+ All Categories
Home > Documents > Application of Multi-Input Volterra Theory to Nonlinear...

Application of Multi-Input Volterra Theory to Nonlinear...

Date post: 26-Apr-2018
Category:
Upload: trinhdieu
View: 226 times
Download: 1 times
Share this document with a friend
7
Application of Multi-Input Volterra Theory to Nonlinear Multi-Degree-of-Freedom Aerodynamic Systems Maciej Balajewicz, Fred Nitzsche, and Daniel Feszty Carleton University, Ottawa, Ontario K1S 5B6, Canada DOI: 10.2514/1.38964 This paper presents a reduced-order-modeling approach for nonlinear, multi-degree-of-freedom aerodynamic systems using multi-input Volterra theory. The method is applied to a two-dimensional, 2 degree-of-freedom transonic airfoil undergoing simultaneous forced pitch and heave harmonic oscillations. The so-called Volterra cross kernels are identied and shown to successfully model the aerodynamic nonlinearities associated with the simultaneous pitch and heave motions. The improvements in accuracy over previous approaches that effectively ignored the cross kernels by using superposition are demonstrated. I. Introduction B ECAUSE of the presence of aerodynamic nonlinearities in tran- sonic aeroelasticity, computational uid dynamics (CFD) has become the most reliable tool for its analysis. However, the large computational resources required for such high-delity analysis renders the approach undesirable, especially during the initial and conceptual design stages. As a result, there has been a great deal of interest in reduced-order models (ROMs) of the transonic nonlinear system [13]. A ROM is a simplied mathematical model that captures most of the physics of the more complex system under investigation. Among the many ROM methods available, the most popular include proper orthogonal decomposition, harmonic balance, and the Volterra series. Although research into the Volterra series as a ROM for the transonic aerodynamic system has been signicant [211], several critical issues remain unsolved. In his recent review paper, Silva [3] suggested that the application of the Volterra theory to multi-degree-of-freedom systems has been incomplete. He states that An important issue that needs to be addressed is the simultaneous excitation of multiple degrees of freedom to properly identify any nonlinear crosscoupling of the degrees of freedom. All previous applications of the Volterra series to multi-degree-of-freedom aerodynamic systems have been limited to the identication of aerodynamic nonlinearities resulting from individual perturbations of structural modes. Determination of total lift and moment for simultaneous motions required the superposition of the individual nonlinear responses. However, as suggested by Silva [3], the nonlinear nature of the system renders the principle of superposition invalid. This paper attempts to address this issue by proposing the multi- input Volterra series as a viable ROM method for nonlinear multi- degree-of-freedom aerodynamic systems. The multi-input Volterra series is well suited for this purpose and has been success- fully used by researchers in other disciplines [1216]. The multi- input Volterra series differs from the classical single-input Volterra series through its inclusion of Volterra cross kernels. These cross kernels capture the coupling dynamics between the degrees of freedom of the nonlinear system. The applicability of the multi-input Volterra series to nonlinear aerodynamic systems is illustrated by modeling the transonic, unsteady, two-dimensional, 2 DOF airfoil. II. Volterra Theory The Volterra theory of nonlinear systems is quite mature and several texts are available [17,18]. It was rst applied to nonlinear engineering problems by Wiener [19] and rst applied to subsonic and transonic aerodynamic systems by Tromp and Jenkins [4] and Silva [5], respectively. This section provides a brief summary of the Volterra theory of single- and multi-input nonlinear systems both in continuous and discrete-time domains. A. Single-Input Volterra Theory The output yt, of a continuous-time, causal, time-invariant, fading memory, nonlinear system , due to a single-input xt yt fxtg (1) can be modeled using the pth-order Volterra series yt X p i1 H i Z t 1 H 1 t x d Z t 1 Z t 1 H 2 t 1 ;t 2 x 1 x 2 d 1 d 2 . . . Z t 1 Z t 1 H p t 1 ; ... ;t p Y p i1 fx i d i g (2) where the pth-order Volterra operator H p , is dened as a p-fold convolution between the input xt and the pth-order Volterra kernel H p t; ... ;t. The identication of Volterra kernels is key to the synthesis of a Volterra ROM. However, analytical derivations of the Volterra kernels in continuous time are only possible if analytical, closed-form expressions of the inputoutput relationship of the nonlinear system are available. Unfortunately, many engineering applications of interest including aerodynamic applications lack such closed-form formulations and, instead, rely on numerical solutions of the nonlinear system . As a result, identication of the Volterra kernels involves the processing of discrete-time outputs due to Presented as Paper 2324 at the 49th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference and 16th AIAA/ ASME/AHS Adaptive Structures Conference, Schaumburg, IL, 710 April 2008; received 5 June 2008; revision received 10 September 2009; accepted for publication 10 September 2009. Copyright © 2009 by Maciej Balajewicz. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission. Copies of this paper may be made for personal or internal use, on condition that the copier pay the $10.00 per-copy fee to the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923; include the code 0001-1452/10 and $10.00 in correspondence with the CCC. Graduate Student, Department of Mechanical and Aerospace Engineer- ing; currently Graduate Student, Department of Mechanical Engineering and Materials Science, Duke University, Durham, North Carolina 27708; [email protected]. Student Member AIAA. Professor, Department of Mechanical and Aerospace Engineering. Senior Member AIAA. Assistant Professor, Department of Mechanical and Aerospace Engineering. Member AIAA. AIAA JOURNAL Vol. 48, No. 1, January 2010 56 Downloaded by STANFORD UNIVERSITY on July 21, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.38964
Transcript

Application of Multi-Input Volterra Theory to NonlinearMulti-Degree-of-Freedom Aerodynamic Systems

Maciej Balajewicz,∗ Fred Nitzsche,† and Daniel Feszty‡

Carleton University, Ottawa, Ontario K1S 5B6, Canada

DOI: 10.2514/1.38964

This paper presents a reduced-order-modeling approach for nonlinear, multi-degree-of-freedom aerodynamic

systems using multi-input Volterra theory. The method is applied to a two-dimensional, 2 degree-of-freedom

transonic airfoil undergoing simultaneous forced pitch and heave harmonic oscillations. The so-called Volterra

cross kernels are identified and shown to successfully model the aerodynamic nonlinearities associated with the

simultaneous pitch and heave motions. The improvements in accuracy over previous approaches that effectively

ignored the cross kernels by using superposition are demonstrated.

I. Introduction

B ECAUSE of the presence of aerodynamic nonlinearities in tran-sonic aeroelasticity, computational fluid dynamics (CFD) has

become the most reliable tool for its analysis. However, the largecomputational resources required for such high-fidelity analysisrenders the approach undesirable, especially during the initial andconceptual design stages. As a result, there has been a great deal ofinterest in reduced-order models (ROMs) of the transonic nonlinearsystem [1–3]. A ROM is a simplified mathematical model thatcaptures most of the physics of the more complex system underinvestigation. Among the many ROM methods available, the mostpopular include proper orthogonal decomposition, harmonicbalance, and the Volterra series. Although research into the Volterraseries as a ROM for the transonic aerodynamic system has beensignificant [2–11], several critical issues remain unsolved. In hisrecent review paper, Silva [3] suggested that the application ofthe Volterra theory to multi-degree-of-freedom systems has beenincomplete. He states that “An important issue that needs to beaddressed is the simultaneous excitation of multiple degrees offreedom to properly identify any nonlinear crosscoupling of thedegrees of freedom.” All previous applications of the Volterra seriesto multi-degree-of-freedom aerodynamic systems have been limitedto the identification of aerodynamic nonlinearities resulting fromindividual perturbations of structural modes. Determination of totallift andmoment for simultaneous motions required the superpositionof the individual nonlinear responses. However, as suggested bySilva [3], the nonlinear nature of the system renders the principle ofsuperposition invalid.

This paper attempts to address this issue by proposing the multi-input Volterra series as a viable ROM method for nonlinear multi-degree-of-freedom aerodynamic systems. The multi-input Volterraseries is well suited for this purpose and has been success-

fully used by researchers in other disciplines [12–16]. The multi-input Volterra series differs from the classical single-input Volterraseries through its inclusion of Volterra cross kernels. These crosskernels capture the coupling dynamics between the degrees offreedom of the nonlinear system. The applicability of the multi-inputVolterra series to nonlinear aerodynamic systems is illustrated bymodeling the transonic, unsteady, two-dimensional, 2 DOF airfoil.

II. Volterra Theory

The Volterra theory of nonlinear systems is quite mature andseveral texts are available [17,18]. It was first applied to nonlinearengineering problems by Wiener [19] and first applied to subsonicand transonic aerodynamic systems by Tromp and Jenkins [4] andSilva [5], respectively. This section provides a brief summary of theVolterra theory of single- and multi-input nonlinear systems both incontinuous and discrete-time domains.

A. Single-Input Volterra Theory

The output y�t�, of a continuous-time, causal, time-invariant,fading memory, nonlinear system �, due to a single-input x�t�

y�t� ��fx�t�g (1)

can be modeled using the pth-order Volterra series

y�t� �Xpi�1

Hi

�Zt

�1H1�t� ��x��� d�

�Zt

�1

Zt

�1H2�t � �1; t � �2�x��1�x��2� d�1 d�2

..

.

�Zt

�1� � �Zt

�1Hp�t � �1; . . . ; t � �p�

Ypi�1fx��i� d�ig (2)

where the pth-order Volterra operator Hp, is defined as a p-foldconvolution between the input x�t� and the pth-order Volterra kernelHp�t; . . . ; t�. The identification of Volterra kernels is key to thesynthesis of a Volterra ROM. However, analytical derivations of theVolterra kernels in continuous time are only possible if analytical,closed-form expressions of the input–output relationship of thenonlinear system � are available. Unfortunately, many engineeringapplications of interest including aerodynamic applications lack suchclosed-form formulations and, instead, rely on numerical solutionsof the nonlinear system �. As a result, identification of the Volterrakernels involves the processing of discrete-time outputs due to

Presented as Paper 2324 at the 49th AIAA/ASME/ASCE/AHS/ASCStructures, Structural Dynamics, and Materials Conference and 16th AIAA/ASME/AHS Adaptive Structures Conference, Schaumburg, IL, 7–10 April2008; received 5 June 2008; revision received 10 September 2009; acceptedfor publication 10 September 2009. Copyright © 2009 byMaciej Balajewicz.Published by the American Institute of Aeronautics and Astronautics, Inc.,with permission. Copies of this paper may be made for personal or internaluse, on condition that the copier pay the $10.00 per-copy fee to the CopyrightClearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923; includethe code 0001-1452/10 and $10.00 in correspondence with the CCC.

∗Graduate Student, Department of Mechanical and Aerospace Engineer-ing; currently Graduate Student, Department of Mechanical Engineering andMaterials Science, Duke University, Durham, North Carolina 27708;[email protected]. Student Member AIAA.

†Professor, Department ofMechanical and Aerospace Engineering. SeniorMember AIAA.

‡Assistant Professor, Department of Mechanical and AerospaceEngineering. Member AIAA.

AIAA JOURNALVol. 48, No. 1, January 2010

56

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

July

21,

201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.3

8964

specifically tailored training inputs. Consequently, the discrete-timeversion of the Volterra series using discrete-time Volterra operatorsand kernels is preferred. For a uniformly sampled, discrete-timerepresentation of the system

y�n� ��fx�n�g (3)

where for n� 0; 1; . . . ; N

x�n� � x�t�jt�n�T � x�n�T�y�n� � y�t�jt�n�T � y�n�T� (4)

The output y�n� of a nonlinear system� due to a single input x�n� canbe modeled using the pth-order, discrete-time, Volterra series

y�n� �Xpi�1

Hi

�Xnk�0

H1�n � k�x�k�

�Xnk1�0

Xnk2�0

H2�n � k1; n � k2�x�k1�x�k2�

..

.

�Xnk1�0� � �Xnkp�0

Hp�n � k1; � � � ; n � kp�Ypi�1x�ki� (5)

where the pth-order, discrete-time Volterra operatorHp is defined asa p-fold discrete-time convolution between the input x�n� and thepth-order, discrete-time Volterra kernelHp�n; . . . ; n�. Because of theexponentially increasing difficulty inherent in identifying higher-order, discrete-time Volterra kernels, most applications, includingaerodynamic ones, use a truncated, second-order (p� 2) Volterraseries

y�n� �Xnk�0

H1�n � k�x�k� �Xnk1�0

Xnk2�0

H2�n � k1; n � k2�x�k1�x�k2�

(6)

Unfortunately, such a low-ordered truncation of the Volterra seriesrestricts its ROM capability to weakly nonlinear systems only.Although precise and generally applicable definitions of weaknessare lacking, the general consensus is that a weakly nonlinear systemis one whose output only weakly diverges from the output predictedby linearized models.

B. Single-Input Volterra Kernel Identification

In this paper, Silva’s familiar impulse identificationmethod for thefirst- and second-order discrete-time Volterra kernels is used:

�H1�n� � 2�f��n�g � 12�f2��n�g (7)

XNk�0�2�2H2�n; n � k� ��f��n� � ��n � k�g

��f��n�g ��f��n � k�g� (8)

where ��n� is an impulse function of magnitude �

��n� ��� n� 0

0 n ≠ 0(9)

The identification of the first-order kernel H1�n� is straightforward.Only two outputs due to inputs ��n� and 2��n� are required. Theidentification of the second-order kernel H2�n; n� is more involvedbecause multiple outputs due to inputs ��n� and ��n � k� for k�0; 1; . . . ; N must be computed. Because of symmetry, for all k

H2�n; n � k� �H2�n � k; n� (10)

C. Multi-Input Volterra Theory

The output y�t� of a continuous-time, causal, time-invariant,fading memory, nonlinear system � due to m inputs

y�t� ��fx1�t�; x2�t�; . . . ; xm�t�g (11)

can be modeled using the pth-order, multi-input Volterra series

y�t� �Xpi�1

Hmi

�Xmj�1

�Zt

�1Hj

1�t� ��xj���d��

�Xmj1�1

Xmj2�1

�Zt

�1

Zt

�1Hj1j2

2 �t� �1; t� �2�xj1��1�xj2 ��2�d�1 d�2�

..

.

�Xmj1�1� � �Xmjp�1

�Zt

�1� � �Zt

�1Hj1 ���jpp �t� �1; . . . ; t� �p�

Ypi�1fxji ��i�d�ig

�(12)

where the pth-order, multi-input Volterra operatorHmp is defined as a

mp-fold summation of p-fold convolution integrals between thevarious combinations of inputs x1�t�; x2�t�; . . . ; xm�t�, and the pth-

order,multi-input Volterra kernelHj1 ���jpp �t; . . . ; t�. Notice the appear-

ance of superscripts on the pth-order, multi-input Volterra kernel

Hj1 ���jpp �t; . . . ; t�. These superscripts identify to which inputs the

kernel corresponds. For example, a third-order kernel H5123 �t; t; t�

corresponds to inputs x5�t�, x1�t�, and x2�t�. Volterra kernels

Hj1 ���jpp �t; . . . ; t�, where j1 � j2 � � � � � jp, are called Volterra direct

kernels. Volterra kernels with superscripts that do not match arecalled Volterra cross kernels. The presence of these Volterra crosskernels differentiates the multi-input Volterra series from the single-input Volterra series summarized in Sec. II.A.

For a uniformly sampled, discrete-time representation of thesystem �

y�n� ��fx1�n�; x2�n�; . . . ; xm�n�g (13)

where n� 0; 1; . . . ; N

Xmi�1fxi�n� � xi�t�jt�n�T � xi�n�T�g

y�n� � y�t�jt�n�T � y�n�T� (14)

The output y�n� of a nonlinear system � due to m inputs can bemodeled using the pth-order, discrete-time, multi-input Volterraseries

y�n� �Xpi�1

Hmi

�Xmj�1

�Xnk�0

Hj1�n � k�xj�k�

�Xmj1�1

Xmj2�1

�Xnk1�0

Xnk1�0

Hj1j22 �n � k1; n � k2�xj1 �k1�xj2 �k2�

..

.

�Xmj1�1� � �Xmjp�1

�Xnk1�0� � �Xnkp�0

Hj1 ���jpp �n � k1; � � � ; n � kp�

Ypi�1fxji �ki�g

�(15)

where the pth-order, discrete-time, multi-input Volterra operatorHmp is defined as a mp-fold summation of p-fold discrete-time

BALAJEWICZ, NITZSCHE, AND FESZTY 57

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

July

21,

201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.3

8964

convolutions between the various combinations of inputs x1�n�;x2�n�; . . . ; xm�n� and the pth-order, multi-input, discrete-time

Volterra kernel Hj1 ���jpp �n; . . . ; n�. The difficulties associated with

the identification of higher-order, multi-input, discrete-time Volterradirect and cross kernels are identical to those associated with theidentification of higher-order, single-input, discrete-time Volterrakernels. In other words, only a second-order (p� 2), multi-inputVolterra series is practical

y�n� �Xmj�1

�Xnk�0

Hj1�n � ��xj�k�

�Xmj1�1

Xmj2�1

�Xnk1�0

Xnk1�0

Hj1j22 �n � k1; n � k2�xj1 �k1�xj2 �k2�

�(16)

D. Multi-Input Volterra Kernel Identification

As in the previous section dealing with the single-input Volterraseries, we use the impulse identification method for the first- andsecond-order direct and cross kernels:

Xmj�1

��jH

j1�n� � 2�f�j�n�g �

1

2�f2�j�n�g

�(17)

Xmj1�1

Xmj2�1

�XNk�0�2�j1�j2H

j1j22 �n; n � k� ��f�j1 �n� � �j2 �n � k�g

��f�j1 �n�g ��f�j2 �n � k�g��

(18)

where �j�n� is an impulse function corresponding to the jth input

�j�n� ���j n� 0

0 n ≠ 0(19)

Because of symmetry, for all k

Hj1j22 �n; n � k� �H

j1j22 �n � k; n� for j1 � j2

Hj1j22 �n; n � k� �H

j2j12 �n � k; n� for j1 ≠ j2 (20)

III. Multi-Input Volterra Reduced-Order Modelof 2 Degree-of-Freedom Airfoil

The 2 DOF oscillations in both pitch � and heave h fullycharacterize the unsteady motion of a two-dimensional airfoil ofchord c immersed in transonic flow, as illustrated in Fig. 1a. Theaerodynamics of such an airfoil are described using the coefficient ofliftCL andmomentCM, as shown in Fig. 1b. However, for the sake ofbrevity, in this paper we limit our discussion to the aerodynamicmoment output only. Unlike the pitch degree of freedom, only a

heave velocity _h is capable of producing aerodynamic loading.Therefore, from a Volterra series ROM point of view, it is more

appropriate to use _h, instead of h, as one of the inputs to the aero-dynamic system:

CM�n� � CFDf��n�; _h�n�g (21)

To capture the nonlinear effects of the transonic flow regime, thetraditional Volterra ROM approach has been to superimpose twosecond-order Volterra ROMs, one per each degree of freedom:

CM�n� �Xnk�0

H�1 �n � k���k� �

Xnk�0

H_h1 �n � k� _h�k�

�Xnk1�0

Xnk2�0

H��2 �n � k1; n � k2���k1���k2�

�Xnk1�0

Xnk2�0

H_h _h2 �n � k1; n � k2� _h�k1� _h�k2� (22)

where H�1 , H

��2 and H

_h1 , H

_h _h2 are the first- and second-order direct

Volterra kernels corresponding to the pitch and heave degrees offreedom.However, as suggested by Silva [3], the nonlinearities of thesystem render the superposition principle invalid. The multi-inputVolterra series is a more appropriate ROM for this system due to itsinclusion of Volterra cross kernels, which are capable of modelingthe nonlinearities associated with the simultaneous perturbation ofthe pitch and heave degrees of freedom. The second-order, multi-input Volterra series ROM of this system is formed by setting j1 � �and j2 � _h in Eq. (16):

CM�n� �Xnk�0

H�1 �n � k���k� �

Xnk�0

H_h1 �n � k� _h�k�

�Xnk1�0

Xnk2�0

H��2 �n � k1; n � k2���k1���k2�

�Xnk1�0

Xnk2�0

H_h _h2 �n � k1; n � k2� _h�k1� _h�k2�

�Xnk1�0

Xnk2�0

H� _h2 �n � k1; n � k2���k1� _h�k2�

�Xnk1�0

Xnk2�0

H_h�2 �n � k1; n � k2� _h�k1���k2� (23)

where H� _h2 and H

_h�2 are the second-order Volterra cross kernels. All

the required first- and second-order Volterra direct and cross kernelsin Eq. (23) are identified using the impulse identification methodsummarized in Sec. II.D:

��H�1 �n� � 2CFDf���n�g � 1

2CFDf2���n�g

� _hH_h1 �n� � 2CFDf� _h�n�g � 1

2CFDf2� _h�n�g

(24)

XNk�0�2�2�H��

2 �n; n � k� � CFDf���n� � ���n � k�g

� CFDf���n�g � CFDf���n � k�g�XNk�0�2�2_hH

_h _h2 �n; n � k� � CFDf� _h�n� � � _h�n � k�g

� CFDf� _h�n�g � CFDf� _h�n � k�g�XNk�0�2��� _hH� _h

2 �n; n � k� � CFDf���n� � � _h�n � k�g

� CFDf���n�g � CFDf� _h�n � k�g�XNk�0�2� _h��H

_h�2 �n; n � k� � CFDf� _h�n� � ���n � k�g

� CFDf� _h�n�g � CFDf���n � k�g� (25)

where ���n� and � _h�n� are impulse functions corresponding to the

pitch � and heave velocity _h inputs:

���n� ���� n� 0

0 n ≠ 0� _h�n� �

�� _h n� 0

0 n ≠ 0(26)

Fig. 1 The 2 DOF, two-dimensional airfoil system.

58 BALAJEWICZ, NITZSCHE, AND FESZTY

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

July

21,

201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.3

8964

IV. Description of Test Case

To demonstrate the applicability of the multi-input Volterra seriesas a ROM method for nonlinear, multi-degree-of-freedom aero-dynamic systems, we chose to model a symmetric NACA 0012airfoil oscillating about a nonzero static angle of attack:

�� �0 � �� sin�2k��� _h� �_h sin�2k _h�� (27)

where k� !c=2U1, � �U1t=c. The static and dynamic pitchamplitudes, �0 and ��, have units of degrees, whereas the dynamic

heave velocity amplitude�_h is dimensionless because _h�

d�h=c�=d�. When assigning values to the parameters in Eq. (27),several requirements must be satisfied:

1) For the prescribed motion, reliable experimental data shouldexist to validate the CFD output.

2) The nonlinear lift and moment response due to these motionsmust be weakly nonlinear.

To the authors’ knowledge, no experimental data for simultan-eously forced pitch and heave oscillations at transonic conditionsexist. However, the experimental data are numerous for forcedpitch motions alone. An often-cited experimental data set of lift andmoment values for forced pitchmotions at transonic conditions is the1982 AGARD Compendium of Unsteady Aerodynamic Measure-ments [20]. This AGARD report introduces the so-called CT2 testcase:

�� 3:16 deg�4:59 deg sin�2k��� (28)

at a Mach number M of 0.6 and a pitch reduced frequency ofk� � 0:0811. The aerodynamics of this test case are characterizedby the formation of a strong and highly dynamic shock waveexperiencing Tijdeman and Seebass’s [21] type-B shock motion.Unfortunately, strong and highly dynamic shock waves are knownto produce highly nonlinear lift and moment response. Because asecond-order Volterra ROM can only handle weak nonlinearities,the dynamic pitch motion must be decreased so that weaker, less-dynamic shock waves are formed. Observing the pressure contoursfrom the AGARD reports, this condition is satisfied at dynamic pitchamplitudes of approximately 1 deg.

Because the aim of this paper is to present a method of modelingmulti-degree-of-freedom aerodynamic systems using themulti-inputVolterra series, the airfoil motion must include heave. The maindifference between the application of the multi-input Volterra seriesover the single-input Volterra series is the identification of crosskernels. Becausewewish to focus on these cross kernels, it is helpfulto select a dynamic heavevelocity amplitude that would yield a heavevelocity direct kernel approximately equal in magnitude to the pitchdirect kernel. After several systematic CFD runs, it was found that adimensionless heave velocity amplitude of 0.018 would achieve thisrequirement. Hence, the following 2 DOF motion was chosen:

�� 3:16 deg�1 deg sin�2k��� _h� 0:018 sin�2k _h�� (29)

whereM� 0:6, k� � 0:0811, and kh � 0:8k�.

V. CFD Using the Carleton Multiblock Solver

All CFD results presented in this paper were carried out usingthe Carleton multiblock (CMB) CFD code. The CMB code is aderivative of a code originally developed at the University ofGlasgow, specifically tailored for transonic, time-marching aero-elastic analysis. The aerodynamics of the airfoil were modeled usingthe inviscid Euler equations. For further details, refer to Dubuc et al.[22] and Badcock et al. [23].

The NACA 0012 airfoil domain was discretized using a C-type180 33 Euler grid with 130 nodes on the airfoil. The surface nodeswere at a distance of approximately 0:001coff the airfoil surface. Themesh extended into the far field approximately 10c in all directions.The unsteady solutions were solved using a dimensionless time step�� � 1:96, which corresponds to 20 time steps per period of pitchoscillation. This choice of mesh and time step was based on several

mesh refinement studies carried out by Dubuc et al. [22], whichshowed that no significant accuracy improvements are gained athigher spatial or temporal discretizations.

Figure 2 compares the experimental and CFD outputs of theAGARD test case as described by Eq. (28); very good agreement wasobtained. Errors are likely associated with the neglect of viscousforces and uncertainties in the experimental data [22,23].

VI. Results and Discussion

A. Volterra Kernel Identification

As stated in the theoretical derivation of the Volterra series inSec. III, the first- and second-order kernels identified using Eqs. (24)and (25) are illustrated in Figs. 3 and 4, respectively. From sensitivitystudies summarized in Secs. VI.C and VI.D, both the first- andsecond-order kernels are identified using �� � 2, �� � 0:5 deg,and � _h � 0:009. On a single Intel Pentium 3.2 GHz desktop ma-chine running RedHat 2.6.9 with 1 GB of RAM, the identification ofa single first-order and a single second-order kernel requires approxi-mately 51 s (0.85min) and 550 s (9.1min) of CPU time, respectively.Because of the quickly decaying nature of the second-order kernels,only the first 10 terms are identified, that is, in Eq. (25), N � 10.

B. Multi-Input Volterra Reduced-Order Model of Transonic Airfoil

Figures 5a and 5b illustrate the CFD and Volterra ROM momentcoefficient outputs due to the individual application of the pitch andheave inputs of Eq. (31), respectively. It should be noted that themoment coefficient due to the steady pitch of 3.16 deg has beensubtracted from the unsteady results presented in this section. In bothcases, the Volterra ROM, which includes both the first- and second-order kernels, performs better at modeling the nonlinear CFD output.

Figure 6 illustrates the CFD moment coefficient output due to thesimultaneous application of pitch and heave inputsCFDf�� hg andthe superposition of outputs due to individual application of the pitchand heave inputsCFDf�g � CFDfhg.We can clearly verify that, dueto the nonlinear nature of the aerodynamic system, the superpositionof the individual outputs does not accurately predict the output due tothe simultaneous application of inputs. Note the clear presence of a

Fig. 2 CFD validation of AGARD test case.

Fig. 3 First-order Volterra kernels of the transonic airfoil.

BALAJEWICZ, NITZSCHE, AND FESZTY 59

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

July

21,

201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.3

8964

beat frequency in the aerodynamic moment coefficient resultingfrom the input frequency ratio being nearly equal to one. Figure 7compares the simultaneous CFDf�� hg and superimposedCFDf�g � CFDfhg outputs with a Volterra ROM that only includes

the direct kernels H�1 , H

_h1 , H

��2 , and H

_h _h2 . As expected, the

superimposed output CFDf�g � CFDfhg is well predicted, but thesimultaneous output CFDf�� hg is not.

Figure 8 shows how the inclusion of cross kernels H_h�2 and H� _h

2

significantly improves the accuracy of theVolterra ROM inmodelingthe simultaneous response.

We can quantify the modeling performance of the Volterra ROMsusing the L2 relative error norm, defined as

%Error�

��������������������������������������������������������������������CFDf�� hg � Volterra ROM�2

CFDf�� hg2

s 100 (30)

Error norms for the variousVolterra ROMs covered in this section aresummarized in Table 1.As expected, the inclusion of the cross kernelssignificantly increases the accuracy of the Volterra ROM.

Figure 9 shows the frequency content of the CFDf�� hg outputand the variousVolterra kernels thatmake up theVolterraROMs. The

first-order kernelsH�1 andH

_h1 capture the fundamental, linear output

frequencies k=k� � 1 and k=k� � 0:8. The second-order, direct

kernels H��2 and H

_h _h2 capture the first three integer harmonics

k=k� � 2 (i.e., k=k� � 2 1), k=k� � 1:6 (i.e., k=k� � 2 0:8), and

Fig. 4 Second-order Volterra kernels of the transonic airfoil.

Fig. 5 Moment coefficient outputs due to individual application of

pitch and heave motions.

Fig. 6 Superimposed and simultaneous pitch and heave motion

outputs.

Fig. 7 Volterra ROM using direct kernels only against superimposed

and simultaneous pitch and heave motion CFD output.

Fig. 8 Volterra ROM using direct kernels only and a Volterra ROMusing both direct and cross kernels against simultaneous pitch and heave

motion CFD output.

Table 1 Volterra ROM modeling error norms

ROM % error

Volterra fH�1 ; H

_h1 g 30.6

VolterrafH�1 ; H

_h1 ; H

��2 ; H

_h _h2 g 21.7

Volterra fH�1 ; H

_h1 ; H

��2 ; H

_h _h2 ; H

� _h2 ; H

_h�2 g 12.1

60 BALAJEWICZ, NITZSCHE, AND FESZTY

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

July

21,

201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.3

8964

k=k� � 0 (nonlinear contribution to the static response), while the

cross kernels H� _h2 and H

_h�2 capture the first two intermodulation

harmonics k=k� � 1:8 (i.e., k=k� � 1� 0:8) and k=k� � 0:2 (i.e.,k=k� � 1 � 0:8).

C. Effect of Time Step and Impulse Magnitude on Kernel

Identification Accuracy

As demonstrated by Raveh [8] and more recently by Grewal andZimcik [24], the impulse identification method of Eqs. (24) and (25)can be very sensitive to the choice of time step and impulseamplitude. Hence, to ensure accurate Volterra kernel identification, asensitivity analysis with respect to time step and impulse amplitudewas performed. However, due to time constraints, only pitch impulseamplitude �� sensitivity was analyzed. Figure 10a shows error normsfor a first- and second-order Volterra ROM whose kernels werecalculated using�� � 2 and four different pitch impulse amplitudes:�� � 0:1, 0.5, 1, and 2.5 deg. Here, the L2 relative error norm isdefined as

%Error�

�����������������������������������������������������������CFDf�g � Volterra ROM�2

CFDf�g2

s 100 (31)

where CFDf�g is the CFD output due to a pitch input as specifiedby Eq. (29); �� 3:16 deg�1 deg sin�2k���, where M � 0:6,k� � 0:0811, and �� � 1:96. Observing Fig. 10a, it is clear thatoptimal performance is achieved with �� � 0:5 deg.

Figure 10b shows error norms for afirst- and second-orderVolterraROM whose kernels were calculated using �� � 0:5 deg and fourdifferent time steps: �� � 0:5, 1, 2 and 4. Observing Fig. 10b, it isclear that optimal performance is achieved with �� � 2.

Noting that the optimal pitch impulse amplitude of �� � 0:5 degequals half the dynamic pitch amplitude ��, it was assumed that a

heave velocity amplitude � _h � 0:009� 1=2 � �_h would yield similaroptimal performance. However, due to time constraints, this was notverified.

D. Effect of Pitch and Heave Frequency Separation

Figure 11 illustratesVolterraROMerror norms for inputs specifiedby Eq. (29):

�� 3:16 deg�1 deg sin�2k��� _h� 0:018 sin�2k _h�� (32)

where M� 0:6 and k� � 0:0811 for four different heave reducedfrequencies: k _h � 0:2k�, 0:4k�, 0:8k� and 1:6k�. TheVolterra ROMswere calculated using the optimal parameters determined earlier:�� � 2, �� � 0:5 deg and � _h � 0:009. The L2 relative error normsare defined as

%Error�

��������������������������������������������������������������������CFDf�� hg � Volterra ROM�2

CFDf�� hg2

s 100 (33)

In all cases, inclusion of Volterra cross kernels H� _h2 and H

_h�2

significantly improves modeling performance.

E. Curse of Dimensionality

The total CPU time tID required to identify a second-order Volterraseries ROM of an N degree-of-freedom system equals

tID � N � tH1� N2 � tH2

(34)

where tH1and tH2

are CPU times required to identify a singlefirst-order and a single second-order Volterra kernel, respectively.Because, in general, tH2

tH1, the total identification time can

becomevery high and render theVolterra ROMapproach impracticalfor higher degree-of-freedom systems.

However, its possible that under certain circumstances thisproblem can be avoided. For example, consider the problem oftransonic mode-coupling instability, flutter, of a complete aircraftconfiguration governed by N modal degrees of freedom. It is wellknown that, in many cases, the onset of flutter, and the modesinvolved in the instability, can be predicted using linear (dynamicallylinear) methods [1]. For example, linear, first-order Volterra ROMshave been demonstrated to successfully predict transonic flutter ofcomplete aircraft configurations [3]. Identification of a second-orderVolterra ROM is only necessary when the evolution of the instabilityis of interest, for example, determination of limit cycle oscillationamplitude and frequency. For such a scenario, it would only benecessary to identify second-order kernels corresponding to the twomodes involved in the instability.Hence, the total computational timefor such a model would equal

Fig. 9 Frequency content of CFD and individual Volterra kernels.

Fig. 10 Sensitivity analysis of Volterra ROM accuracy with respect to

time step and impulse magnitude. Fig. 11 Volterra ROM modeling error norms vs heave frequency.

BALAJEWICZ, NITZSCHE, AND FESZTY 61

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

July

21,

201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.3

8964

tID � N � tH1� 22 � tH2

(35)

where only four second-order kernels, two direct and two crosskernels, are necessary. Identification of N first-order kernels isrelatively trivial.

VII. Conclusions

The multi-input Volterra series differs from the classical Volterraseries through its inclusion of cross kernels. These cross kernelscapture the intermodulation harmonics when multiple degreesof freedom of a nonlinear system are perturbed simultaneously.The applicability of the multi-input Volterra series to nonlinear aero-dynamic systems was demonstrated using the transonic, unsteady,two-dimensional, 2 DOF NACA 0012 airfoil. For the specificAGARD test case analyzed, the identified cross kernels significantlyincreased the modeling accuracy of the Volterra ROM.

References

[1] Dowell, E., Edwards, J., and Strganac, T., “Nonlinear Aeroelasticity,”Journal of Aircraft, Vol. 40, No. 5, 2003, pp. 857–874.doi:10.2514/2.6876

[2] Lucia, D., Beran, P., and Silva, W., “Reduced-Order Modeling: NewApproached for Computational Physics,” Progress in Aerospace

Sciences, Vol. 40, 2004, pp. 51–117.doi:10.1016/j.paerosci.2003.12.001

[3] Silva, W., “Identification of Nonlinear Aeroelastic Systems Based onthe Volterra Theory: Progress and Opportunity,” Nonlinear Dynamics,Vol. 39, No. 1–2, 2005, pp. 25–62.doi:10.1007/s11071-005-1907-z

[4] Tromp, J. C., and Jenkins, J. E., “A Volterra Kernel IdentificationScheme Applied to Aerodynamic Reactions,” AIAA Paper 90-2803,Aug. 1990.

[5] Silva, W., “Application of Nonlinear Systems Theory to TransonicUnsteadyAerodynamicResponses,” Journal of Aircraft, Vol. 30,No. 5,1993, pp. 660–668.doi:10.2514/3.46395

[6] Silva, W., “Reduced-order Models Based on Linear and NonlinearAerodynamic Impulse Responses,” AIAA Paper 1999-1262, 1999.

[7] Raveh, D. E., Levy, Y., and Karpel, M., “Aircraft Aeroelastic Analysisand Design Using CFD-Based Unsteady Loads,” AIAA Paper 2000-1325, 2000.

[8] Raveh, D. E., “Reduced Order Models for Nonlinear UnsteadyAerodynamics,” AIAA Journal, Vol. 39, No. 8, Aug. 2001, pp. 1417–1429.doi:10.2514/2.1473

[9] Silva, W., Beran, P. S., Cesnik, C. E. S., Kurdila, R. E., Prazenica, A.,Librescu, R. J., Marzocca, P., and Raveh, D., “Reduced-OrderModeling: Cooperative Research and Development at the NASALangley Research Center,” International Forum on Aeroelasticity andStructural Dynamics Paper 2001-008, 2001.

[10] Lucia, D. J., Beran, P. S., and Silva, W. A., “Aeroelastic SystemDevelopment Using Proper Orthogonal Decomposition and Volterra

Theory,” AIAA Paper 2003-1922, 2003.[11] Silva, W., Hong, M. S., Bartels, R. E., Piatak, D. J., and Scott, R. C.,

“Identification of Computational and Experimental Reduced-OrderModels,” International Forum on Aeroelasticity and StructuralDynamics Paper 2003-US-39, 2003.

[12] Marmarelis, P. Z., and Naka, K. I., “Identification of Multi-InputBiological Systems,” IEEE Transactions on Biomedical Engineering,Vol. BME-21, March 1974, pp. 88–101.doi:10.1109/TBME.1974.324293

[13] Grzymacz, N. M., and Koch, C., “Functional Properties of Models forDirection Selectivity in the Retina,” Synapse, Vol. 1, No. 5, 1987,pp. 417–434.doi:10.1002/syn.890010506

[14] Worden, K., Manson, G., and Tomlinson, G. R., “A Harmonic ProbingAlgorithm for the Multi-Input Volterra Series,” Journal of Sound and

Vibration, Vol. 201, No. 1, 1997, pp. 67–84.doi:10.1006/jsvi.1996.0746

[15] Agarossi, L., Agarossi, L., Bellini, S., Canella, A., andMigliorati, P., “AVolterra Model for the High Density Optical Disc,” Proc. IEEE

International Conference on Acoustics, Speech and Signal Processing,edited by S. Bellini, Vol. 3, IEEE Publications, Piscataway, NJ, 1998,pp. 1605–1608.

[16] Chatterjee, A., and Vyas, N. S., “Nonlinear Parameter Estimation inMulti-Degree-of-Freedom Systems UsingMulti-Input Volterra Series,”Mechanical Systems and Signal Processing, Vol. 18, 2004, pp. 457–489.doi:10.1016/S0888-3270(03)00016-5

[17] Schetzen,M., The Volterra andWiener Theories of Nonlinear Systems,Wiley, New York, 1980.

[18] Rugh,W. J.,Nonlinear System Theory. The Volterra/Wiener Approach,John Hopkins Univ. Press, Baltimore, MD, 1981.

[19] Wiener, N., “Response of a Nonlinear Device to Noise,” RadiationLaboratory, Massachusetts Inst. of Technology, Rept. 129 Cambridge,MA, 1942.

[20] “Compendium of Unsteady Aerodynamic Measurements,” AGARDTech. Rept. R-702, 1982.

[21] Tijdeman, H., and Seebass, R., “Transonic Flow Past OscillatingAirfoils,” Annual Review of Fluid Mechanics, Vol. 12, 1980, pp. 181–222.doi:10.1146/annurev.fl.12.010180.001145

[22] Dubuc, L., Cantariti, F.,Woodgate,M., Gribben, B., Badcock,K. J., andRichards, B. E., “Solution of the Euler Unsteady Equations UsingDeforming Grids,” Univ. of Glasgow, Aero. Rept. 9704, 1997.

[23] Badcock, K. J., Richards, B. E., and Woodgate, M. A., “Elements ofComputational Fluid Dynamics on Block Structured Grids UsingImplicit Solvers,”Progress inAerospace Sciences, Vol. 36,No. 5, 2000,pp. 351–392.doi:10.1016/S0376-0421(00)00005-1

[24] Grewal, A., and Zimcik, D., “Application of Reduced Order Modellingto Fluid-Structure Interaction Analysis,” AIAA Paper 2008–5961,Sept. 2008.

J. CooperAssociate Editor

62 BALAJEWICZ, NITZSCHE, AND FESZTY

Dow

nloa

ded

by S

TA

NFO

RD

UN

IVE

RSI

TY

on

July

21,

201

3 | h

ttp://

arc.

aiaa

.org

| D

OI:

10.

2514

/1.3

8964


Recommended