Dynamic Aeroelastic Analysis Using Reduced-Order Modeling

2y ago
7 Views
2 Downloads
1.54 MB
20 Pages
Last View : 1m ago
Last Download : 3m ago
Upload by : Oscar Steel
Transcription

Dynamic Aeroelastic Analysis Using Reduced-Order Modelingwith Error EstimationBrandon M. Lowe and David W. Zingg †Institute for Aerospace Studies, University of Toronto,4925 Dufferin St, Toronto, Ontario, M3H 5T6, CanadaThis paper presents a methodology for dynamic aeroelastic analysis of aircraft based onmodel order reduction with error estimation. A projection-based model order reduction approach is used to create an aerodynamic reduced-order model (ROM) which is coupled to astructural model to create an aeroelastic ROM. The governing aerodynamic equations are thelinearized semi-discrete Euler equations. Flutter analysis is conducted by analyzing the eigenvalues of the aeroelastic ROM. A dual-weighted residual-based error estimator is presentedwhich approximates the error in the eigenvalues obtained from the reduced eigenproblemrelative to the eigenvalues from the high-dimensional aeroelastic model. The error estimator thus allows for the construction of aeroelastic ROMs with select eigenvalues that satisfya user-prescribed accuracy. The aerodynamic ROM is constructed using approximate highdimensional aeroelastic eigenvectors computed using the two-sided Jacobi-Davidson algorithm.Dynamic aeroelastic analyses are presented for a two degree of freedom structural model andfor the AGARD 445.6 wing test case. The error estimator is shown to have good agreementwith the exact error. For the test cases presented in this work, the cost of computing the flutterpoint at a given Mach number is equivalent to the cost of approximately 4 to 5 steady nonlinearflow evaluations of the high-dimensional Euler equations.I. IntroductionIrecent years, a strong emphasis has been placed on the need for increased aircraft fuel efficiency. Efficient designscan be conceived with the aid of aerodynamic shape optimization and multidisciplinary optimization tools basedon high-fidelity computational fluid dynamics (CFD) [1]. However, the resulting unconventional aircraft are typicallydesigned to fly at transonic Mach numbers, where a dip in the flutter boundary typically occurs due to shock formationand motion.Linear aerodynamic methods (such as the doublet lattice method) have been widely used for flutter predictions in thepast due to their low computational cost. However, these methods fail to model nonlinear flow features accurately inthe transonic regime [2]. This lack of accuracy has led to overly conservative aircraft designs and hence reduced fuelefficiency [3]. Alternatively, time-dependent high-dimensional CFD simulations present a means to model aerodynamicsaccurately for dynamic aeroelastic analysis. However, due to the unsteady nature of the problem, the large number ofdegrees of freedom, and the number of flight conditions to consider, the use of CFD for flutter predictions in a designcontext remains intractable. As a result, numerous CFD-based alternatives have been proposed for flutter predictions. Anumber of researchers have applied a Hopf point calculation method to obtain the flutter point of CFD-based aeroelasticsystems [4–6]. Jacobson et al. [7] presented a linearized frequency-domain approach for predicting the onset of flutter.Their approach uses a p-k flutter analysis method with generalized aerodynamic forces obtained from the solutionsof the linearized Reynolds-averaged Navier-Stokes (RANS) equations in the frequency domain. He et al. [8] havepresented a time-spectral approach to solve for the flutter point of the aeroelastic equations. They simultaneously solvefor flutter speed index and the flutter frequency along with the flow states at all time instances. Opgenoord et al. [9] havedeveloped a low-order aerodynamic model which is calibrated using two-dimensional high-fidelity CFD simulationsand uses strip theory to extend its applications to three-dimensional geometries.A large breadth of literature has also been devoted to the application of model order reduction for flutter analysis[10–27]. These methods aim to generate reduced-order models (ROMs) which accurately capture the dynamic behaviorof the high-dimensional model (HDM) in a fraction of the computational time. In the case of flutter analysis, the CFDPhD Candidate, University of Toronto, brandon.lowe@mail.utoronto.caof Toronto Distinguished Professor of Computational Aerodynamics and Sustainable Aviation, Director, Centre for Research inSustainable Aviation, dwz@oddjob.utias.utoronto.ca† University1

model constitutes the HDM. We do not provide an exhaustive survey of the field, but here we describe a few notableapplications of ROMs for aeroelastic analyses. Hall et al. [10] and Thomas et al. [11] were among the first to use aproper orthogonal decomposition (POD)-based ROM for aeroelastic modeling. Using the linearized Euler equationsin the frequency-domain, they constructed ROMs which provide flutter boundaries with good agreement to solutionsfrom the HDM. A few years later, Lieu et al. [14, 15] modeled the dynamic aeroelastic behavior of a complete aircraftconfiguration using a POD-based ROM constructed from the frequency-domain linearized Euler equations. Beran et al.[12] used a POD-based ROM to predict the onset of limit-cycle oscillations for a nonlinear panel in a two-dimensionalflow. Silva et al. [19, 20] performed flutter predictions by constructing aerodynamic ROMs based on aerodynamicforces from high-dimensional CFD simulations using the eigensystem realization algorithm [28]. Argaman and Raveh[26] have modeled an aeroelastic system as a multi-output autoregressive process with model parameters identifiedusing aeroelastic responses from CFD solutions. Previously, the authors of this paper have also presented a ROM-basedapproach for flutter predictions [27]. In this previous work, an aerodynamic ROM was constructed using a reducedbasis created by applying POD to snapshots obtained by exciting the structural states during a single unsteady flow solve.The aerodynamic ROM was then coupled to the structural model, and the eigenvalues of the resulting aeroelastic ROMwere analyzed to predict the onset of flutter.All of the aforementioned ROM-based approaches are capable of predicting flutter. However, none are capable ofestimating the accuracy of their flutter prediction relative to the original aeroelastic HDM. In this paper, we present aROM-based aeroelastic analysis approach with an error estimator capable of providing a user-prescribed level of accuracyfor flutter predictions relative to the aeroelastic HDM. Dynamic aeroelastic analysis is performed by analyzing theeigenvalues of the aeroelastic ROM. The dual-weighted residual (DWR) method [29] is used to provide an error estimatefor the eigenvalues. The error estimator provides the user with a measure of accuracy for the approximated aeroelasticeigenvalues and also serves as an indicator to determine at which flight conditions to update the reduced basis. This leadsto an automated ROM training procedure for which the user needs only to supply some parameters for initial snapshots,a range of relevant flight conditions, and the desired level of accuracy in the eigenvalues. Furthermore, an extensionof the eigenvalue error estimator is presented to obtain approximate errors for the predicted flutter point. Similar tothe work by Lowe and Zingg [27], we rely on the linearized semi-discrete Euler equations as our high-dimensionalaerodynamic system; however the reduced basis is now constructed using approximate eigenvectors of the aeroelasticHDM obtained using the two-sided refined Jacobi-Davidson algorithm [30–32]. Our use of eigenvectors and thetwo-sided Jacobi-Davidson algorithm for model order reduction is similar to the work presented by Benner et al. [33].However, in our work, the left eigenvectors are not used for the test basis in the construction of the ROM, but are ratherused to construct a second ROM for error estimation. We also note that the methodology presented herein is not limitedto the use of the linearized Euler equations, and can be extended to the linearized RANS equations.It has been shown that the DWR method provides effective error estimation for aerodynamic problems [34] as wellas eigenvalue problems [35]. The DWR method requires the solution to a dual (or adjoint) problem in a subspacenot spanned by the reduced basis used to create the aeroelastic ROM. We approximate dual solutions with a secondaerodynamic ROM created using a reduced basis trained on dual solutions. As will be discussed in Section IV, the dualsolution to the aeroelastic eigenproblem is in fact the left eigenpair of the system. Thus, we denote the aeroelastic ROMused for the main aeroelastic analysis as the primal ROM, and the aeroelastic ROM used for the error analysis as thedual ROM.The remainder of this paper is divided into the following sections. In Section II, we present the high-dimensionalequations used to model aeroelastic behavior, including the structural and aerodynamic models. Section III describesthe model order reduction approach used to create the primal aeroelastic ROM. In Section IV, we give an overviewof the error estimator derivation, and describe the procedure to create the dual aeroelastic ROM. Section V presentsour methodology for training the primal and dual reduced bases for accurate flutter predictions. Section VI gives theapproach for obtaining the flutter point once the ROM has been sufficiently trained, and also presents the extension ofthe error estimator to compute the error in the predicted flutter point. Lastly, Section VII contains the results obtainedusing the methodology derived in this paper.II. High-Dimensional Semi-discrete Aeroelastic ModelThis section presents the high-dimensional semi-discrete equations used to model dynamic aeroelastic behavior.An overview of the structural model is first presented. Subsequently, the linearized semi-discrete Euler equationsare derived and used to form the aerodynamic model. Finally, the structural and aerodynamic models are coupled toform the high-dimensional monolithic aeroelastic model. For ease of presentation, certain quantities associated to the2

aerodynamic model are indicated with the subscript a, whereas those for the structural models are presented with thesubscript s.A. Structural ModelFor dynamic aeroelastic analysis, structural states are commonly expanded into a modal series consisting of thesummation of free vibration modes weighted by time-dependent generalized displacements. Neglecting contributionsfrom damping, one may write the resulting equations of motion as follows:3 2 u s,tot s u s,tot fs,tot3C s2(1)where Cs is the time variable for the structural model, u s,tot is the vector of generalized displacements, fs,tot is the vectorof generalized forces applied to the structural grid, and s is a diagonal matrix of squared natural frequencies. We wishto linearize equation (1) about a steady-state solution. The relevant states are expanded as:u s,tot u s,0 u s ,(2)u a,tot u a,0 u a ,where u a,tot is the vector of aerodynamic states, and steady-state quantities are denoted with the subscript 0. Interest liesin modeling the state fluctuations u s and u a , and their impact on the fluctuations of the applied forces: fs fs,tot fs,0 .Forces are computed on the aerodynamic grid, which does not correspond to the structural grid. Additionally, theaerodynamic equations are typically nondimensionalized and thus applied forces require a dimensionalization factor.We introduce the force scaling factor ff , the force transfer matrix )f , and the displacement transfer matrix )d , such thatfs ff )f fa ,(3)x a,surf )d u s ,where fa is the vector of aerodynamic force fluctuations computed on the aerodynamic surface grid, and xa,surf is thevector of aerodynamic surface grid node coordinate fluctuations. The transfer matrices )f and )d may be computed in anumber of ways. For complex structural geometries, popular methods include Rendall and Allen’s radial basis functioninterpolation [36], and Brown’s rigid link method [37]. For two degree of freedom structural models acting at a singlepoint, the transfer matrix can be formed using a small angle approximation. As the applied forces are a function of boththe aerodynamic and structural states, we can use the transfer matrices to exchange information between structural andaerodynamic grids. Expanding the fluctuations of the force into a Taylor series and truncating high-order terms gives, m fam fafs ff )f fa ff )f)d u s ua .(4)mxa,surfmu aThe linearized structural model thus becomes,where:s ff )f3 2 us 3C s2s usm fa)dmxa,surf s ,(5) B u a , B ff )fm fa.mu aB. Aerodynamic ModelThe governing aerodynamic equations used for dynamic aeroelastic analysis are the linearized semi-discrete Eulerequations. Before linearization, the Euler equations are discretized in space using a second-order summation-by-partsapproach, and are then put into an arbitrary Lagrangian-Eulerian form. This leads to the following set of ordinarydifferential equations:3 1 u a,tot Xa u a,tot (C a ), xa,vol,tot (C a ), x§ a,vol,tot (Ca ) .(6)3C aHere xa,vol,tot and x§ a,vol,tot are the vectors of volume grid node coordinates and velocities, respectively. Additionally, C ais the time variable for the aerodynamic model, Xa is the residual vector, and 1 is a diagonal matrix of inverse metricJacobians of the transformation to curvilinear coordinates. In order to preserve a uniform flow in the presence of adeforming grid, the evolution of the 1 is governed by the Geometric Conservation Law [38].3

The linearized semi-discrete Euler equations are obtained by assuming the flow states, grid node coordinates, andgrid node velocities can be represented as linear fluctuations about a nonlinear steady-state:u a,tot u a,0 u a ,xa,vol,tot xa,vol,0 xa,vol ,(7)x§ a,vol,tot x§ a,vol .Steady-state grid node velocities are assumed to be zero. The resulting linearized semi-discrete Euler equations are: m Xam Xam3 1m Xa1 3u a u x u (8)am a,surfa,0m x§ a,surf 0,03C amu amxa,volm x§ a,vol 3C am x§ a,volwheremis a linear mesh movement matrix such that:xa,vol x§ a,vol m x a,surf ,(9)m x§ a,surf .Here xa,surf and x§ a,surf are the vectors of aerodynamic surface grid node coordinate fluctuations and velocities. Fordetails on the derivation of the linearized Euler equations, see Lowe and Zingg [27].The aerodynamic model is nondimensionalized in time differently than the structural model. To address this, thefollowing time scaling factor is introduced:Cs ft C a .(10)Additionally, as for the force transfer discussed above, we must transfer displacements from the structural grid to theaerodynamic surface grid. We use the displacement transfer matrix )d such that,3xa,surf3u s ft)d,3C a3C sxa,surf )d u s ,(11)Introducing equations (10) and (11) into (8), the aerodynamic model is given by, a3u a 3C sa ua a,1 u s a,23u s,3C swhere the system matrices are: a ft01,a m Xa,mu a a,1 m Xamxa,volm)d , a,2 ft (12) m3 1m Xau a,0 m x§ a,vol 3C am x§ a,volm)d .C. Monolithic Aeroelastic ModelWe wish to couple the linear structural and aerodynamic models into a single monolithic aeroelastic system.Introducing the notation,3u s u§ s ,(13)3C swe write equations (5), (12), and (13) into a complete set of aeroelastic equations,22 u§ s 3 2 00 0 37 s 37s66 7 667 3 6 7 6707007606u 7 667 3C s 6 s 7 6760 0 a 76u a 7 6 a,2 a,17a5454 5 4where is the identity matrix. In a more convenient form, we write this asM3u Au,3C s2 u§ s 36 76 76us 7 ,6 76u a 74 5(14)(15) where u) u§ )s u)s u)a .Analogous to the “p-method” used in flutter analysis, we expand the states into simple exponential functions in timeas follows,u v exp ( Cs ) .(16)This leads to the following generalized eigenvalue problem to determine the stability of the system:(AM) v 0.(17)We refer to equation (17) as the high-dimensional eigenproblem, and we denote any eigenvalue and (right) eigenvectorv which satisfies (17) as a truth eigenvalue and truth eigenvector.4

III. Model Order ReductionDue to the use of the linearized Euler equations, the aeroelastic model (15) is of very high dimension and thusremains impractical for fast dynamic aeroelastic analysis. If instead of using the high-dimensional aerodynamic model(12) we approximate the behavior of the aerodynamics with a ROM, the aeroelastic eigenproblem becomes tractable.This section presents an overview of the application of projection-based model order reduction to the aerodynamicmodel used in this work. The approach used to form the reduced basis is discussed in Section V below.To begin, assume we possess a set of a reduced basis vectors stored column-wise in the reduced basis {51 , 52 , ., 5 a }. In projecting equation (12) onto the reduced basis, one approximates the aerodynamic solution in thereduced space [39]:u a ũ a ,(18)where ũ a 2 R a is a vector of modal coefficients. Inserting this approximation into the high-dimensional aerodynamicmodel (12), we obtain the residual,r ROM a3 ũ a3C sũ aa(19) a,2 u§ s . a,1 u sEnforcing the Galerkin condition, which states that the residual is orthogonal to the reduced space, leads to:), r ROM 0,(20)where , is a matrix which approximates a continuous inner-product in Euclidean space. The resulting aerodynamicROM is,3 ũ a), a ) , a ũ a ) , a,1 u s a,2 u§ s .(21)3C sTo differentiate this aerodynamic ROM from that used for the error estimator, we denote equation (21) as the primalaerodynamic ROM.Coupling the primal aerodynamic ROM to equations (5) and (13) forms the (primal) aeroelastic ROM,2666066040000) , a32 3 27 3 6 u§ s 7 676 7 676u 7 67 3C s 6 s 7 676ũ a 7 654 5 40s0), a,2), a,1 s0),aAs before, we write this in a more convenient form,M̃3 ũ Ãũ,3C s37777752 u§ s 36 76 76us 7 ,6 76ũ a 74 5(22)(23) where ũ) u§ )s u)s ũ)a . One can immediately see that the matrices ) , a and ) , a are now of order a , which is typically orders of magnitude smaller than the number of degrees of freedom in the high-dimensional s , the stability of the aeroelastic system can now be determined by theaerodynamic model. Assuming ũ ṽ exp Ctractable eigenproblem, Ã M̃ ṽ 0.(24)In general, due to the approximation of the right eigenvector in the reduced space,v ṽ,with26 6606604005(25)0 37707 .775

IV. Error EstimationThe stability of the aeroelastic system is determined by the eigenvalues of the high-dimensional eigenproblem(17). In order to ensure that the aeroelastic ROM provides sufficiently accurate results relative to the high-dimensionalmodel, we wish to estimate the error between the eigenvalues obtained from the reduced eigenproblem (24) and thecorresponding truth eigenvalues. If the estimated error exceeds a user-prescribed tolerance, then the aerodynamic ROMcan be trained to produce a more accurate eigenvalue. The following section presents the error estimator used in thiswork. The estimator is based on the work of Heuveline and Rannacher [35], who presented a DWR-based error estimatorfor eigenvalue problems. An overview of the approach is presented in this section; the reader is referred to Bangerth andRannacher [40] for details of the derivation.A. Dual-Weighted Residual-Based Error EstimationTo begin, we present the primal and dual problems of interest. The primal problem is the high-dimensional righteigenvalue problem from equation (17), rewritten here for convenience, with an extra normalization condition on theeigenvector,Av Mv,v Mv 1,(26)where a superscript denotes the conjugate transpose. For flutter analysis, we are specifically interested in theeigenvalue; thus we define the output functional as:(27)( , w) w Mw.Due to the normalization condition, this output functional provides the eigenvalue: ( , v) v Mv . For thisoutput functional, the dual solution is obtained from the high-dimensional left eigenproblem with an added normalizationcondition:(v du ) A (v du ) M,(v du ) Mv 1,(28)where v du is the truth left eigenvector. For further details see [40]. ṽ .Now, assume we have solved the reduced (primal) eigenproblem (24) for an approximate right eigenpair ,Introducing this approximate eigenpair into the high-dimensional right eigenproblem (26), we obtain the followingresidual: ṽ. r A M(29)Given the truth left eigenvector v du , the error in the solution is given by the equation, (v du ) , r R (2) ,(30)where R (2) is a second-order remainder term, and, as before, , approximates an inner product [40].Obtaining the error through equation (30) (with the remainder term neglected) remains impractical because werequire the truth left eigenvector v du from the high-dimensional left eigenproblem (28). To reduce computational cost, itwould be ideal to approximate the left eigenvector using a ROM. Importantly, the primal aeroelastic ROM (22) cannotbe used because the Galerkin condition states that the residual (29) is orthogonal to the subspace spanned by the primalreduced basis. Instead, in this work we obtain an approximate left eigenvector with a second ROM, denoted as the dualROM.B. Dual ROM and Error EstimatorTo obtain the dual aeroelastic system, we transpose the state and mass matrices from the primal high-dimensionalaeroelastic model (14), giving266606604002u§ du 3 2 00 376 s 7 67 3 6 du 7 6076u 7 67 3C s 6 sdu 7 6 )s76u a 7 6 a 54 5 4 s00 )a,2 377 )a,1 77) 7a 52u§ du 36 s 76 du 76u s 7 .6 du 76u a 74 5(31)Note that all matrices are real and that the matrices s and a are symmetric and thus do not require a transpose. Thedual aerodynamic model is then,3u du) du§s . a a )a u du(32)a s u3C s6

As for the primal aerodynamic ROM, we approximate the dual aerodynamic solution in a reduced space spanned by thereduced basis du :du duu duũ a ,(33)a where ũ dua is a vector of modal coefficients. From this approximation, we obtain the residual,dur ROM aEnforcing the Galerkin condition du )du ), adudu 3 ũ adu duũ a)a3C s s) u§ dus .(34)du, r ROM 0 leads to the dual aerodynamic ROM,dudu 3 ũ a3C s du ),du duũ a)a du ), s) u§ dus .(35)Armed with this dual aerodynamic ROM, we can form the dual aeroelastic ROM,26660660400du00), a32 du 3 27 3 6u§ s 7 676 du 7 676u s 7 676 du 7 63Csdu 76ũ a 7 654 5 40sdu ), s)00In a more convenient form, we write this as3 ũ du Ãdu ũ du .3C sThis directly leads to the reduced dual generalized eigenproblem, Ãdu du M̃du ṽ du 0,M̃du )a,2 )a,1du ),dudu)a37777du 752u§ du 36 s 76 du 76u s 7 .6 du 76ũ a 74 5(36)(37)(38)where ṽ du is a reduced approximation of the truth left eigenvector v du such thatv du du ṽ du ,(39)with2 00 37667 60(40)0 7.67du 760 045Introducing the left eigenvector approximation (39) into equation (30), and neglecting the remainder term, we obtain theerror estimator [ used in this work: [ du ṽ du , r.(41) duV. Reduced Basis TrainingIn this section, we define our approach to constructing an aeroelastic ROM capable of predicting aeroelasticeigenvalues to a user-prescribed accuracy. But first, we define the parameter on which our aeroelastic ROMs depend.Aerodynamic parameters including the Mach number and angle of attack are held fix during our analyses. Similarly,structural parameters such as the structural mass and Young’s modulus are also held fix. The parameter for theaeroelastic ROMs is the variable which affects the coupling between both the aerodynamic and structural models. For 1our purposes, the aeroelastic parameter can be either the dynamic pressure @ 1 or the speed index 1 l, where 1U is the freestream speed, 1 is the root semichord, l U is the uncoupled first torsional frequency, and is the mass ratio.The results presented in Section VII use aeroelastic systems formulated using the speed index. But to keep the approachmore general, we denote both @ 1 and as the “aeroelastic parameter” and use the placeholder .The error estimator (41) allows us to approximate the error in any eigenvalue obtained from the reduced aeroelasticeigenproblem (24). However, not all eigenvalues obtained from the aeroelastic ROM are important for flutter analysis.Certain eigenmodes in the aeroelastic system are closely associated to the modes of the structural model, and it is one ofthese eigenmodes that will become unstable. At low aeroelastic parameter values, these are easily determined because7

ImaginaryMode 4Mode 3Increasing q Mode 2Mode 1RealFig. 1 Example of an aeroelastic root locus plot for increasing dynamic pressure, showing only importantaeroelastic modes.each eigenvalue has an imaginary part near the natural frequency of its associated structural mode. As the aeroelasticparameter is increased, the eigenvalues change in value. Figure 1 demonstrates an example root locus for increasingdynamic pressure for an aeroelastic system with four structural modes.Based on this knowledge, we wish to create a reduced basis that spans the subspace of the eigenvectors of theseimportant eigenmodes for a range of aeroelastic parameter values. In this pursuit, we require three algorithmic elements:a means of constructing an initial reduced basis, a robust eigenvalue tracking algorithm, and a means of approximatingtruth eigenvectors. The eigenvalue tracking algorithm is required to know which eigenvalues are relevant for flutteranalysis and necessitate error estimation. A means of approximating truth eigenvectors is required in order to train thereduced basis to obtain better approximations of aeroelastic eigenvalues. In this section, we present our approach toeach of the aforementioned elements.A. Reduced Basis InitializationWhen initializing the primal and dual reduced bases, and du , the goal is not to create aeroelastic ROMs whichcan accurately predict the aeroelastic eigenvalues. Rather, the goal is to create primal and dual ROMs which canestimate the error in the eigenvalues with sufficient accuracy. In this work, the initial primal and dual reduced bases areconstructed from solutions of the primal and dual high-dimensional aerodynamic models in the frequency-domain. Toobtain these equations, the aerodynamic and structural vectors are assumed to be harmonic,’’u a (C) ū a, 9 4 8 lC ,u s (C) ū s, 9 4 8 lC ,(42)99where ū a, 9 and ū s, 9 are the aerodynamic and structural Fourier coefficients, respectively. Inserting these expressionsinto the high-dimensional linearized aerodynamic model (12), we obtain the governing primal aerodynamic equationsin the frequency-domain,(8l a(43)a ) ū a, 9 a,1 8l a,2 ū s, 9 .Using the same approach, we can convert the high-dimensional dual aerodynamic model (32) to the frequency-domain, )du) du8l a(44)a ū a, 9 8l a ū s, 9 .To obtain initial snapshots for the reduced bases, equations (43) and (44) are solved for several aeroelastic parametervalues. At each aeroelastic parameter value, these equations are solved for l equal to the natural frequencies of thestructure. At each natural frequency, the corresponding generalized velocity and displacement are set to unity and theinverse frequency, respectively, in the input vectors ū s, 9 and ū dus, 9 . Note that changing the aeroelastic parameter valueresults in changes to the system matrices in equations (43) and (44) due to different time and force scaling factors ft andff .8

We have found that solving equations (43) and (44) for five aeroelastic parameter values produces initial ROMswith good error estimation capability. Moreover, the equations are solved only to a relative tolerance of 10 1 to obtainthe initial snapshots. Once all of the snapshots of ū a, 9 and ū dua, 9 have been obtained, they are orthogonalized usingthe modified Gram-Schmidt procedure and used to create the initial reduced bases and du . Note that the real andimaginary parts of the snapshots are included as separate vectors in and du to produce only real valued reducedbases.B. Eigenvalue TrackingOnce the aeroelastic ROMs have been initialized, the reduced eigenproblem (24) can be used to approximateaeroelastic eigenvalues. At aeroelastic parameter values near zero, the eigenvalues relevant for flutter analysis are thosewith an imaginary part near a natural frequency of the structure. As the aeroelastic parameter value is increased, thevalues of these eigenvalues change, which leads to the need for tracking. Assume we have an approximate eigenvalue ( 1) known to be important for flutter analysis at the aeroelastic parameter increment 1. At increment , aftersolving the eigenproblem (24), we have a number of approximate eigenvalues 8( ) for 8 1, 2, ., and we wish to knowwhich eigenvalue 8( ) is associated to the known eigenvalue ( 1) . Several algorithms have been proposed for thispurpose [41, 42]. In this work, we use a mixture of three metrics to track the eigenvalues.The first tracking metric is based on the use of the error estimator (41). At increment , we have the approximateeigenvalues 8( ) , and approximate right and left eigenvectors, ṽ 8( ) and ṽ du( ), respectively. Assuming the eigenvectors8vary smoothly with changing aeroelastic parameter value, we can use the left eigenvector at the previous incrementassociated to the known eigenvalue ( 1) to estimate the error in the new 8 th approximate eigenvalue, f8err w A 8( ) w M ṽ 8( ) ,(45)where w A du ṽ du( 1) w M du ṽ du( ,A ( ) ,1) ,M ( ) ,and ṽ du( 1) is the approximate left eigenvector for the known eigenvalue ( 1) . One can see that if we replace thevector ṽ du( 1) with ṽ du( ) in equation (45) above, we regain the standard error estimator from (41). Thus the onlydifference between the error estimators (45) and (41) is the use of a left eigenvector at a previous aeroelastic parametervalue. The vectors w A and w M need only be computed once at each aeroelastic parameter increment, and therefore theterm f err9 is efficient to compute for all approximate eigenvalues.The second metric is th

aerodynamic model are indicated with the subscript a, whereas those for the structural models are presented with the subscript s. A. Structural Model For dynamic aeroelastic analysis, structural states are commonly expanded into a modal series consisting of the summation of free vibration mo

Related Documents:

MODELING STATE-SPACE AEROELASTIC SYSTEMS USING A SIMPLE MATRIX POLYNOMIAL APPROACH FOR THE UNSTEADY AERODYNAMICS RTO-AVT-154 09-3 dynamic pressure, not seen in the NASTRAN analysis. Now the concern led to aeroelastic modeling issues and issues with the flutter-analysis methods.

ses, such as motion analysis for flexible bodies. NX Nastran facilitates links to motion simulation tools, such as NX Motion, RecurDyn, Adams, SIMPACK and MATLAB to simplify simulation workflows. Aeroelastic analysis Aeroelastic analysis enables analysis of structural models in the presence of

Vibration theory. Steady and unsteady flows. Mathematical foundations of aeroelasticity, static and dynamic aeroelasticity. Linear unsteady aerodynamics, non-steady aerodynamics of lifting surfaces. Stall flutter. Aeroelastic problems in civil engineering structures. Aeroelastic problems of

calculate unsteady aerodynamics in frequency domain ignoring the bending effect of the deflected wing. And then, the aeroelastic stability analysis of the system under a given load condition is successively carried out. Comparing with the linear

The Aerodynamic Analysis and Aeroelastic Tailoring of a Forward-Swept Wing. (Under the direction of Dr. Charles E. Hall, Jr.) The use of forward-swept wings has aerodynamic benefits at high angles of attack and in supersonic regimes. These consist of reduction in wave drag, profile drag, and increased high angle of attack handling qualities.

Recent applications of higher-order spectral (HOS) methods to nonlinear aeroelastic phenomena are presented. Applications include the analysis of data from a simulated nonlinear pitch and plunge apparatus and from F-18 flight flutter tests. A MATLAB model of the Texas A&M University's Nonlinear Aeroelastic Testbed Apparatus (NATA) is used to

This paper offers an alternative approach to aeroelastic analysis, where the sensitivity of aerodynamic damping with respect to main flow and structural parameters is quantified through a surrogate-model-based investigation. The parameters are chosen based on previous studies and are represented by a uniform distribution within applicable .

asset management industry, that in the future will need to move these resources within its boundaries. handling compliance some Regulatory challenges In the past few years, regulatory compliance has constantly been at the top of asset manager’s agenda. Currently, the most debated regulation is the upcoming Market in Financial Instruments Directive (MiFID II), as it covers many areas of the .