results for Analytic Functions

    In this section, the root-mean-square error (RMSE) is used for comparison purposes and is computed onequispaced Cartesian meshes using 1012 = 10, 201 nodes in 2D, 86 = 262, 144 nodes in 6D, and 310 = 59, 049nodes in 10D. Note that lines are plotted between markers in the result figures as a visual aide, but do notnecessarily reflect performance at intermediate points.

    Rosenbrock Function

    The Rosenbrock function serves as a verification case for the correct implementation of the MFSPCE methodand results are provided in Figure 1 for two and ten dimensions.

    10-8

    10-6

    10-4

    10-2

    100

    102

    5 9 17 33

    RM

    SE

    Number of HF training points

    HFMF

    HFSMFS

    (a) Two dimensions

    10-6

    10-4

    10-2

    100

    102

    104

    21 201 1201 5281

    RM

    SE

    Number of HF training points

    HFMF

    HFSMFS

    (b) Ten dimensions

    0

    100

    200

    300

    400

    500

    600

    700

    5 9 17

    RM

    SE

    Number of HF training points

    HFMF

    HFSMFS

    (c) Two dimensions (zoom in)

    0

    2000

    4000

    6000

    8000

    10000

    12000

    21 201 1201

    RM

    SE

    Number of HF training points

    HFMF

    HFSMFS

    (d) Ten dimensions (zoom in)

    Figure 1. Verification of multifidelity PCE implementation on Rosenbrock function in two and ten dimensions.

    6

  • Once sufficient training data are gathered to construct a fourth-order polynomial expansion, the sur-rogate model should match the truth function exactly with the error dropping accordingly. Imposing anoversampling ratio of two, the standard (full order) PCE surrogate model with high-fidelity training dataalone (HF) requires 30 training points to match the exact function in two dimensions and 2002 points in tendimensions. For the multi-fidelity (MF) models some of these points may be LF. The number of LF trainingpoints in two dimensions is four levels higher (ie. when using 17 HF points, 97 LF points are used) and inten dimensions it is one level higher (e.g., 201 HF are combined with 1201 LF points). The shown resultsare in agreement with the fundamental expectations, demonstrating the correctness of the implementation.

    For the sparse polynomial chaos expansion models, a fourth-order polynomial is requested (and a first-order one for the additive and multiplicative corrections), and the expectation is that the LASSO finds thepertinent zero and non-zero PCE coefficients. As displayed in Figure 1, this works well for the high-fidelitysparse (HFS) as well as multi-fidelity sparse (MFS) polynomial chaos expansion models yielding an RMSEthat is lower than the corresponding standard PCE. In particular, the two-dimensional MFS model is ordersof magnitude better. It should be noted that once the number of training points is large enough to obtainthe exact fourth-order polynomial, the sparse models are not quite as good as the full order PCEs (compare

    solid to dashed lines) due to the extra non-zero term λ||β̂||1 in equation (11).

    Exponential-sine Function

    For the sum of exponential-sine products, the multi-fidelity surrogate model essentially needs the freedomto multiply the LF function by the polynomial expansion of the exponential function and subtract off thecross-dimensional terms. While it is still expected that multi-fidelity data will be helpful, it will be not aspowerful as in the Rosenbrock function example since the required bridge function is not a simple polynomialitself. The number of LF training points in the considered two and six dimensions is always three and onelevel higher, respectively. For the sparse polynomial chaos expansion models, many different orders for thepolynomials are combined with various values for λ to find the combination with the lowest RMSE. Theresulting best values are shown in Tables 2 and 3 for the two- and six-dimensional case, respectively.

    Table 2. Best order and λ values for two-dimensional exponential-sine function.

    # of HF points # of LF points λ Order, P Additive, R Multiplicative, Q

    5 0 10−6 4 – –

    9 0 10−3 5 – –

    17 0 10−3 6 – –

    33 0 10−6 7 – –

    65 0 10−6 10 – –

    5 33 10−3 9 2 2

    9 33 10−3 10 2 5

    17 65 10−6 5 4 2

    33 97 10−6 9 6 5

    65 161 10−8 9 6 6

    Table 3. Best order and λ values for six-dimensional exponential-sine function.

    # of HF points # of LF points λ Order, P Additive, R Multiplicative, Q

    13 0 10−6 4 – –

    73 0 10−3 4 – –

    257 0 10−3 5 – –

    737 0 10−6 6 – –

    13 73 10−4 4 3 2

    73 257 10−2 8 4 3

    257 737 10−6 8 5 4

    737 1889 10−6 9 5 5

    7

  • The results for the exponential-sine function with these settings are plotted in Figure 2 for two and sixdimensions. One can infer, generally speaking, that the sparse PCE surrogate models outperform the fullorder ones (compare solid to dashed lines), and the multi-fidelity models are better than the mono-fidelityones (compare red to black lines). In six dimensions this is especially true for 73 HF training points (and257 LF, if requested) and in two dimensions for 33 HF training points (and 97 LF ones).

    10-5

    10-4

    10-3

    10-2

    10-1

    100

    5 9 17 33 65

    RM

    SE

    Number of HF training points

    HFMF

    HFSMFS

    (a) Two dimensions

    0

    0.05

    0.1

    0.15

    0.2

    0.25

    0.3

    13 73 257 737

    RM

    SE

    Number of HF training points

    HFMF

    HFSMFS

    (b) Six dimensions

    Figure 2. RMSE comparisons for multidimensional exponential-sine function.

    IV. Aeroelastic Flutter Analysis

    In an aeroelastic flutter analysis it is necessary to capture the complex and coupled physical phenomenapresent in the operating environment and flight regime of modern aircraft. Here, a process developed byRumpfkeil and Beran30 will be employed. The Computational Aircraft Prototype Syntheses (CAPS)31

    program is used to map a given set of design parameters into the required aerodynamic and structuralanalysis models and to generate computational grids. Some resulting entities for the AGARD 445.6 dynamicaeroelastic test case model are shown in Figure 3.

    The employed structural solver is the Automated STRuctural Optimization System (ASTROS).32 AS-TROS can perform static, modal, and transient linear FEA, and has an internal aerodynamics capability forstatic and dynamic aeroelastic analyses which is employed in this work as low-fidelity model. A number-of-panel independent solution can be achieved using one-hundred aerodynamic panels as shown in Figure 3(a).For the actual flutter analysis the p-k method implemented in ASTROS is utilized. The structural modelitself is very simple since it contains no ribs and spars. The membrane thickness is set to 4.65mm and thematerial is assumed to be mahogany timber which is isotropic with a Young’s modulus of 3.2 · 109 Pa, shearmodulus of 4.1 · 108 Pa, Poisson’s ratio of 0.31, and density of 586 kg/m3. A grid convergence study forthe first three modal frequencies resulted in a relatively grid independent solution for the mesh shown inFigure 3(c) containing 279 grid nodes, 242 quadrilateral and 312 triangular elements.

    This structural model is also used within the modal structural decomposition approach in NASA’s Fully-Unstructured Navier-Stokes 3D (FUN3D)33 code. The Euler mode serves in this paper as a high-fidelitymodel. Three mode shapes are included in the dynamic analysis, and the transfer of mode shapes fromthe structural mesh to the fluid surface mesh is handled by CAPS. FUN3D performs the required volumemesh deformation internally via a linear elastic analogy driven by the surface mesh displacements.34 Agrid convergence study was conducted yielding satisfactory CFD volume meshes with approximately 47, 000nodes and 250, 000 tetrahedrals. A CFD surface mesh can be seen in Figure 3(b). The optimized second-order backward difference (BDF2OPT) scheme is employed for temporal discretization. The time step sizeis selected to have eighteen steps per cycle of the highest modal frequency and a total of 600 time stepsare used to simulate the unsteady aeroelastic behavior with an initial perturbation to “kick” the elasticresponse. In order to be able to automatically determine the dynamic pressure at the flutter threshold, qf ,a bisection method is employed by varying the flutter velocity until a neutral lift coefficient oscillation isachieved based on a normalized linear trendline through a Hilbert transformation yielding an instantaneousenvelope amplitude.30

    8

  • (a) Panels for low-fidelity aerodynamics. (b) OML for FEA and CFD as well as CFD surface mesh.

    (c) Internal structure for FEA.

    XY

    Z

    (d) Fluid domain for CFD.

    Figure 3. Representations of the AGARD 445.6 wing for multi-fidelity and multi-disciplinary analyses. Thickness tochord ratio is exaggerated in (b) for better visualization.

    Figure 4 shows ASTROS and FUN3D simulation results for qf at zero angle of attack as a function ofMach number together with the experimental data.3 One can observe a decent agreement among ASTROS(green lines), FUN3D (red lines) and the experimental data (black circles).

    V. Flutter Databases

    Rumpfkeil and Beran30 considered the variations of the critical flutter dynamic pressure, qf , with changesin Mach number (0.6 ≤M ≤ 1.2), angle of attack (0◦ ≤ α ≤ 5◦) and thickness to chord ratio (4% ≤ tc ≤ 8%).For each case an “exact” database was obtained from high-fidelity (FUN3D) analyses on a Cartesian meshand was used for comparisons and error assessment. One low-fidelity (ASTROS) simulation runs about150 times faster on one core than the corresponding high-fidelity simulation on six cores. Here, the newlydeveloped MFSPCE framework will be applied to the same flutter database generation problem using twoand three inputs.

    9

  • Figure 4. AGARD 445.6 flutter design space with one input. The green and red lines are ASTROS and inviscid FUN3Dresults, respectively, and the black circles are experimental data.3

    V.A. Two-dimensional Models

    Figure 5 compares the two fidelity levels in the domain of interest if only two inputs, namely Mach numberand angle of attack, are varied. For the truth data a Cartesian mesh of 15× 6 = 90 nodes is employed. Onecan see that the lower fidelity trends match the high-fidelity ones, which is encouraging for the use of amulti-fidelity approach.

    When building the MFSPCE models the number of LF training points is chosen to be always fourgrid levels higher (compare with Table 1). Many different orders for the actual sparse polynomial chaosexpansion as well as additive and multiplicative polynomials are combined with various values for λ to findthe combination with the lowest RMSE. The resulting best values are shown in Table 4.

    Table 4. Best order and λ values for two dimensions.

    # of HF points # of LF points λ Order, P Additive, R Multiplicative, Q

    5 0 10−2 4 – –

    9 0 10−2 5 – –

    17 0 10−1 7 – –

    33 0 100 7 – –

    5 33 100 12 7 2

    9 65 100 14 5 6

    17 97 10−2 12 3 6

    33 97 10−2 13 3 5

    Figure 6 shows the exact and MFSPCE model using 17 HF and 97 LF training points. It can be inferredthat the MFSPCE model is in good agreement with the exact model and is able to capture the transonicdip behavior very well, demonstrating its ability to model highly non-linear functions.

    10

  • Figure 5. AGARD 445.6 flutter design space with two inputs. The green and red surfaces are ASTROS and inviscidFUN3D results, respectively, and the black spheres are experimental data.3

    Figure 6. Two-dimensional exact (white) and MFSPCE model (red). HF training points are shown as black and LFones as white spheres.

    11

  • Lastly, Figure 7 exhibits the quantitative performance with and without the enhancement via low-fidelitydata as well as employing compressed sensing or using full-order polynomials. As one can infer, the sparse

    0

    200

    400

    600

    800

    1000

    1200

    1400

    1600

    5 9 17 33

    RM

    SE

    Number of HF training points

    HFLFMF

    HFSMFS

    500

    1000

    1500

    2000

    2500

    3000

    3500

    4000

    5 9 17 33

    Max

    Err

    or

    Number of HF training points

    HFLFMF

    HFSMFS

    Figure 7. RMSE (left) and maximum error (right) for various PCE surrogate models as a function of number of HFtraining points in two dimensions.

    PCE surrogate models outperform the full order ones (compare solid to dashed lines) and the multi-fidelitymodels are better than the mono-fidelity ones (compare red to black lines). For just informational purposesthe green line represents the RMSE when the model is built with only LF training points.

    V.B. Three-dimensional Models

    Figure 8 shows isosurfaces of both fidelity level simulation results for qf as a function of Mach number, angleof attack and thickness to chord ratio. For the truth data a Cartesian mesh of 13× 2× 5 = 130 nodes is

    Figure 8. AGARD 445.6 flutter design space with three inputs. The green and red edges are ASTROS and inviscidFUN3D results, respectively. Shown are three isosurfaces with qf = 3750Pa, qf = 7500Pa and qf = 11250Pa.

    12

  • employed. One can see that the low-fidelity trends match the high-fidelity ones which is encouraging for theuse of a multi-fidelity approach. In this three-dimensional case the number of LF training points is alwaysone grid level higher. For the sparse polynomial chaos expansion models again many different orders for thevarious polynomials are combined with various values for λ to find the combination with the lowest RMSEand the best values are shown in Table 5.

    Table 5. Best order and λ values for three dimensions.

    # of HF points # of LF points λ Order, P Additive, R Multiplicative, Q

    7 0 100 7 – –

    19 0 100 3 – –

    39 0 100 3 – –

    7 19 10−6 5 1 3

    19 39 100 5 1 4

    39 87 100 3 1 1

    Figure 9 shows the exact and MFSPCE model using 19 HF and 39 LF training points. The surrogatemodel does not agree that well with the truth model.

    Figure 9. Three-dimensional exact (white edges) and MFSPCE model (red edges). HF training points are shown asblack and LF ones as white spheres.

    Finally, Figure 10 displays the quantitative performance with and without the enhancement via low-fidelity data and compressed sensing. The sparse PCE surrogate models marginally outperform the fullorder ones (compare solid to dashed lines) and the multi-fidelity models are somewhat better than themono-fidelity ones (compare red to black lines).

    13

  • 2000

    2500

    3000

    3500

    4000

    4500

    5000

    7 19 39

    RM

    SE

    Number of HF training points

    HFLFMF

    HFSMFS

    6000

    7000

    8000

    9000

    10000

    11000

    12000

    13000

    14000

    15000

    16000

    7 19 39

    Max

    Err

    or

    Number of HF training points

    HFLFMF

    HFSMFS

    Figure 10. RMSE (left) and maximum error (right) for various PCE surrogates as a function of number of HF trainingpoints in three dimensions.

    V.C. Leave-One-Out Error Performance

    Table 6 shows the results for the two-dimensional flutter database if instead of minimizing the RMSE, whichusually cannot be computed, the leave-one-out error, ELOO, is minimized to decide which λ-value and ordersfor the actual sparse polynomial chaos expansion as well as additive and multiplicative polynomials shouldbe used. These values can be compared with the ones given in Table 4 to see that the results vary.

    Table 6. Best order and λ values for two dimensions based on leave-one-out error.

    # of HF points # of LF points λ Order, P Additive, R Multiplicative, Q

    5 0 100 4 – –

    9 0 10−4 8 – –

    17 0 10−4 7 – –

    33 0 10−4 8 – –

    5 33 10−4 12 7 4

    9 65 10−1 11 2 2

    17 97 10−1 12 4 7

    33 97 10−1 12 7 6

    This leads to different errors in the final MFSPCE surrogate models as displayed in Figure 11. The solidlines are the results for the MFSPCE models already displayed in Figure 7 using the values from Table 4.The dashed lines are the results if using the values from Table 6 instead, that is, the parameter search isguided by the leave-one-out error instead of the RMSE. For completeness, the dotted lines are using thevalues from Table 6 but the leave-one-out error (ELOO) instead of the RMSE is plotted. As one can infer bycomparing solid and dashed lines of the same color, basing the parameter values on the leave-one-out erroryields worse MFSPCE models, especially for the cases with few HF training points. This is expected sincethe leave-one-out error only converges to the true RMSE for a large number of training points. Interestingly,ELOO itself is lower than the actual RMSE (compare dotted and dashed lines) except for the case with fivetraining points and only HF data.

    VI. Conclusions

    This paper has presented multi-fidelity sparse polynomial chaos expansion (MFSPCE) models of criticalflutter dynamic pressures for the AGARD 445.6 aeroelastic test case model as a function of Mach number,angle of attack and thickness to chord ratio. The highest fidelity level were body fitted Euler flow solutionsand the lower level were panel solutions, both combined with the same modal structural solver. Fairlyaccurate surrogate models of the flutter database could be constructed using only twenty or so high-fidelityEuler training points. Before these flutter databases were generated the MFSPCE model framework was

    14

  • 0

    200

    400

    600

    800

    1000

    1200

    1400

    1600

    5 9 17 33

    RM

    SE

    / E

    LOO

    Number of HF training points

    HFSMFS

    HFS based on ELOOMFS based on ELOO

    HFS all ELOOMFS all ELOO

    Figure 11. RMSE and leave-one-out (LOO) error for various PCE surrogate models as a function of number of HFtraining points in two dimensions.

    successfully verified with analytic test functions. Overall, multi-fidelity models yield more accurate resultscompared to using high-fidelity data alone especially at the beginning of the simulation when only a handfulHF data points are available. The sparse PCEs tend to be a factor of two or so better than the full-ordercounterparts in terms of accuracy as measured by the root-mean-square error (RMSE) compared to a truthmodel.

    Acknowledgments

    This work was supported by the US Air Force Summer Faculty Fellowship Program and the US Air ForceResearch Laboratory (AFRL). The third author acknowledges support of US Air Force Office of ScientificResearch.

    References

    1Bendiksen, O. O., “Transonic Limit Cycle Flutter of High-Aspect-Ratio Swept Wings,” Journal of Aircraft , Vol. 45, No.5, 2008, pp. 1522–1533.

    2Bendiksen, O. O., “High-Altitude Limit Cycle Flutter of Transonic Wings,” Journal of Aircraft , Vol. 46, No. 1, 2009,pp. 123–136.

    3Carson Yates, Jr., E., “AGARD Standard Aeroelastic Configurations for Dynamic Response I - Wing 445.6,” Tech. Rep.AGARD 765, 1988.

    4Isukapalli, S., Roy, A., and Georgopoulos, P., “Efficient sensitivity/uncertainty analysis using the combined stochasticresponse surface method and automated differentiation: Application to environmental and biological systems,” Risk Analysis,Vol. 20, 2000, pp. 591–602.

    5Kim, N., Wang, H., and Queipo, N., “Adaptive reduction of random variables using global sensitivity in reliability-basedoptimisation,” International Journal of Reliability and Safety, Vol. 1, 2006, pp. 102–119.

    6Roderick, O., Anitescu, M., and Fischer, P., “Polynomial regression approaches using derivative information for uncer-

    15

  • tainty quantification,” Nuclear Science and Engineering, Vol. 164, No. 2, 2010, pp. 122–139.7Boopathy, K. and Rumpfkeil, M. P., “A Unified Framework for Training Point Selection and Error Estimation for

    Surrogate Models,” AIAA Journal , Vol. 53, 2015, pp. 215–234.8Bryson, D. E. and Rumpfkeil, M. P., “All-at-Once Approach to Multifidelity Polynomial Chaos Expansion Surrogate

    Modeling,” Aerospace Science and Technology, Vol. 70C, 2017, pp. 121–136.9Gano, S., Renaud, J. E., and Sanders, B., “Variable fidelity optimization using a kriging based scaling function,” AIAA

    Paper, 2004-4460, 2004.10Eldred, M. S., Giunta, A. A., Collis, S. S., Alexandrov, N. A., and Lewis, R., “Second-order corrections for surrogate-based

    optimization with model hierarchies,” AIAA Paper, 2004-4457, 2004.11Ng, L. W. T. and Eldred, M. S., “Multifidelity Uncertainty Quantification Using Non-Intrusive Polynomial Chaos and

    Stochastic Collocation,” AIAA Paper, 2012-1852, 2012.12Bryson, D. E. and Rumpfkeil, M. P., “Comparison of Unified and Sequential-Approximate Approaches to Multifidelity

    Optimization,” AIAA Paper, 2017-0131, 2017.13Han, Z. H., Goertz, S., and Zimmermann, R., “Improving variable-fidelity surrogate modeling via gradient-enhanced

    kriging and a generalized hybrid bridge function,” Aerospace Science and Technology, Vol. 25, No. 1, 2013, pp. 177–189.14Alexandrov, N. M., Lewis, R. M., Gumbert, C. R., Green, L. L., and Newman, P. A., “Approximation and Model

    Management in Aerodynamic Optimization with Variable-Fidelity Models,” Journal of Aircraft , Vol. 38(6), 2001, pp. 1093–1101.

    15Choi, S., Alonso, J. J., Kroo, I. M., and Wintzer, M., “Multi-fidelity design optimization of low-boom supersonic businessjets,” AIAA Paper, 2004-1530, 2004.

    16R. M. Lewis, S. G. N., “A multigrid approach to the optimization of systems governed by differential equations,” AIAAPaper, 2000-4890, 2000.

    17Blatman, G. and Sudret, B., “Adaptive sparse polynomial chaos expansion based on least angle regression,” Journal ofComputational Physics, Vol. 230(6), 2011, pp. 2345–2367.

    18Doostan, A. and Owhadi, H., “A non-adapted sparse approximation of PDEs with stochastic inputs,” Journal of Com-putational Physics, Vol. 230(8), 2011, pp. 3015–3034.

    19Davenport, M. A., Duarte, M. F., Eldar, Y., and Kutyniok, G., Introduction to compressed sensing. In: Eldar YC,Kutyniok G, editors. Compressed sensing: theory and applications, Cambridge University Press, 2011.

    20Jakeman, J. D., Eldred, M. S., and Sargsyan, K., “ Enhancing l1-minimization estimates of polynomial chaos expansionsusing basis selection,” Journal of Computational Physics, Vol. 289, 2015, pp. 18–34.

    21Ng, L. W. T., Huynh, D. B. P., and Willcox, K. E., “Multifidelity Uncertainty Propagation for Optimization UnderUncertainty,” 12th ATIO and 14th AIAA/ISSMO Conference, 2012.

    22Shah, H., Hosder, S., Koziel, S., Tesfahunegn, Y. A., and Leifsson, L., “Multi-fidelity robust aerodynamic design opti-mization under mixed uncertainty,” Aerospace Science and Technology, Vol. 45, 2015, pp. 17–29.

    23Palar, P. S., Tsuchiya, T., and Parks, T. G., “Multi-Fidelity Non-Intrusive Polynomial Chaos Based on Regression,”Computer Methods in Applied Mechanics and Engineering, Vol. 305.

    24Robinson, T. D., Eldred, M. S., Willcox, K. E., and Haimes, R., “Surrogate-Based Optimization Using MultifidelityModels with Variable Parameterization and Corrected Space Mapping,” AIAA Journal , Vol. 46, No. 11, 2008, pp. 2814–2822.

    25Li, Y., Anitescu, M., Roderick, O., and Hickernell, F., “Orthogonal Bases for Polynomial Regression with DerivativeInformation in Uncertainty Quantification,” International Journal for Uncertainty Quantification, Vol. 1, No.4, 2011, pp. 297–320.

    26Burkardt, J., “SGMGA: Sparse Grid Mixed Growth Anisotropic Rules,” 2010.27Salehi, S., Raisee, M., Cervantes, M. J., and Nourbakhsh, A., “Efficient Uncertainty Quantification of Stochastic CFD

    Problems Using Sparse Polynomial Chaos and Compressed Sensing,” Computers and Fluids, Vol. 154, 2017, pp. 296–321.28Curtin, R. R., Cline, J. R., Slagle, N. P., March, W. B., Ram, P., Mehta, N. A., and Gray, A. G., “MLPACK: A Scalable

    C++ Machine Learning Library,” Journal of Machine Learning Research, Vol. 14, 2013, pp. 801–805.29Allen, D. M., “The relationship between variable selection and data agumentation and a method for prediction,” Tech-

    nometrics, Vol. 16(1), 1974, pp. 125–7.30Rumpfkeil, M. P. and Beran, P., “Multi-Fidelity Surrogate Models for Flutter Database Generation,” Proceedings of the

    Tenth International Conference on Computational Fluid Dynamics, ICCFD10, Barcelona, Spain, July, 2018.31Alyanak, E., Durscher, R., Haimes, R., Dannenhoffer, J. F. I., Bhagat, N., and Allison, D., “Multi-fidelity, Gemoetry-

    centric Multi-disciplinary Analysis for Design,” AIAA Paper, 2016-4007, 2016.32Johnson, E. H. and Venkayya, V. B., “Automated structural optimization system (ASTROS) Volume I - Theoretical

    manual,” Tech. Rep. AFWAL-TR-88-3028, Air Force Wright Aeronautical Laboratories, 1988.33Carlson, R. B. J., Derlaga, J., Gnoffo, P., Hammond, D., Jones, W., Kleb, B., Lee-Rausch, E., Nielsen, E., Park, M.,

    Rumsey, C., Thomas, J., and Wood, W., “FUN3D Manual: 12.7,” NASA Technical Memorandum 218761, 2015.34Biedron, R. T. and Thomas, J. L., “Recent Enhancements to the FUN3D Flow Solver for Moving-Mesh Applications,”

    AIAA Paper, 2009-1360, 2009.

    16