140 91 51MB
English Pages 408 [396] Year 2011
Conference Proceedings of the Society for Experimental Mechanics Series
For other titles published in this series, go to www.springer.com/series/8922
Tom Proulx Editor
Civil Engineering Topics, Volume 4 Proceedings of the 29th IMAC, A Conference on Structural Dynamics, 2011
Editor Tom Proulx Society for Experimental Mechanics, Inc. 7 School Street Bethel, CT 06801-1405 USA [email protected]
ISSN 2191-5644 e-ISSN 2191-5652 ISBN 978-1-4419-9315-1 e-ISBN 978-1-4419-9316-8 DOI 10.1007/978-1-4419-9316-8 Springer New York Dordrecht Heidelberg London Library of Congress Control Number: 2011923969 © The Society for Experimental Mechanics, Inc. 2011 All rights reserved. This work may not be translated or copied in whole or in part without the written permission of the publisher (Springer Science+Business Media, LLC, 233 Spring Street, New York, NY 10013, USA), except for brief excerpts in connection with reviews or scholarly analysis. Use in connection with any form of information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or hereafter developed is forbidden. The use in this publication of trade names, trademarks, service marks, and similar terms, even if they are not identified as such, is not to be taken as an expression of opinion as to whether or not they are subject to proprietary rights. Printed on acid-free paper Springer is part of Springer Science+Business Media (www.springer.com)
Preface
Civil Structures Topics represents one of six clusters of technical papers presented at the 29th IMAC, A Conference and Exposition on Structural Dynamics, 2011 organized by the Society for Experimental Mechanics, and held in Jacksonville, Florida, January 31 - February 3, 2011. The full proceedings also include volumes on Advanced Aerospace Applications; Modal Analysis, Linking Models and Experiments, Rotating Machinery, Structural Health Monitoring, Shock and Vibration; and Sensors, Instrumentation, and Special Topics. Each collection presents early findings from experimental and computational investigations on an important area within Structural Dynamics. The current volume on Civil Structures Topics includes studies on Human Loading, Vibration Control, Bridges, Modal Parameter Identification, System Identification, Modal Testing, Experimental Techniques, and Processing Modal Data. Understanding the dynamic response of large civil structures improves design and safety, extends life, and reduces maintenance. IMAC has become a principle technical venue for dissemination of the latest techniques devoted to field testing of civil structural systems and components, processing of response data and identification of dynamic structural properties, calibration and validation of numerical structural models, and assessment of structural condition based on dynamic properties. The organizers would like to thank the authors, presenters, session organizers and session chairs for their participation in this track.
Bethel, Connecticut
Dr. Thomas Proulx Society for Experimental Mechanics, Inc
Contents
1
Multi-dimensional Calibration of Impact Dynamic Models L.G. Horta, M.C. Reaves, M.S. Annett, K.E. Jackson, NASA Langley Research Center
2
Dynamic Analysis of Railway Bridges by Means of the Spectral Method G. Catania, S. Sorrentino, University of Bologna
3
Advanced Operational Modal Analysis Methods for Linear Time Periodic System Identification M.S. Allen, University of Wisconsin-Madison; S. Chauhan, Brüel & Kjær Sound & Vibration Measurement A/S; M.H. Hansen, Risø National Laboratory
1 21
31
4
Normalization of Complex Modes from Mass Perturbations D.-W. Seo, F. Moghim, D. Bernal, Northeastern University
45
5
Numerical Continuation of Periodic Orbits for Harmonically Forced Nonlinear Systems M.W. Sracic, M.S. Allen, University of Wisconsin-Madison
51
6
Identification of the Tensile Force in Tie-rods of Historical Constructions H.T.M. Luong, L.F. Ramos, R. Aguilar, University of Minho
71
7
Data Fusion for System Identification of the Humber Bridge M. Döhler, INRIA; B. Jaishi, The University of Sheffield; L. Mevel, INRIA; J.M.W. Brownjohn, The University of Sheffield
83
8
Human-induced Vibration Propagation on a Composite Floor System O.F. Tigli, McNamara/Salvia Inc. Consulting Engineers
99
9
Mitigation of Low-velocity, Wind-induced Vibration of an Architectural Spire O.F. Tigli, McNamara/Salvia Inc. Consulting Engineers; L. Caracoglia, Northeastern University
109
10
Closed-loop System Identification of Alstom 3MW Wind Turbine C.E. Carcangiu, I.F. Balaguer, ALSTOM Wind; S. Kanev, ECN Wind Energy; M. Rossetti, ALSTOM Wind
121
11
Reduced Order Modal-space Wireless Control of Civil Engineering Structures R.A. Swartz, Michigan Technological University
129
viii
12
Wind Turbine Structural Damping Control for Tower Load Reduction A. Rodríguez T., C.E. Carcangiu, I. Pineda, Alstom Wind; T. Fischer, B. Kuhnle, M. Scheu, Universität Stuttgart; M. Martin, Alstom Wind
13
Identification of Dynamic Properties of Open-deck Viaducts Under Passing Train Loads K. Matusoka, K. Kaito, Osaka University; T. Watanabe, M. Sogabe, Railway Technical Research Institute
141
155
14
Risk Assessment of Dolphin Protected Bridge Pier Considering Collision Point Analysis H.-J. Kim, J. Lim, W. Park, H.-M. Koh, Seoul National University
163
15
OMAX Testing of a Steel Bowstring Footbridge E. Reynders, D. Degrauwe, M. Schevenels, G. De Roeck, Katholieke Universiteit Leuven; P. Van den Broeck, KaHo Sint-Lieven; K. Deckers, P. Guillaume, Vrije Universiteit Brussel
173
16
Bridge Pier Scour Measurement by Means of Bragg Grating Arrays: Laboratory Tests and Field Applications F. Ballio, A. Cigada, G. Crotti, F. Inzoli, S. Manzoni, Politecnico di Milano
183
Numerical Investigation on the Measurement Uncertainty in Operational Modal Analysis of a Civil Structure A. Cattaneo, S. Manzoni, M. Vanali, Politecnico di Milano
191
Measurement and Application of Bouncing and Jumping Loads Using Motion Tracking Technology V. Racic, J.M.W. Brownjohn, A. Pavic, The University of Sheffield
201
17
18
19
Operational Modal Analysis for Guangzhou New TV Tower Y. Niu, P. Kraemer, C.-P. Fritzen, University of Siegen
20
On the use of Shape Memory Alloys Dampers to Reduce the Vibration Amplitudes of Civil Engineering Cables L. Dieng, LCPC; V. Torra, CIRG-DFA-ETSECCPB; P. Pilvin, Université de Bretagne
21
Dynamic Behavior of ‘Palazzo Lombardia’ Helisurface, Comparison Between Operational and Experimental Modal Analysis G. Busca, A. Cigada, Politecnico di Milano; E. Mola, ECSD S.r.l.; F. Mola, M. Vanali, Politecnico di Milano
211
221
235
22
Human-structure Interaction: Effects of Crowd Characteristics K.A. Salyards, R.J. Firman, III, Bucknell University
247
23
Delamination Detection in Concrete Plates Using Output-only Vibration Measurements S. Xing, M.W. Halling, P.J. Barr, Utah State University
255
24
Semi-active Control of Staircase Vibration Under Human Excitation R. Philp, P. Reynolds, D.S. Nyawako, The University of Sheffield
263
ix
25
A System Identification Based Approach for Estimating Modal Frequencies of Bridges Under Environmental Influence H. Nandan, Virginia Polytechnic Institute and State University; M.P. Singh, National Science Foundation Arlington
273
26
Feature Assimilation in Structural Health Monitoring Applications Saurabhu Prabhu, Jordan Supler, Sez Atamturktur, Clemson University
27
Decentralized Modal Analysis and System Identification Using Embedded Markov Parameter Extraction Within Distributed Wireless Sensor Networks J. Kim, J.P. Lynch, University of Michigan
297
Comparison Study of Output-only Subspace and Frequency-domain Methods for System Identification of Base Excited Civil Engineering Structures J. Kim, J.P. Lynch, University of Michigan
305
28
285
29
Footbridge Response Predictions and Their Sensitivity to Stochastic Load Assumptions L. Pedersen, C. Frier, Aalborg University
313
30
Modal Parameter Variations due to Earthquakes of Different Intensities R.P. Carreño, R.L. Boroschek, University of Chile
321
31
Operational Modal Analysis of a Historic Tower in Bari M. Diaferio, D. Foti, M. Mongelli, University of Bari; N.I. Giannoccaro, University of Salento; P. Andersen, Structural Vibration Solutions A/S
335
32
Monitoring of a Movable Bridge Mechanical Components for Damage Identification Using Artificial Neural Networks T. Dumlupinar, F.N. Catbas, University of Central Florida
343
33
Ambient Vibration Based Modal Identification of a Flexible Retaining Wall C. Rainieri, A. Dey, C. Laorenza, G. Fabbrocino, F. Santucci de Magistris, University of Molise
349
34
Structural Health Monitoring of a Bridge Model Using ARX Models T. Terrell, M. Gul, F.N. Catbas, University of Central Florida
357
35
Structural Health Monitoring and Evaluating Structural Performance of a Stadium H.O. Sazak, F.N. Catbas, M. Gul, University of Central Florida
365
36
Estimation of Damping for one of the New European Court Towers in Luxembourg R. Brincker, Aarhus University; A. Brandt, University of Southern Denmark; C.T. Georgakis, J.H. Roldsgaard, Technical University of Denmark; D. Schilz, INCA Ingenieurs Conseils Associes
373
37
Use of Wireless Smart Sensors for Detecting Human Falls Through Structural Vibrations B.T. Davis, J.M. Caicedo, S. Langevin, V. Hirth, University of South Carolina
383
38
Structural Health Indices for Steel Truss Bridges R.K. Giles, R. Kim, B.F. Spencer, Jr., L.A. Bergman, University of Illinois at Urbana-Champaign; C.K. Shield, University of Minnesota; S.C. Sweeney, U.S. Army ERDC-CERL
391
Multi-Dimensional Calibration of Impact Dynamic Models Lucas G. Horta, Mercedes C. Reaves, Martin S. Annett, and Karen E. Jackson NASA Langley Research Center Hampton, VA 23681
ABSTRACT NASA Langley, under the Subsonic Rotary Wing Program, recently completed two helicopter tests in support of an in-house effort to study crashworthiness. As part of this effort, work is on-going to investigate model calibration approaches and calibration metrics for impact dynamics models. Model calibration of impact dynamics problems has traditionally assessed model adequacy by comparing time histories from analytical predictions to test at only a few critical locations. Although this approach provides for a direct measure of the model predictive capability, overall system behavior is only qualitatively assessed using full vehicle animations. In order to understand the spatial and temporal relationships of impact loads as they migrate throughout the structure, a more quantitative approach is needed. In this work impact shapes derived from simulated time history data are used to recommend sensor placement and to assess model adequacy using time based metrics and orthogonality multi-dimensional metrics. An approach for model calibration is presented that includes metric definitions, uncertainty bounds, parameter sensitivity, and numerical optimization to estimate parameters to reconcile test with analysis. The process is illustrated using simulated experiment data.
NOMENCLATURE gi(t) i,k,l m n p t v y E Et I L M1 M2 N
scalar function with time variation integer indices number of measured locations number of impact shapes used in reconstruction parameter vector Time response vector used for calibration response vector modulus of elasticity tangent modulus identity matrix total number of time samples calibration Metric 1 calibration Metric 2 number of solutions
T. Proulx (ed.), Civil Engineering Topics, Volume 4, Conference Proceedings of the Society for Experimental Mechanics Series 7, DOI 10.1007/978-1-4419-9316-8_1, © The Society for Experimental Mechanics, Inc. 2011
1
2
Q Qe Gi 'T Vi . ) < ,) Prob max
( )T
scalar with 2-norm of response vector for analysis scalar with 2-norm of response vector for experiment ith impact shape contribution to response sample time ith singular value ith impact shape vector at measured locations shape matrix at un-measured locations shape matrix at measured locations normalized impact shape matrices probability maximum value relational operators maximum of all maxima minimum of all minima matrix transpose 1
INTRODUCTION
As computational capabilities continues to improve and the cost associated with test programs continue to increase, certification of future rotorcraft will rely more on computational tools along with strategic testing of critical components. Today, MIL-STD 1290A (AV) [1] encourages designers of rotary wing vehicles to demonstrate compliance with the certification requirements for impact velocity and volume loss by analysis. Reliance on computational tools, however, will only come after rigorous demonstration of the predictive capabilities of existing computational tools. NASA, under the Subsonic Rotary Wing Program, is sponsoring the development and validation of such tools. Jackson in [2] discussed detailed requirements and challenges associated with certification by analysis. Fundamental to the certification effort is the demonstration of verification, validation, calibration, and algorithms for this class of problems. Work in this paper deals with the model calibration of systems undergoing impact loads. The process of model calibration, which follows the verification and validation phases, involves reconciling differences between test and analysis. Most calibration efforts combine both heuristics and quantitative methods to assess model deficiencies, to consider uncertainty, to evaluate parameter importance, and to compute required model changes. Calibration of rotorcraft structure models presents particular challenges because the computational time, often measured in hours, limits the number of solutions obtainable in a timely manner. Oftentimes, efforts are focused on predicting responses at critical locations as opposed to assessing the overall adequacy of the model. For example Kamat in [3] conducted a survey, which at the time, studied the most popular finite element analysis codes and validation efforts by comparing impact responses from a UH-1H helicopter drop test. Similarly, Wittlin and Gamon in [4] used the KRASH analysis program for data correlation of the UH-1H helicopter. Another excellent example of a rotary wing calibration effort is that of Cronkhite and Mazza in [5] comparing results from a U.S. Army’s Composite helicopter with simulation data from the KRASH analysis program. Recently, Tabiei, Lawrence, and Fasanella [6] reported on a validation effort using anthropomorphic test dummy data from crash tests to validate an LS-DYNA finite element model. Common to all these calibration efforts is the use of scalar deterministic metrics.
3
One complication with calibration efforts of nonlinear models is the lack of universally accepted metrics to judge model adequacy. Work by Oberkampf et al. [7] and later Schwer et al. [8] are two noteworthy efforts that provide users with metrics to evaluate nonlinear time histories. Unfortunately, seldom does one see them used to assess model adequacy. In addition, the metrics as stated in [7] and [8] do not consider the multi-dimensional aspect of the problem explicitly. A more suitable metric for multi-dimensional calibration exploits the concept of impact shapes as proposed by Anderson et al. in [9] and demonstrated by Horta et al. in [10]. Aside from the metrics themselves, the verification, validation, and calibration elements, as described by Roach [11], Oberkampf [12], Thaker [13], and Atamturktur [14], must be adapted to rotorcraft problems. Because most applications in this area use commercially available codes, it is assumed that code verification and validation have been addressed elsewhere. Thus, this work concentrates on calibration elements only. Fundamental to the success of the model calibration effort is a clear understanding of the ability of a particular model to predict the observed behavior in the presence of modeling uncertainty. The approach proposed in this paper is focused primarily on model calibration using parameter uncertainty propagation and quantification, as opposed to a search for a reconciling solution. The process set forth follows a three-step approach. First, Analysis of Variance (ANOVA) as described in references [15-18] is used for parameter selection and sensitivity. To reduce the computational burden associated with variance based sensitivity estimates, response surface models are created and used to estimate time histories. In our application, the Extended Radial Basis Functions (ERBF) response surface method, as described by Mullur [19-20], has been implemented and used. Second, after ANOVA estimates are completed, uncertainty propagation is conducted to evaluate uncertainty bounds and to gage the ability of the model to explain the observed behavior by comparing the statistics of the 2norm of the response vector between analysis and test. If the model is reconcilable according to the metric, the third step seeks to find a parameter set to reconcile test with analysis by minimizing the prediction error using the optimization scheme proposed Regis and Shoemaker [21]. To concentrate on the methodology development, simulated experiment data has been generated by perturbing an existing model. Data from the perturbed model is used as the target set for model calibration. To keep from biasing this study, changes to the perturbed model were not revealed until the study was completed. In this paper, a general description of the helicopter model will first be presented, followed by a description of basic model calibration elements. These elements include time and spatial multi-dimensional metrics, parameter selection, sensitivity using analysis of variance, and optimization strategy for model reconciliation. Other supporting topics discussed are sensor placement to assure proper evaluation of multi-dimensional orthogonality metrics, prediction of unmeasured responses from measured data, and the use of surrogates for computationally efficiency. Finally, results for the helicopter calibrated model are presented and, at the end, the actual perturbations made to the original model are revealed for a quick assessment. 2
DESCRIPTION OF HELICOPTER TEST ARTICLE
A full-scale crash test of an MD-500 helicopter, as described by Annett and Polanco in [22], was conducted at the Landing and Impact Research (LandIR) Facility at NASA Langley Research Center (LaRC). Figure 1a shows a photograph of the test article while it was being prepared for test, including an experimental dynamic energy absorbing honeycomb structure underneath the fuselage designed by Kellas [23]. The airframe, provided by the US Army's Mission Enhanced Little Bird
4
(MELB) program, has been used for civilian and military applications for more than 40 years. NASA Langley is spearheading efforts to develop analytical models capable of predicting the impact response of such systems. 3
LS-DYNA MODEL DESCRIPTION
To predict the behavior of the MD-500 helicopter during a crash test, an LS-DYNA [24] finite element model (FEM) of the fuselage, as shown in Figure 1b, was developed and reported in [22]. The element count for the fuselage was targeted to not exceed 500,000 elements, including seats and occupants; with 320,000 used to represent the energy absorbing honeycomb and skid gear. Shell elements were used to model the airframe skins, ribs and stiffeners. Similarly, the lifting and pullback fixtures, and the platform supporting the data acquisition system (mounted in the tail) were modeled using rigid shells. Concentrated masses like the rotor, tail, and fuel were modeled using ballast. For materials, the fuselage section is modeled using Aluminum 2024-T3 with elastic-plastic properties, whereas the nose is fiberglass and the engine fairing is Kevlar fabric. Instead of using the complete “as-tested” FEM model, this study uses a simplified model created by removing the energy absorbing honeycomb, skid gears, anthropomorphic dummies, data acquisition system, and lifting/pull-back fixtures. After these changes, the resulting simplified model is shown in Figure 2. Even with all these components removed, the simplified model had 27,000 elements comprised primarily of shell elements to represent airframe skins, ribs and stiffeners. The analytical test case used for calibration, simulates a helicopter crash onto a hard surface with vertical and horizontal speeds of 26 ft/sec and 40 ft/sec, respectively. For illustration, Figure 3 shows four frames from an LS-DYNA simulation as the helicopter impacts the hard surface.Equation Chapter 1 Section 4 Equation Chapter (Next) Section 1 4
PROBLEM FORMULATION
Calibration of models is a process that requires analysts to integrate different methodologies in order to achieve the desired end goal which is to reconcile prediction with observations. Starting with an analytical model that incorporates the physics of the test article, this model is initially judged based on some pre-established calibration metrics. Although there are no universally accepted metrics, the work in this paper uses two metrics; one that addresses the predictive capability of time responses and a second metric that addresses multi-dimensional spatial correlation of sensors for both test and analysis data. After calibration metrics are established, the next step in the calibration process involves parameter selection and uncertainty estimates using engineering judgment and available data. With parameters selected and uncertainty models prescribed, the effect of parameter variations on the response of interest must be established.
If parameter variations are found to
significantly affect the response of interest, then calibration of the model can proceed to determine a parameter set to reconcile the model. These steps are described in more detail, as follows. 4.1
Time Domain Calibration Metrics
Calibration metrics provide a mathematical construct to assess fitness of a model in a quantitative manner. Work by Oberkampf [7] and Schwer [8] set forth scalar statistical metrics ideally suited for use with time history data. Metrics in terms of mean, variance, and confidence intervals facilitate assessment of experimental data, particularly when probability statements are sought. For our problem, instead of using response predictions at a particular point, a vector 2-norm of the system response is used as a function of time. An important benefit of using this metric is that it provides for a direct measure
5
of multi-dimensional closeness of two models. In addition, when tracked as a function of time, closeness is quantified at each time step.
Because parameters are uncertain, statistical measures of the metric need to be used to conduct assessments. With limited information about parameter uncertainty a uniform distribution function, which is the least informative distribution function, is the most appropriate representation to model parameter uncertainty. This uncertainty model is used to create a family of N equally probable parameter vectors, where N is arbitrarily selected. From the perspective of a user, it is important to know the probability of being able to reconcile measured data with predictions, given a particular model for the structure and parameter uncertainty. To this end, let Q(t , p)
v(t , p)
be a scalar time varying function, in which the vector v is defined
2
as the 2-norm of the response at time t, using parameter vector p. Furthermore, let V (t ) over all parameter variations, and let V (t )
min Q(t , p) be the minimum value p
max Q(t , p) be the maximum value. Using these definitions and N LS-DYNA p
solutions corresponding to equally probable parameter vectors, a calibration metric can be established to bound the probability of test values falling outside the analysis bounds as; M 1=Prob(V (t ) Qe (t ) Qe (t ) ! V (t )) 1 / N
(4.1)
where Qe (t ) is the 2-norm of responses from the experiment. Note that N controls tightness of the bounds and also the number of LS-DYNA solutions required. The use of norms, although convenient, tends to hide the spatial relationships that exist between responses at different locations in the model. In order to study this spatial multi-dimensional dependency explicitly, a different metric must be established. 4.2
Spatial Multi-Dimensional Calibration Metric
Spatial multi-dimensional dependency of models has been studied in classical linear dynamic problems in terms of mode shapes or eigenvectors resulting from a solution to an eigenvalue problem. Unfortunately, the nonlinear nature of impact problems precludes use of any simple eigenvalue solution scheme. Alternatively, an efficient and compact way to study the spatial relationship is by using a set of orthogonal impact shapes basis vectors. Impact shapes, proposed by Anderson [9] and later by Horta [10], are computed by decomposing the time histories using orthogonal decomposition. For example, time histories from analysis or experiments can be decomposed using singular value decomposition as n
y( x, t )
¦V I ( x)g (t ) i i
(4.2)
i
i 1
In this form, the impact shape vector
Ii
sized m x 1 contains the spatial distribution information for m sensors, g(t) contains
the time modulation information, V contains scalar values with shape participation factors, and n is the number of impact shapes to be included in the decomposition, often truncated based on allowable reconstruction error. Although Eq. (4.2) is written in continuous time form, for most applications, time is sampled at fixed intervals such that t
k 'T where the integer
k=0,…,L and 'T is the sample time. From Eq. (4.2), the fractional contribution of the i impact shape to the total response th
is proportional to G i , defined as;
6
Gi
Vi
n
¦V
(4.3)
l
l 1
Mimicking the approach used in classical dynamic problems, impact shapes can now be used to compare models using the orthogonality. Orthogonality, computed as the dot product operation of vectors (or matrices), quantifies the projection of one vector onto another. If the projection is zero, vectors are orthogonal, i.e., uncorrelated. This same idea applies when comparing test and analysis impact shapes. Numerically, the orthogonality metric is computed as; M2
)T
u( y G ) u( y G lt )@ .
(2)
Note that within the present studyUt is assumed to be constant; however, piecewise-constant or other distributionsUt([ ) may be considered and adopted in the following developments. Equation 2 contains the unit step distribution u(), Lt and lt are the length and the width of the strip modelling the train, and G is the distance between the side of the strip and the edge y = 0 of the plate, as shown in Fig. 1. The second term on the right-hand side of Eq. 1 describes the inertial action of the load. The total acceleration may be expressed in the following general form:
d 2w dt 2
w2 w w2 w w2w w2w w2w w2w ww ww v 2x 2 2v x v y v 2y 2 2v x 2v y ax ay 2 wxwy wxwt wywt wx wy wt wx wy
(3)
where vx, vy, ax, ay express the velocities and accelerations of the travelling load in the x and y directions respectively [1]. Considering a train travelling at constant speed v in the x direction, Eq. 3 reduces to: d 2w dt 2
2 w2 w w2 w 2 w w 2v v wxwt wt 2 wx 2
(4)
The first term of the right-hand side of Eq. 4 expresses the influence of vertical acceleration of the moving load, the second term the influence of Coriolis acceleration, and the third term the influence of track curvature [1].
y Lb Lt
vx
lt
G
0
lb
x Fig. 1 Model scheme
23
The functional of the total potential energy of the coupled system can be written as the sum of a term U due to the strain energy plus a term V representing the potential of all applied loads (including the inertial forces): 3 U V
(5)
The potential of the strain energy can be written in terms of second order derivatives of the out-of-plane displacement w: l L
U
1b b D ª wxx2 wyy2 2Q wxx wyy 2(1 Q ) wxy2 º¼ dxdy , D 2 ³0 ³0 ¬
Eh3 12(1 Q 2 )
(6)
where the subscripts denote differentiation with respect to the spatial variables and D is the flexural stiffness of the plate, expressed as a function of Young’s modulus E, Poisson’s ratio Q and thickness h [7]. In the adopted formulation the inertial forces are included in the potential of applied loads V as follows: lb Lb
³ ³ ( U ww wp) dxdy
V
(7)
b
0 0
where Ub is the mass per unit area of the plate and p is the load in Eq. 1. The out-of-plane displacement w is expressed by means of a linear combination of shape functions, selected as products of homogeneous uniform prismatic beam eigenfunctions I: w
N
¦q
n
In ([ ,K ) w I7 q
(8)
n 1
where q is the generalized coordinate vector. Introducing the displacement expansion in the quadratic functional 3, and imposing its stationarity, yields the following algebraic eigenproblem:
2r Įȕ['C] q ȕ 2 [K r Į 2 'K ] q [M r 'M ] q
rg f
(9)
with: r
Ut , Į Ub
v , ȕ Lbȕ
1 L2b
D
(10)
Ub
where E is a frequency parameter and D is a dimensionless parameter depending on the speed v. The matrices in square brackets in Eq. 9 can be regarded as dimensionless quantities, and they can be computed according to the following integrals: lb Lb
M
7 ³ ³ (I I ) dxdy , 'M
į lt x1
0 0
³
7 ³ (I I ) dxdy , 'C
į x0
į lt x1
Lb
³ ³ (I I
T x
) dxdy
į x0
lb Lb
K
L4b ³ ³ [Ixx IxxT Iyy IyyT Q (Ixx IyyT Iyy IxxT ) 2(1 Q )(Ixy IxyT )] dxdy
(11)
0 0
'K
2 L2b
į lt x1
į lt x1
į x0
į x0
³
T ³ (I Ixx ) dxdy , f
³ ³ I dxdy
In Eq. 11 the integration interval [x0, x1] is time-dependent. Introducing the ratio between the lengths Lt and Lb:
ȝ then x0 and x1 vary according to Tab. 1.
Lt Lb
(12)
24
Table 1 Time-dependent interval of integration
case ȝ 1
case ȝ ! 1
case ȝ 1
0 d vt Lt
x1 ® ¯ x0
vt
Lt d vt Lb
x1 ® ¯ x0
vt
x1 Lb d vt Lb Lt ® ¯ x0
Lb
0
vt Lt
x1 ® ¯ x0
vt
x1 Lb d vt 2 Lb ® ¯ x0
Lb
0 d vt Lb
0
vt Lb
vt Lt
0 d vt Lb
x1 ® ¯ x0
vt
Lb d vt Lt
x1 ® ¯ x0
Lb
x1 Lt d vt Lt Lb ® ¯ x0
Lb
0
0
vt Lt
To model energy dissipation within the structure, a dimensionless damping matrix C may be defined by means of the plate modal matrix ) (mass normalized) and eigenvalues Zn (computed from the M and K matrices), and considering a modal damping ratio ] equal for all modes: §Z · 2ȗC 2ȗĭ T diag ¨ n ¸ ĭ1 © ȕ ¹
§ Z · ĭ T diag ¨ 2ȗ n ¸ ĭ1 ȕ ¹ ©
(13)
Introducing Eq. 13 in Eq. 9 yields:
2ȕ[ȗC r Į 'C] q ȕ 2 [K r Į 2 'K ] q [ M r 'M ] q
rg f
(14)
Equation 14 is a reduced order discretized model with time-dependent coefficients, which can be solved numerically. 3 Numerical results Some numerical examples are presented for studying the dynamic behaviour of the model described in section 2. The influence of parameters v, r, P, E, ] governing Eq. 14 is highlighted by studying time responses w(t) and dynamic response functions H of the dimensionless frequency D (playing the role of ‘frequency response functions’) defined according to:
H ( x , y; D )
max t > w(t ) @ ws
(15) x , y ;D
where ws is the static deflection due to the load centered in Lb/2. Numerical solutions of Eq. 14 are computed using the Runge-Kutta algorithm, expanding the solution w (Eq. 8) with 4 u 2 beam eigenfunctions (4 pinned – pinned eigenfunctions along x direction and 2 free – free eigenfunctions along y direction). Realistic values for parameter E are computed by means of the empirical expression: ȕ
2S a
O12
LbJ
(16)
based on large collections of experimental data [2], where a and J are parameters depending on the kind of bridge considered, as reported in Tab. 2. The values (in Hz) of the first natural frequency and of parameter E for different kinds of bridges are reported as functions of the length Lb in Fig.2.
25
25
18 General bridges Steel truss bridges Steel plate girder bridges with ballast Concrete bridges with ballast Concrete bridges without ballast
14 12
15
E [rad/s]
f1 [Hz]
20
General bridges Steel truss bridges Concrete bridges with ballast
16
10
10 8 6 4
5
2 0 10
20
30
40
50
60
70
80
90
100
0 10
20
30
40
50
60
70
80
90
100
Lb [m]
Lb [m]
Fig. 2 First natural frequency f1 [Hz] (left) and frequency parameter E [rad/s] (right) as functions of the length Lb for different kinds of bridges Table 2 Parameters in Eq.16, as reported in [2] a
J
General bridges (average case)
133
0.9
Steel truss bridges
307
1.1
Steel plate girder bridges with ballast
59
0.7
Steel plate girder bridges without ballast
208
1
Concrete bridges with ballast
190
1.1
Concrete bridges without ballast
225
1.2
Kind of bridge
Effect of the speed of the load As a reference case study, the following values for the parameters are assumed: Plate: Lb = 50 m, lb = 10 m, E = 5 rad/s, ] = 0.05. Moving load: P = 1.4, lt = 2.5 m, G = 1.5 m, r = 0.5. Time responses w(t) are computed at coordinate x = Lb/2, y = lb/2 with speed v varying from 30 m/s to 50 m/s (108 Km/h to 180 Km/h), as shown in figure 4. Maximum deflection at different points Parameter values are assumed as in the reference case. Response functions H(D) are computed at different points (x, y) along the structure, as reported in Fig. 3. The frequency parameter varies from 0.1 up to 1, i.e. v varies from 40 Km/h up to 400 Km/h. Functions H(x, y; D) show a peculiar undulating trend, not significantly affected by the choice of coordinate x. Effect of the mass of the load Parameter values are assumed as in the reference case with v = 40 m/s = 144 Km/h, varying r from 0.1 to 1. Response functions w(t) and H(D) are computed in x = Lb/2, y = lb/2, as reported in Fig. 4. Effect of the length of the load Parameter values are assumed as in the reference case with v = 40 m/s = 144 Km/h, varying P from 0.1 to 2. Response functions w(t) and H(D) are computed in x = Lb/2, y = lb/2, as reported in Fig. 5. Parameter P (related to Lt) is able to significantly affect the behaviour of H(D). Note that the plots of H(D) in the case P t 1 are superimposed.
26
0.5
x 10
-3
1.1 x = Lb/2; y = lb/2
0
1.06
-1.5
x = Lb/2; y = lb
H
w [m]
x = 3Lb/4; y = lb/2 x = Lb/2; y = 0
-1
1.04
-2 -2.5
1.02
v = 40 m/s v = 50 m/s v = 30 m/s
-3 -3.5
x = Lb/4; y = lb/2
1.08
-0.5
0
1
2
3
4
1 0.1
5
0.2
0.3
0.4
0.5
t [s]
D
0.6
0.7
0.8
0.9
1
0.7
0.8
0.9
1
Fig. 3 Effect of parameter v on w(t) (left); H(D) at different points (x, y) (right) 1
x 10
-3
1.15 r = 1.0 r = 0.8 r = 0.5 r = 0.1
0 -1
1.1
-3
H
w [m]
-2
-4
1.05
-5
r = 0.5 r = 1.0 r = 0.1
-6 -7
0
1
2
3
4
1 0.1
5
t [s]
0.2
0.3
0.4
0.5
D
0.6
Fig. 4 Effect of parameter r on w(t) (left); effect of parameter r on H(D) (right) 0.5
x 10
-3
1.45
1.35
-0.5
1.3
-1
1.25
-1.5
H
w [m]
P = 2.0 P = 1.0 P = 0.5 P = 0.1
1.4
0
1.2
-2
P = 0.2 P = 0.6 P = 1.0 P = 1.4 P = 1.8
-2.5 -3 -3.5
0
1
2
3
t [m]
4
1.15 1.1 1.05
5
1 0.1
0.2
0.3
0.4
0.5
D
0.6
0.7
Fig. 5 Effect of parameter P on w(t) (left); effect of parameter P on H(D) (right)
0.8
0.9
1
27
1.1
1.1
1.08
E=5
0.5
rad/s
E=3
0.5
rad/s
E=8
0.5
rad/s
]=0 ] = 0.01
1.08
] = 0.05 ] = 0.10
1.06
] = 0.20 ] = 1.00
H
H
1.06
1.04
1.04
1.02
1.02
1 0.1
0.2
0.3
0.4
0.5
D
0.6
0.7
0.8
0.9
1 0.1
1
0.2
0.3
0.4
0.5
D
0.6
0.7
0.8
0.9
1
Fig. 6 Effect of parameter E on H(D) (left); effect of parameter ] on H(D) (right) 0.5
x 10
-3
1.35
1.25
-0.5
1.2
-1
H
w [m]
F = 1/2 F = 1/6 F = 1/48 F = 1/480
1.3
0
1.15
-1.5
F = 1/480
1.1
F = 1/48
-2
F = 1/6
1.05
F = 1/2
-2.5
0
0.5
1
1.5
2
2.5
3
3.5
1 0.1
0.2
0.3
t [s]
0.4
0.5
D
0.6
0.7
0.8
0.9
1
Fig. 7 Effect of partially distributed load on w(t) (left); effect of partially distributed load on on H(D) (right) Effect of structural stiffness Parameter values are assumed as in the reference case, varying E from 3 rad/s to 8 rad/s. Response functions H(D) are computed in x = Lb/2, y = lb/2, as reported in Fig. 6. The plots are almost superimposed. Effect of structure damping Parameter values are assumed as in the reference case, varying ] from 0 to 1. Response function H(D) is computed in x = Lb/2, y = lb/2, as reported in Fig. 6. Raising ] reduces the amplitude of oscillation of H(D), until its behaviour becomes monotonic (however this is not the case for real bridge structures). Effect of partially distributed load Parameter values are assumed as in the reference case, with Lt = 24 m and v = 40 m/s = 144 Km/h. Different loading distributions are compared: the continuous one (as represented in Fig. 1) and ‘partial’ distributions consisting of two shorter sections in which the load is distributed. The assumed ‘partial’ distributions are given by: 0d[ d
Lt
F
L · § and ¨ Lt t ¸ d [ d Lt F¹ ©
(17)
28
with F 0.5 (F = 0.5 yields the ‘continuous’ distribution). Since for the continuously distributed load it is assumed r0 = 0.5, for the ‘partially distributed’ load described by Eq. 17 r0 increases to r = 1/(2F) u r0. Response functions w(t) and H(D) are computed in x = Lb/2, y = lb/2 for different values of F (1/6, 1/48, 1/480) as reported in Fig. 7. Load distribution variations such as that described in Eq. 17 may dramatically affect the behaviour of the response function H(D). Effect of time dependent matrices Parameter values are assumed as in the reference case, with v = 40 m/s = 144 Km/h. The effect of neglecting the time dependent matrices 'M, 'C, 'K on the solution w(t) is evaluated by introducing a relative error, according to:
HM
w(t ) > w(t ) @'M
0
(18)
max t w(t )
where [w(t)]'M = 0 refers to the solution computed assuming 'M = 0 in Eq. 14. Similarly, HC and HK can be defined, considering 'C = 0 and 'K = 0. 6
x 10
-3
1.1 ' M = 'C = 'K = 0
4 1.08
2
1.06
-2
H
H
0
1.04
-4
HC
-6
HK
-8
HM
-10
Exact 'C = 'K = 0 'C = 0
1.02
H Tot
0
0.5
1
1.5
2
t [s]
2.5
3
3.5
4
1 0.1
0.2
0.3
0.4
0.5
D
0.6
0.7
0.8
0.9
1
Fig. 8 Effect of neglecting time dependent matrices 'M, 'C, 'K on w(t) (left) and on H(D) (right) The error functions HM, HC and HK are plotted versus time in Fig. 8, where HTot represents the total error, assuming 'M, 'C and 'K equal to 0 at the same time. The smallest, and negligible contribution to the error HTot appears to be HC, while the main contribution is due to HM. Fig. 8 also shows the effect on H(D) of neglecting the time dependent matrices 'M, 'C, 'K. Again, the effect of neglecting 'C is very small. The total error, on the contrary, can be significant.
4 Discussion Function H(x, y; D) appears to be an effective tool for studying the dynamic behaviour of a structure crossed by travelling loads with constant speed, in some way equivalent to a ‘frequency response function’ for time-varying coefficient systems. This function shows peculiar undulating trends (Fig. 3), influenced by the parameters governing Eq. 14. The response can be evaluated at any coordinate point (x, y) of the plate, making it possible to study the variation of structural deflection also along the y coordinate (Fig. 3). Mass parameter r can produce important shifts in magnitude, but not in shape (Fig. 4). On the contrary, length parameter P controls both shape and magnitude of H(D) (Fig. 5), but only in the case 0 < P d 1. The damping parameter ] has the effect of progressively smoothing the oscillation of H(D), until it becomes monotonic (Fig. 6, though the latter limit case is not realistic for actual bridge structures): in general, the reduction in amplitude becomes particularly significant at high speed. Frequency parameter E, within the range of real bridges, scarcely affects the behaviour of H(D) (Fig. 6), so H may be considered independent from E. Changes in the spatial distribution of the load can produce dramatic variations in H(D) (Fig. 7): this result should highlight the importance of properly modelling the ballast, directly influencing the load distribution on the actual structure. The contribution to the solution of the time dependent matrices 'M, 'C and 'K is globally not negligible (Fig. 8), however the effect of 'C is usually very small in comparison with the contributions of 'K, and especially of 'M.
29
5 Conclusions and future work In this study, the dynamical behaviour of railway bridges crossed by travelling trains was investigated by adopting a simplified model, i.e. a plate loaded by a travelling distributed mass, solved by means of the Rayleigh-Ritz method. The effects of each of the model governing parameters was studied introducing a dynamic function of the travelling speed, equivalent to a ‘frequency response function’ for time-varying coefficient systems. This function can be an effective tool for studying the dynamic behaviour of a structure crossed by travelling loads, since the travelling speed is the most important parameter influencing the dynamic stresses in railway bridges, which in general increase with increasing speed. In particular, it was shown how different spatial distributions of the load can deeply influence the dynamic response of the structure, highlighting the importance of properly modelling the ballast. Future work will thus concern this significant problem.
Acknowledgments This study was developed within the INTERMECH laboratory with the contribution of the Regione Emilia Romagna Assessorato Attività Produttive, Sviluppo Economico, Piano telematico, PRRIITT misura 3.4 azione A Obiettivo 2.
References [1]
Fryba L., Vibration of Solids and Structures under Moving Loads, 3rd edition, Telford, 1999.
[2]
Fryba L., Dynamics of Railway bridges, Telford, 1996.
[3]
Lin Y.H., “Vibration analysis of Timoshenko beams traversed by moving loads”, Journal of Marine Science and Technology 2 (4), pp. 25-35, 1994.
[4]
Stancioiu D., Ouyang H., Mottershead J.E., “Vibration of a continuous beam excited by a moving mass and experimental validation”, Journal of Physics, Conference series 181, 1999.
[5]
Adetunde I.A., “Dynamical Behavior of Euler-Bernoulli Beam Traversed by Uniform Partially Distributed Moving Masses”, Reasearch Journal of Applied Sciences 2 (4), pp. 476-483, 2007.
[6]
Lin Y.H., Cho C.H., “Vibration suppression of beam structures traversed by multiple moving loads using a damped absorber”, Journal of Marine Science and Technology 1 (1), pp. 39-48, 1993.
[7]
Timoshenko S., Young D.H., Weaver W., Vibration problems in engineering, 4th edition, Wiley, 1974.
Advanced Operational Modal Analysis Methods for Linear Time Periodic System Identification Mathew S. Allen, Department of Engineering Physics University of Wisconsin Madison, WI 53706 [email protected]
Shashank Chauhan Brüel & Kjær Sound & Vibration Measurement A/S Skodsborgvej 307 DK-2850 Nærum, Denmark [email protected] &
Morten Hartvig Hansen Wind Energy Department Risø National Laboratory P.O. Box 49, DK-4000 Roskilde Denmark
ABSTRACT Often structural dynamic systems cannot be modeled with constant stiffness, mass and damping. For example, wind turbines, helicopters, turbomachinery, and a variety of nonlinear structures linearized about a periodic limit cycle all may contain time-periodic terms in their equations of motion even if the equations are still linear. Linear time periodic systems such as these may exhibit parametric resonance, where the damping in the system is negative at certain rotational frequencies, leading to catastrophic failure. The authors previously presented an extension of operational modal analysis to linear time periodic systems. The previous work introduced a new type of spectrum, dubbed the harmonic autospectrum, discussed how to interpret the spectra, and showed how the simple peak picking method could be used to extract an estimate for the linear time-periodic model of a system from measurements. This paper builds on that work, revealing how more advanced operational modal analysis methods can be extended to linear time-periodic systems. Curve fitting approaches for both the harmonic autospectra and the positive harmonic spectra are applied to simulated measurements from two time-periodic systems, and the OMA based Enhanced Mode Indicator Function (EMIF) method is used to extract the modal parameters from the enhanced positive power spectrum. These extensions are found to provide more accurate estimates of the damping of the modes of the time-periodic systems, and to provide good estimates of the mode shapes of the systems so long as the measurements stand out clearly above the noise. Application of the complex mode indicator function an the EMIF algorithm makes it possible to separate the forward and backward whirling modes of a wind turbine, which is difficult since each of these modes is manifest at several harmonics due to the anisotropy in the tower supporting the turbine.
T. Proulx (ed.), Civil Engineering Topics, Volume 4, Conference Proceedings of the Society for Experimental Mechanics Series 7, DOI 10.1007/978-1-4419-9316-8_3, © The Society for Experimental Mechanics, Inc. 2011
31
32
1. INTRODUCTION
Many important structural systems require time-varying terms in their equations of motion. For example, the stiffness matrix of a two-bladed wind turbine [1] or any wind turbine with non-identical blades [2] depends on the angle of the rotor, and if the rotor speed is constant the equations of motion become linear time periodic (LTP). Linear time periodic equations of motion are commonly used to model other kinds of rotating machinery [3], for studying the stability of various classes of nonlinear systems linearized about a periodic limit cycle , and they have also been used to accelerate laser Doppler vibrometry measurements using the Continuous Scan (CSLDV) approach [4]. In a prior work, the authors presented a operational modal analysis (OMA) methodology for linear time periodic systems and validated it using simulated measurements from a Mathieu oscillator and a rotating wind turbine [5, 6]. Measurements from the latter were simulated using HAWC2 [7], a high fidelity simulation code that includes the random loading that the wind applies to the rotating turbine. The LTP-OMA methodology was subsequently applied to continuous scan laser vibrometry measurements from a few different structures [8], revealing that the methodology is capable of extracting detailed mode shapes from real experimental measurements. The authors’ original work [5, 6] presented the LTP-OMA methodology and focused on interpreting spectral measurements from the time-periodic systems. The simple peak-picking method was used to extract time-periodic models of the systems from the autospectrum of the output. Several more advanced techniques exist for interrogating and curve fitting operational modal analysis measurements from linear time invariant systems. This work shows how some of those methods can be applied to measurements from linear time periodic systems. Methods for curve fitting the spectra to a modal model, both the full harmonic power spectra as well as the positive harmonic spectra, and the two methods are compared. The OMA based Enhanced Mode Indicator Function (EMIF) method [9], an extension to the complex mode indicator function (CMIF) is also applied to the measurements and its performance is evaluated. The following section reviews the LTP-OMA identification methodology and shows how the advanced OMA methods can be adapted to measurements from the time periodic system. In Section 3 the methods are applied to simulated measurements from two linear time periodic systems, a Mathieu oscillator and a 5MW wind turbine in operation. Section 4 presents the conclusions. 2. THEORY
The state space equations of motion of a linear time-periodic (LTP) system can be written as follows, x = A(t ) x + B(t )u (1) y = C (t ) x + D (t )u
where A(t+TA) = A(t) and the other matrices are periodic as well with the same period. The fundamental frequency of the time-periodic system is denoted ωA = 2π/TA. The state transition matrix (STM) [10] gives the free response of such a system at time t via the relationship, x(t ) = Φ (t , t0 ) x(t0 ) . (2) One can also write the forced response of the system in terms of the STM as follows. t
y (t ) = C (t )Φ (t , t0 ) x(t0 ) + ∫ C (t )Φ (t ,τ ) B (τ ) u(τ )dτ t0
(3)
33
In the absence of degenerate roots, the state transition matrix of an LTP system can be represented as a modal sum [11, 12], N
Φ (t , t0 ) = ∑ψ r (t ) Lr (t0 )T eλr ( t −t0 )
(4)
r =1
where λr is the rth Floquet exponent of the state transition matrix, ψr is the rth time-periodic −T mode vector of the STM and Lr is the rth column of Ψ (t ) = [ψ 1 (t ) ψ 2 (t ) ] . The Floquet exponents of an LTP system are analogous to the eigenvalues of a Linear Time Invariant (LTI) system, which can be written in terms of the damping ratio ζr and natural frequency ωr as
λr = −ζ rωr + iωr 1 − ζ r 2 for an underdamped mode. In a prior work [5, 6], the authors used this equation to derive an expression for the output auto spectrum of a linear time periodic system in terms of the modal parameters of the state transition matrix. In order to achieve this, first an exponentially modulated version of the output signal is developed as follows, Y(ω ) = ⎡⎣
y−1 (ω )T
y0 (ω )T
y1 (ω )T
⎤⎦
T
(5)
where yn (ω ) =
∞
∫ y(t )e
( − iω −inω A ) t
dt
(6)
−∞
is simply the Fourier transform of y(t), shifted in frequency by nωA where n is an integer. Then, the harmonic autospectrum S yy (ω ) = E ( Y(ω )Y(ω ) H ) is given by the following equation when
D(t) = 0. N
∞
N
⎡⎣ S yy (ω ) ⎤⎦ = ∑ ∑ ∑
∞
∑
r =1 l =−∞ s =1 k =−∞
Cr ,l W(ω ) r , s ,l ,k Cs ,k H
[iω − (λr − ilω A )][iω − (λs − ikω A )]
H
(7)
The matrix W(ω) is a function of the input spectrum and is assumed to be reasonably flat. The dominant terms in the summation above are those for which iω − (λr − ilω A ) and iω − (λs − ikω A ) are both minimum at the same frequency. If the sidebands for mode r do not overlap with those for mode s, then the largest terms occur when r=s and l=k and the expression becomes the following. N
∞
⎡⎣ S yy (ω ) ⎤⎦ ≈ ∑ ∑
r =1 l =−∞
Cr ,l W(ω ) r ,l Cr ,l H
[iω − (λr − ilω A )][iω − (λr − ilω A )]
This is a summation of terms with ( iω − λ )( iω − λ )
H
H
(8)
in the denominator, or squared modal
contributions, so the harmonic autospectrum has the same mathematical form as the output autospectrum of a multi-output linear time invariant system. However, there are a few important differences. First, the expression contains a summation over both the modes, whose eigenvalues are λr, and also a summation over the harmonics of ωA using the integer index l. As a result, the autospectrum of the LTP system will have peaks near each natural frequency, ωr, and also at the frequencies ωr ± lωA for any integer l. Second, the mode vectors Cr ,l are actually the Fourier coefficients of the time-varying mode shapes of the LTP system as discussed in [5, 6]. These vectors could have infinite dimension, but in practice some finite number of terms Np is
34
sufficient to accurately describe the mode shape. Hence, if the system as No outputs, then the vectors Cr ,l have length No×Np, where Np is the number of modulations used to form Y(ω) in eq. (5).
Many different vectors Cr ,l can be identified, but each contains the same Fourier
coefficients of the observed mode vectors shifted by l. Specifically, if the time varying mode shape at the output is C (t )ψ r (t ) =
∞
∑C
n =−∞
r ,n
e jnω At
(9)
then Cr ,l is given by. Cr ,l = ⎡⎣
Cr ,−1−l T
Cr , − l T
Cr ,1−l T
⎤⎦
T
(10)
Additional details of this theory can be found in [5, 6]. In a previous work [5], the authors used the peak picking method together with the expressions developed above to estimate the natural frequencies and the mode vectors Cr ,l of the LTP system. This work explores other more advanced methods, including curve fitting and advance spatial domain tools, such as CMIF, based methods for the purpose of modal parameter estimation of LTP systems. 2.1. Curve fitting autospectra
Since the harmonic autospectrum, (or equivalently harmonic power spectral density, here denoted HPSD) of an LTP system was shown in eq. (8) to have same mathematical form as an LTI system (i.e. it is a summation of modal contributions squared), it can be put into a convenient form for curve fitting. Assuming underdamped modes λr+(N/2) = λr*, where ()* denotes complex conjugate, and contribution due to each underdamped mode can be collected and written as follows after a partial fraction expansion. N /2 ∞ A r ,l A * r ,l B r ,l B* r ,l (11) ⎡⎣ S yy (ω ) ⎤⎦ ≈ ∑ ∑ + + + ⎡⎣iω − (λr − ilω A )* ⎤⎦ [ −iω − (λr − ilω A )] ⎡⎣ −iω − (λr − ilω A )* ⎤⎦ r =1 l =−∞ [ iω − (λr − ilω A ) ]
Alternatively, this could be written as a matrix polynomial model with complex coefficients in even powers of (iω). Various algorithms can fit a model of this form to a set of measurements. For this work the Algorithm of Mode Isolation (AMI) [13] was modified to fit a model of the form described above and used to extract modes from the harmonic autospectrum. 2.2. Positive Power Spectra
The power spectrum of a signal possesses four quadrant symmetry, and hence it includes information of system poles twice. This can also be seen in eq. (11), which contains each of the poles, once having positive damping and other having negative damping (note that same is true for the complex conjugate pole as well). These extra negatively damped poles can lead to difficulty when estimating the modal parameters of the system. This problem can be dealt with by considering positive power spectra instead of power spectra. The inverse FFT of the power spectrum is found, and then a rectangular window is used in the time domain to eliminate the negatively damped modes by retaining only the decaying part of the impulse response function. Thus PPS functions have the same form as a typical frequency response function (FRF) has, making them more suitable for modal parameter estimation. The process of estimating PPS functions from power spectra is described in detail in [9]. In this work, the same approach is used to convert a set of harmonic power spectra into positive harmonic power spectra. Once this has been done, the negatively damped poles are no
35
longer present and the positive harmonic power spectrum (pHPSD or S + yy (ω ) ) can be written in the following mathematical form. N /2 ∞ A r ,l A * r ,l ⎡⎣ S + yy (ω ) ⎤⎦ ≈ ∑ ∑ + ⎡⎣iω − (λr − ilω A )* ⎤⎦ r =1 l =−∞ [ iω − (λr − ilω A ) ]
(12)
The pHSPD has the usual magnitude and phase relationship as a frequency response function, so it can often be more accurately curve fit. In the following the AMI algorithm in [13, 14] was used to fit a model of the form above to the power spectra in order to construct the time periodic modes of the system. Since the pHSPD has the same mathematical form as a frequency response function, no modification to the AMI algorithm was necessary when doing this. 2.3. EMIF based Method
The enhanced mode indicator function (EMIF) method was proposed for operational modal analysis in [9]. Here that approach is adapted to parameter estimation on harmonic power spectral measurements (HPSD). OMA-EMIF method works on positive power spectra, which are defined in section 2.2. One begins by choosing a frequency range of interest and also the number of modes to be identified in the chosen frequency range. The number of modes to be identified is typically equal to the number of dominant peaks in the CMIF plot over the chosen frequency range. The positive power spectra in the chosen frequency range are then assembled in an augmented matrix form as shown below, where Ntot is the number of responses and Nf is number of frequency lines in chosen frequency range. For an LTP system there are Np responses for each measurement point, so Ntot=NoNp.
[ A0 ]N
tot
× ( N tot N f )
(
)
+ … S 21 (ωi ) … S21+ ωi + N f
(
)
…
⎡ S11+ (ωi ) … S11+ ωi + N f ⎢ ⎢ = ⎢ + ⎢ S pq (ωi ) … S pq+ ωi + N f ⎣
(
…
…
…
)
(
)
(
)
… S p+1 (ωi ) … S p+1 ωi + N f ⎤ ⎥ ⎥ ⎥ + … … … S pq ωi + N f ⎥ ⎦
(13)
This matrix now has all of the spatial information contained in the autospectrum, over all of the frequency lines that have been included in [A0]. A singular value decomposition of this augmented matrix yields left and right singular vectors along with the singular values. Supposing that Nb modes are to be indentified from the measurements, the first Nb dominant left singular vectors are chosen to create an enhanced positive power spectra in the chosen frequency range of interest. This is shown in below, where [U] is a matrix of left singular vectors of the augmented matrix.
ePPS (ω ) = ⎡⎣ S (ω ) ⎤⎦ N × N = [U ]TNb × Ntot ⎡⎣ S + (ω ) ⎤⎦ Ntot × Ntot b tot
(14)
The ePPS functions can now be used to identify the Nb modes by applying a single degree-offreedom unified matrix polynomial [15] algorithm to each of the peaks in the ePPS. The modal vectors obtained from ePPS functions are in a condensed set of coordinates and can be converted back into physical domain by pre-multiplying the obtained modal vectors with [U ]N × N . b
tot
3. Simulated Application Examples 3.1. Mathieu Oscillator
This section applies the proposed methodologies to two different systems. The first is the simple Mathieu oscillator, a spring mass-damper system with time-periodic spring stiffness. The equation of motion for the oscillator is
36
(
)
y + 2ζω 0 y + ω02 + ω12 cos(ωA t ) y =
u (t ) m
(15)
with 2ζω 0 = c / m , ω 02 = k 0 / m , ω12 = k1 / m , and input u(t). The parameters used in this work are the following, m=1, k0=1, k1=0.4, and ωA=0.8 rad/s, which are the same as those used in [5, 6]. 3.1.1.
Curve Fitting HPSD
The response of the Mathieu oscillator to a broadband random input was simulated using time integration, and the harmonic autospectrum (HPSD) was then computed with p=2, so n = 2…2 in eq. (6) so Np=5 and a 5x5 HPSD matrix was produced. The primary column of this matrix (i.e. the (p+1)th or 3rd column in this case) contains all of the information needed to identify the parameters of the system, so only it will be considered in the curve fit. The composite [13], or average, of this 5x1 matrix is shown in Figure 1 with a black line. The composite of a curve fit is also shown as well as the difference between the two. The curve fit can be seen to follow the measurements closely near each of the peaks in the spectrum, but the baseline level of the curve fit is too high so the highest frequency peak is almost obscured by the tails of the dominant modes. Com posite of Residual After Mode Isolation & Refinem ent Data Fit Data-Fit
0
10
-2
10
-4
10
0
0.1
0.2
0.3
0.4 0.5 Frequency (Hz)
0.6
0.7
0.8
Figure 1: Harmonic Autospectrum of Mathieu Oscillator response with: (black line) composite of actual autospectrum, (green dotted) curve fit to the HPSD and (red) composite of the difference between the curve fit and the measurements.
The authors have seen this same phenomena occur when fitting conventional autospectra of structural systems. For those systems the curve fitting routine seems to ascribe more complexity to the modes than is warranted. Better results can sometimes be obtained if the curve fitting routine is forced to fit a real mode model to the measurements, but that is not possible for measurements from an LTP system since the HPSD must be fit to a complex mode model. The
37
authors also noted that the damping obtained for each of the modes seemed to be somewhat sensitive to the number of points around each peak that was used in the curve fit. 3.1.2.
Curve Fit to Positive HPSD (Syy+)
The positive power spectra of the measurements were formed as discussed in Section 2.2. Figure 2 shows a composite of the 5x1 positive power spectrum in the same format used in the previous figure, as well as a curve fit. The positive power spectrum seems to be smoother than the full HSPD, and the fit agrees very well with the measurements. The residual (red line) is small and seems to contain only noise, suggesting that the curve fit has extracted all of the meaningful information from the measurement. The results of the curve fits shown in Figures 1 and 2 are summarized in Table 1, which will be discussed later. Com posite of Residual After Mode Isolation & Refinem ent Data Fit Data-Fit 0
10
-1
10
-2
10
0
0.1
0.2
0.3
0.4 0.5 Frequency (Hz)
0.6
0.7
0.8
Figure 2: Positive Harmonic Autospectrum (Syy+) of Mathieu Oscillator response with: (black line) composite of actual autospectrum, (green dotted) curve fit to the HPSD and (red) composite of the difference between the curve fit and the measurements.
3.1.3.
Application of OMA-EMIF Method to HPSD
The OMA-EMIF method was also applied to the simulated measurements from the Mathieu oscillator. In doing so, the full power spectrum matrix for p=2 was used (rather than the column used in the previous two subsections). A CMIF plot was first created from the HPSD matrix, which is shown in Figure 3. The CMIF plot is far more convenient to view than the individual elements of the harmonic power spectrum matrix, which were shown in [5, 6], yet it retains information regarding the mode shapes of the system that can be used to detect modes with close natural frequencies. In this plot, the peak at 0.982 Hz is the Floquet exponent of the system and other peaks of the same amplitude that appear in the plot are modulations of the Floquet exponent, λ+inωA, for n = -2,…2, where ωA=0.8 rad/s. Other weaker peaks are also visible in the CMIF for other values of n.
38 CMIF of HPSD of Mathieu Oscillator Response
4
10
2
10
0
10
-2
10
-4
10
-6
10
0
0.5
1
1.5
2
2.5 Freq (Hz)
3
3.5
4
4.5
5
Figure 3: CMIF of the HPSD of the Mathieu oscillator response. The individual curves correspond to different singular values of the HPSD matrix at each frequency.
The EMIF methodology was then applied to the measurements, by creating the enhanced positive power spectrum and then fitting low-order modal models to them. The modal paramters of each of the peaks were identified, but for brevity only the mode identified at the 0.982 Hz peak will be considered. 3.1.4 Comparison of Numerical Results
Table 1 shows the result of parameter estimation using various methods. The column labeled “Peak Picking” shows the result obtained by peak picking, which was previously published in [6]. Each of the peaks in the HPSD, pHPSD and CMIF were fit by each method, and they all may contain useful information regarding the mode shapes and eigenvalues of the LTP system. However, for brevity only the result obtained from the primary peak near 0.98 rad/s is compared.
39
Fourier Coefficients of the Mathieu Oscillator Mode Shape C1,n
n -2 -1 0 1 2 λ ωn ζ
Analytical 0.067 -0.220 1 0.092 0.00320
Peak Picking 0.0650 - 0.0090i -0.220 + 0.0061i 1 0.0906 + 0.0011i 0.0028 - 0.0001i
Fit to Harmonic PSD 0.0667 + 0.00233i -0.217 - 0.00151i 1 0.0910 - 0.00126i 0.00340 +0.00050i
Fit to Positive HPSD 0.0652 + 0.00095i -0.219 - 0.00113i 1 0.0902 - 0.00212i 0.00308 +0.00029i
OMA-EMIF 0.063 – 0.014i -0.219+0.009i 1 0.094 + 0.001i 0.003 + 0.0004i
-0.0200 + 0.987i
-0.0270 +0.986i
-0.0169 +0.987i
-0.0172 + 0.988i
-0.0124 + 0.988i
0.987
0.986
0.987
0.988
0.988
0.0203
0.0274
0.0171
0.0174
0.0123
Table 1: Mode shape coefficients identified at the fundamental peak using various methods. The analytical vector C1,0 is shown in the first column.
The results obtained by each of these methods are quite similar. The only large discrepancy between the methods occurs for the damping ratio, which is poorly estimated using the peak picking method as expected. The results obtained using the HPSD and the positive HPSD seem to have similar quality, even though the curve fit on the positive power spectrum seemed to be far more satisfying. The OMA-EMIF agrees well with the other two curve-fitting methods. The result is somewhat surprising if the fit to one of the weak peaks is compared. Table 2 compares the Fourier coefficients obtained at the 2.6 rad/s peak in Figures 1 and 2. This peak was very weakly represented in the measurements, and was seen to be almost buried in the residual in the HPSD curve fit in Figure 1. The estimates of the Fourier coefficients produced by the peak picking method are up to four times larger than the true coefficients, but the curve fitting methods give results with even greater errors. This anomaly is readily explained. The measurements from which the coefficients 2, 3 and 4 were obtained were plotted (not shown here) and it was observed that none of those measurements showed any visual evidence of a mode near 2.6 rad/s. The spurious values seen in the table seem to arise as the least squares procedure attempts to fit a mode to the shoulder of one of the more prominent peaks. On the other hand, a weak peak was visible in the measurement from which the n=1 coefficient was estimated, but even there the peak picking method seems to have achieved a better estimate of the Fourier coefficient. n 0 1 2 3 4 λ ωn
Fourier Coefficients of the Mathieu Oscillator Mode Shape C1,n Analytical
Peak Picking
Fit Harmonic PSD
Fit Positive HPSD
1 0.092 0.003197 0.000039 -0.000015
1 0.0918 + 0.00090i 0.0136 - 0.00054i 0.000824 - 0.000289i 0.000053 - 0.00021i
1 0.0626 + 0.0246i -0.32764 + 0.055243i -0.0137 - 0.0133i -0.000688 + 0.00072i
1 -0.0701 + 0.00662i 0.940 + 0.184i -0.112 - 0.0409i -0.0152 - 0.00348i
-0.0200 + 2.587i
-
-0.0185 + 2.592i
2.587
2.588
2.592
-0.0223 +
2.593i
2.593
Table 2: Mode shape coefficients identified at the 2.6 rad/s peak in the HPSD and pHPSD using various methods.
40
3.2. Simulated Measurements from a Wind Turbine
This section applies the proposed LTP system identification methods to a much more challenging problem, a 5MW wind turbine rotating at constant speed due to aerodynamic forcing. Turbulence in the incoming wind provides a source of broadband random excitation. The system and conditions studied here are identical to those reported in the authors’ prior work [5, 6], where peak picking was used to identify the edgewise modes of the turbine. In that work, some difficulty was reported in distinguishing the forward and backward whirling modes of the turbine since many of their harmonics overlap. Here the advanced LTP system identification algorithms will be applied to that system in an effort to enhance the results. The wind turbine is rotating at a constant rate of 0.201 Hz in these measurements, driven by a wind with logarithmic shear with 18 m/s velocity at hub height. Measurement of the edgewise motion of each of the three blades was simulated, as well as lateral and fore-aft motion of the tip of the tower. The harmonic autospectrum matrix was computed for these five measurements with n=-2…2. The total measurement set is quite difficult to visualize, so two separate CMIFs were created, shown in Figures 4 and 5. Figure 4 shows a CMIF including the two measurement points on the tower and all of their harmonics. As with the results in the previous section, only the primary column of the HPSD matrix was used, so the CMIF is based on a 10x2 matrix. The CMIF clearly shows the presence of the first two modes of the tower, both near 0.27 Hz. The forward whirling mode of the rotor is also visible at 1.16 Hz, and there is a very small peak near 0.76 Hz, where the corresponding backward whirling mode should be. A harmonic of the forward mode also seems to be visible near 1.36 Hz. Finally, there is a large triangular shaped peak in the spectrum near the blade-pass frequency of 0.6 Hz. This is due to the aerodynamic forcing function of the system and is not a mode of the structure, as discussed in [6, 16]. Com plex Mode Indicator Function
-1
Magnitude of Singular Values
10
-2
10
-3
10
-4
10
0
0.2
0.4
0.6
0.8 1 Frequency
1.2
1.4
1.6
1.8
Figure 4: CMIF of the HPSD for tower motion of wind turbine.
Figure 5 shows the CMIF using the three edgewise blade measurements as references. In this spectrum, contamination is clearly visible at the rotor speed and its harmonics: 0.2 Hz, 0.4
41
Hz, 0.6 Hz and (diminished) at 0.8 Hz. The dominant peaks in the spectrum are the edgewise modes of the turbine near 0.96 Hz. These are the same forward and backward whirling modes that were seen in Figure 4, but the blade measurements were taken in the rotating reference frame so those two modes appear at the same frequency in these measurements. The second singular value seems to show a coherent peak at 0.96 Hz, indicating that two modes are present at this peak. Each of these modes are expected to appear at modulations of this frequency: 0.96 + n0.201Hz as well. Any modulations below 0.96 Hz would be obscured by the harmonics of the rotor frequency, but two harmonics are visible at 1.16 and 1.36 Hz. However, only one singular vector is visible at each of those peaks, suggesting that one of the two modes is obscured by noise at those frequencies due to the non-ideal nature of the input spectrum. In order to quantify these modes, the OMA-EMIF algorithm was applied to the peaks at 0.96 Hz and 1.36 Hz. Two modes were identified at the former frequency and one at the latter with natural frequencies 0.9623, 0.9636, and 1.366 Hz. The corresponding mode vectors are shown in Table 6. The magnitude and phase (in degrees) of each Fourier coefficient is shown. The first mode (0.9323 Hz) shows significant harmonics at n=0 and n=2, the latter being about 67% of the former in magnitude. The other harmonics seem to be caused by noise in the measurements. In contrast, the second mode (0.9636 Hz) only has significant Fourier coefficients for n=0; all of its other Fourier coefficients seem to be at the level of the noise. The coefficients identified at 1.366 Hz are due to a modulation of one or both of the first two modes by twice the rotational frequency. Their Fourier coefficients have been shifted (e.g. to account for l in eq. (10)) so that they can be compared with the other two modes. The resulting harmonics agree very well with those for the first mode in both magnitude and phase. This suggests that the mode identified at 1.366 Hz is merely a replicate of the 0.9323 Hz mode due to the modulation in eq. (7). These results agree very well with what would be expected for a turbine such as this. The phases of the n=0 harmonics of the first mode decrease from blade 1 to 2 to 3, revealing that this is a forward whirling mode. Its n=2 harmonics show phase angles that increase from blade 1 to 2 to 3, indicating that the n=2 harmonics are a backward whirling component of this forward whirling mode. This same behaviour was observed for a simulated turbine in [2] and arises due to the anisotropy of the tower. The phases of the n=0 harmonic of the second mode reveal that this mode is a backward whirling mode. None of the other harmonics for that mode show the expected phase relationship, so they are all most likely due to noise.
42 Com plex Mode Indicator Function
3
Magnitude of Singular Values
10
2
10
1
10
0
10
-1
10
0
0.2
0.4
0.6
0.8 1 Frequency
1.2
1.4
1.6
1.8
Figure 5: CMIF of the HPSD for edgewise motion of the three blades of the wind turbine. 0.9623 Hz, M1 0.9636 Hz, M2 1.366 Hz Mag Phs (o) Mag Phs (o) Mag Phs (o) n Blade 0.023 -174 0.123 -158 -2 B1 0.024 -179 0.142 -165 -2 B2 0.024 -175 0.116 -171 -2 B3 0.017 -4 0.040 19 -1 B1 0.006 139 0.036 110 -1 B2 0.018 -78 0.028 -88 -1 B3 1 0 1 0 1 0 0 B1 0.957 -121 1.039 119 0.948 -122 0 B2 0.964 122 1.045 -117 0.946 122 0 B3 0.025 -117 0.015 -123 0.022 -117 1 B1 0.022 -114 0.032 54 0.020 -108 1 B2 0.015 -135 0.036 -29 0.016 -126 1 B3 0.068 -102 0.006 50 0.072 -102 2 B1 0.070 21 0.011 0 0.072 21 2 B2 0.067 138 0.009 15 0.069 138 2 B3 0.002 -73 3 B1 0.003 125 3 B2 0.008 64 3 B3 Table 6: Fourier coefficients identified from the HPSD using the OMA-EMIF method.
4. CONCLUSIONS
This paper has extended several advanced methods for operational modal analysis to linear time periodic systems. Curve fitting methods were demonstrated both on the harmonic autospectrum and the positive harmonic autospectrum. The latter were found to be simpler to curve fit and to produce good agreement between the reconstructed and measured positive autospectra. Even then, both methods were found to identify natural frequencies, damping ratios
43
and mode shapes of similar quality. On the other hand, when the curve fitting methods were applied to very weak peaks in the harmonic autospectra, both methods gave overly large estimates for the Fourier coefficients that were buried in the noise. In this respect, the peak picking method was preferred for these very weak peaks. In any event, it seems advisable to question the accuracy of any Fourier coefficient that is not well represented in the spectrum. The methods were also applied to simulated measurements from a 5MW wind turbine which was rotating at constant speed. Complex Mode Indicator Functions were formed from the measurements and shown to reveal many of the modes of the turbine, including the first edgewise modes of the turbine, which are of special interest since they tend to be lightly damped and hence can limit the life of the turbine. The OMA-EMIF method was also applied to the measurements and was successful in extracting the forward and backward whirling modes, both of which appeared at the same frequency in the blade measurements. This method proved to a significant aide; in the authors previous work [5, 6] the shapes identified by peak picking were shown and with that method it was very difficult to separate those two modes. 5. ACKNOWLEDGEMENT
The first author was supported by the National Science Foundation under Grant CMMI0969224 while completing this work. 6. REFERENCES
[1] [2] [3] [4] [5]
[6] [7] [8] [9] [10]
K. Stol, M. Balas, and G. Bir, "Floquet modal analysis of a teetered-rotor wind turbine," Journal of Solar Energy Engineering-Transactions of the ASME, vol. 124, pp. 364-371, Nov 2002. P. F. Skjoldan and M. H. Hansen, "On the similarity of the Coleman and LyapunovFloquet transformations for modal analysis of bladed rotor structures," Journal of Sound and Vibration, vol. 327, pp. 424-439, 2009. I. Bucher and O. Shomer, "Detecting asymmetry in rotating structures a combined acutation and signal processing approach," in 23rd International Modal Analysis Conference (IMAC XXIII) Orlando, Florida, 2005. M. S. Allen and M. W. Sracic, "A New Method for Processing Impact Excited Continuous-Scan Laser Doppler Vibrometer Measurements," Mechanical Systems and Signal Processing, vol. 24, pp. 721–735, 2010. M. S. Allen, M. W. Sracic, S. Chauhan, and M. H. Hansen, "Output-Only Modal Analysis of Linear Time Periodic Systems with Application to Wind Turbine Simulation Data," in 28th International Modal Analysis Conference (IMAC XXVIII) Jacksonville, Florida, 2010. M. S. Allen, M. W. Sracic, S. Chauhan, and M. H. Hansen, "Output-Only Modal Analysis of Linear Time Periodic Systems with Application to Wind Turbine Simulation Data," Mechanical Systems and Signal Processing, vol. submitted Oct. 2010, 2010. M. O. L. Hansen, J. N. Sørensen, S. Voutsinas, N. Sørensen, and H. A. Madsen, "State of the art in wind turbine aerodynamics and aeroelasticity," Progress in Aerospace Sciences, vol. 42, pp. 285-330, 2006. S. Yang and M. S. Allen, "Output-Only Modal Analysis Using Continuous-Scan Laser Doppler Vibrometry and Application to a 20kW Wind Turbine," in 29th International Modal Analysis Conference (IMAC XXIX) Jacksonville, Florida, 2011. S. Chauhan, R. Martell, D. L. Brown, and R. J. Allemang, "Considerations in the Application of Spatial Domain Algorithms to Operational Modal Analysis," in International Seminar on Modal Analysis Leuven, Belgium, 2006. J. S. Bay, Fundamentals of Linear State Space Systems. Boston: McGraw-Hill, 1999.
44
[11] [12] [13]
[14] [15] [16]
M. S. Allen, "Frequency-Domain Identification of Linear Time-Periodic Systems using LTI Techniques," Journal of Computational and Nonlinear Dynamics vol. 4, 24 Aug. 2009. M. Allen and J. H. Ginsberg, "Floquet Modal Analysis to Detect Cracks in a Rotating Shaft on Anisotropic Supports," in 24th International Modal Analysis Conference (IMAC XXIV) St. Louis, MO, 2006. M. S. Allen and J. H. Ginsberg, "A Global, Single-Input-Multi-Output (SIMO) Implementation of The Algorithm of Mode Isolation and Applications to Analytical and Experimental Data," Mechanical Systems and Signal Processing, vol. 20, pp. 1090–1111, 2006. M. S. Allen and J. H. Ginsberg, "Global, Hybrid, MIMO Implementation of the Algorithm of Mode Isolation," in 23rd International Modal Analysis Conference (IMAC XXIII) Orlando, Florida, 2005. R. J. Allemang and D. L. Brown, "A Unified Matrix Polynomial Approach to Modal Identification," Journal of Sound and Vibration, vol. 211, pp. 301-322, 1998. D. Tcherniak, S. Chauhan, and M. H. Hansen, "Applicability limits of Operational Modal Analysis to Operational wind turbines," in 28th International Modal Analysis Conference (IMAC XXVIII) Jacksonville, Florida, 2010.
Normalization of Complex Modes from Mass Perturbations
Dong-Woo Seo1, Farid Moghim2, Dionisio Bernal3 1
Graduate students, Northeastern University, Civil and Environmental Engineering, Boston, MA 02115, USA 2 Professor, Northeastern University, Civil and Environmental Engineering, Center for Digital Signal Processing, Boston, MA 02115, USA
ABSTRACT Operation modal analysis has become an alternative for structures in which input data is difficult to measure. The sensitivity-based operational mode shape normalization technique was originally introduced by Parloo et al. in 2002. The technique has been widely used and is still receiving a considerable amount of attention by modal analysis researchers. Up to this point attention has been focused on real mode shapes. There are instances, however, where complex modes may have to be used because the modal density or the damping distribution is such that the use of a normal modes model is not warranted. The essential problem in normalizing complex modes using eigenvalue sensitivity is the variability of the real part of the poles. This paper examine two options: 1) neglecting modal complexity and using standard real mode fit and 2) using the complex mode. Results from an experiment using these two alternatives are presented. 1. Introduction The sensitivity-based operational mode shape normalization technique for the scaling of operational mode shapes was introduced by Parloo et al. [1]. In recent years, some variations of this strategy have been proposed [1-4]. The motivation for these studies being to improve accuracy given the experimental limitation with which small changes in eigenvalues can be identified. All of the equations for estimating scale factor developed thus far have the assumption that the mode shapes of the system are real. It is understood, however, that this assumption may not be valid in many cases. The approach originally suggested by Paloo et. al. [1] was not explicitly formulated for the case of complex modes but its extension to this case was conceptually outlined in the original reference and is used here to derive the explicit form of the equation for the scaling in the case of the complex eigensolution [5,6]. In this paper we begin by with an analytical examination of the issue of selecting between real or complex modes and subsequently present some results from an experiment. 2. Mass Change Method 2.1. Some theoretical relations The mass change method consists of performing operational modal analysis on both the original and a modified structure. The modification is carried out by attaching masses to selected locations. The mass matrix of the modified structure can be written as
T. Proulx (ed.), Civil Engineering Topics, Volume 4, Conference Proceedings of the Society for Experimental Mechanics Series 7, DOI 10.1007/978-1-4419-9316-8_4, © The Society for Experimental Mechanics, Inc. 2011
45
46
M
M 0 E M1
(1)
where, M 0 is the mass matrix prior to the perturbation; M 1 is a matrix describing the spatial distribution of the mass modification and E is a scaling parameter with respect to which the rates of change are computed. The arbitrarily normalized mode \ j and the mass normalized one I j are related by a constant
D j , namely
Ij
D j\ j
(2)
The expression that normalizes the modes to unit modal mass based on eigenvalue sensitivity when the eigenvectors are complex is [5]
D 2j
dOj
dE \ M1\ j O j2 T j
(3)
where, O j is the complex eigenvalue. The expression that normalizes the modes to unit modal mass when the eigenvectors are real is [1]
D
2 j
d Z 2j
dE I M1I j Z 2j T j
(4)
where, Z j is the undamped natural frequency and I j is the real arbitrarily scaled mode shape. Eq.3 and Eq.4 are exact but cannot be evaluated without approximation because there is no way to get the exact value of the derivatives using finite difference (unless the mathematical function connecting the eigenvalue to E is known). To minimize the error coming from estimation of the derivatives by finite difference one strives to use as small as possible mass perturbation. Nevertheless, approximation in the identification needs to be considered when selecting the magnitude of the change. Note that variability in the identification may have a larger affect on Eq.3 than on Eq.4 because the real parts of the eigenvalues are known to have a much larger coefficient of variation than the imaginary parts. To get some idea of the possible effect of considering modal complexity we begin by comparing Frequency Response Functions (FRFs) computed using Eq.3 and Eq.4 in the case of a system that is non-classically damped and where the real modes are approximated by normalizing the maximum amplitude of the complex mode to unity and then discarding the imaginary parts. The structure considered is a 3-DOF lumped mass tower where a concentrated dashpot externally grounded is added on the first level. In Fig.1 shows the drive-point FRF for the third coordinate obtained using E 0.01 and 0.1. As expected, the accuracy of the complex mode solution deteriorates some due to the finite difference approximation of the derivative when E = 0.1 but the result is in all cases notably more accurate that the approximation obtained with the normal modes.
47 FRF
0.14
Exact Solution Complex Modes Real Modes
0.1
0.1
0.08
0.08
0.06
0.06
0.04
0.04
0.02
0.02 0.5
1
1.5 2 Frequency (Hz)
2.5
3
Exact Solution Complex Modes Real Modes
0.12
H33
H33
0.12
0 0
FRF
0.14
0 0
3.5
0.5
1
(a)
Figure 1: Comparisons of FRFs among three solutions: (a) is H33 with 0.1
E
1.5 2 Frequency (Hz)
2.5
3
equal to 0.01; (b) is H33 with
3.5
(b)
E
equal to
3. Experiment
3.1. Magnetic damper
In an attempt to introduce modal complexity we added a localized damping force to a 4 story tower structure. The localized damping was realized using magnets that were attached to a fixed structure and copper plates that were affixed to the structure. The principle used is the fact that the motion of the copper plate on the magnetic field induced electrical current. This current then creates its own magnetic field and based on the Lenz’s law the secondary magnetic field opposes the initial field (the field of a moving magnet) and the interaction causes a force that resists the change in magnetic flux which opposite to the relative velocity and is thus dissipative in nature (see Fig. 2).
Figure 2: Time varying magnetic flux [7]
The magnitude of the dissipation increases with: 1) stronger magnetic fields, 2) faster changing fields and 3) using thicker conductive plate and lower resistivity materials, since the movement of the electrons is easier in this case. 4. Experimental Campaign
4.1. Experiment setup
The set up used to create the concentrated dissipation is shown in Fig.3. A copper plate with a thickness of
48
1/8” is used as a conductor sheet. For creating the magnetic field Neodymium magnets were used. These magnets are attached to the piece of the Plexiglas. The Plexiglas is attached to a stand and placed in front of the copper plate which is glued to the floor of the structure.
Figure 3: Close view between the magnets and the structure
The structure considered is the 5-story aluminum tower depicted in Fig.4. The structure is instrumented with 5 accelerometers, one attached one each level. The excitation used to drive the system is a burst random signal supplied by a shaker located on the first level (not shown in the figure).
Figure 4: A 5-story aluminum tower with magnets on the second level
Results for a drive point FRF at coordinate #1 computed using the identified arbitrarily scaled modes normalized using Eqs. 3 and 4 are compared with result experimentally extracted by treating the input as known in Fig.5. As can be seen the results are reasonably good and the match of the complex model is somewhat better than that achieved by the normal modes. It’s worth noting that the added mass used in this case was ȕ = 0.12, which is quite large.
49
10
10
FRF
-3
10 Polymax. Complex Modes Real Modes
-4
10
10
-5
10 10
10
-4
-5
H11
H11
10
FRF
-3
-6
-6
10 -7
0
50
100 Frequency (Hz)
10
150
(a)
-7
-8
0
Polymax. Complex Modes Real Modes 50 100 Frequency (Hz)
150
(b)
Figure 5: Comparisons of FRFs: (a) is with magnets on the second level; (b) is with magnets on the fifth level, both are H11 with E equal to 0.12
To gain some idea on the variability the test previously described was repeated a number of times. The statistics of the results obtained are summarized in Table 1. Table 1: Variability of the complex scaling factor.
Mode Real Part SD CoV 1 0.0853 4.66% 2 0.0822 1.52% 3 0.2401 7.96% 4 0.0544 48.86% 5 0.0701 10.12%
Imag. Part SD CoV 0.0560 0.31% 0.0771 0.13% 0.1731 0.13% 0.1709 0.10% 0.1770 0.08%
5. Concluding Remarks
This study presents a pilot examination of work on the normalization of complex modes obtained from operational modal analysis using the mass change strategy. It was found that while the complex equation can provide improved results when complexity is significant a key issue is the question of identification variability. In particular, the fact that the real part of the complex eigenvalue typically has a large coefficient of variation and thus the estimation of the real part of the eigenvalue derivative is generally very uncertain.
Acknowledgements
This study was initiated as part of the course, CIVG 7343, taught by professor D. Bernal at Northeastern University. The help with the experimental setup provided by Mr. Michael Macneil, the Laboratory Technician of the Department of Civil and Environmental Engineering, is thankfully acknowledged.
50
References
[1] Parloo, E., Verboven, P., Cuillame, P., Vsn Overmerire, M., "Sensitivity-based operational mode shape normalization", Mechanical systems and signal processing, Vol. 16, P: 757-767, 2002. [2] Brinker. R., Andersen. P. A., "Way of getting scaled mode shapes in output-only modal testing", In: 21st International modal anlaysis conference, Kissimmee, Florida, 2003. [3] Bernal, D., "Modal Scaling from Known Mass Perturbation", Journal of engineering mechanics, Vol. 130(9), P: 1083-1088, 2004. [4] Fernández-Canteli M., Brincker, R., Fernández-Canteli, A., "Some methods to determine scaled mode shapes in natural input modal analysis" In: 22nd International modal analysis conference, Feb. 2004. [5] Bernal, D., "Complex eigenvector scaling from mass Perturbation", Submitted to journal of engineering mechanics, 2010. [6] Cadwell, L. H., "Magnetic damping: analysis of an eddy current brake using an airtrack", American journal of physics, Vol. 64(7), P: 917-923, 1996. [7] Wiederick, H. H., Gauthier, N., Campbell, D. A., Rochon, P., "Magnetic braking: simple theory and experiment", American journal of physics, Vol. 55(6) P: 500-503, 1987. [8] Heald, M. A., "Magnetic braking: improved theory", American journal of physics, Vol. 56(6), P: 521-522, 1988. [9] Sodano, H. A., Inman, D. J., Belvin, W. K., "Development of a new passive-active magnetic damper for vibration suppression", Journal of vibration and acoustics, Vol. 128(3), P:318-328, 2006.
Numerical Continuation of Periodic Orbits for Harmonically Forced Nonlinear Systems
Michael W. Sracic Graduate Research Assistant, Ph.D Candidate [email protected] & Matthew S. Allen Assistant Professor [email protected] Department of Engineering Physics University of Wisconsin-Madison 534 Engineering Research Building 1500 Engineering Drive Madison, WI 53706 ABSTRACT A nonlinear structure will often respond periodically when it is excited with a sinusoidal force. Several methods are available that can compute the periodic response for various drive frequencies, which is analogous to the frequency response function for a linear system. The simplest approach would be to compute a sequence of simulations where the equations of motion are integrated until damping drives the system to steady state, but that approach suffers from a number of drawbacks. Recently, numerical methods have been proposed that use a solution branch continuation technique to find the free response of unforced, undamped nonlinear systems for different values of a control parameter. These are attractive because they are built around broadly applicable time-integration routines, so they are applicable to a wide range of systems. However, the continuation approach is not typically used to calculate the periodic response of a structural dynamic system to a harmonic force. This work adapts the numerical continuation approach to find the periodic, forced steady-state response of a nonlinear system. The method uses an adaptive procedure with a prediction step and a mode switching correction step based on Newton-Raphson methods. Once a branch of solutions has been computed, it explains how a full spectrum of harmonic forcing conditions affect the dynamic response of the nonlinear system. The approach is developed and applied to calculate nonlinear frequency response curves for a Duffing oscillator and a low order nonlinear cantilever beam. 1.
Introduction
Many engineering structures exhibit nonlinear behavior. Some common examples include: shock absorbers in automotive suspension systems, bolted joints in aerospace structures and nonlinear elastic tissues in biomechanical systems. Nonlinearities can be due to geometric effects such as buckling, nonlinear constitutive relationships in materials, or dissipative effects such as the damping due to mechanical joints. The nonlinearities greatly affect the dynamic response of a structure and can lead to resonance phenomena that are not predicted by linear theory. One needs to accurately characterize the nonlinear dynamic response to accurately predict the structure’s response and avoid structural failure. Many nonlinear systems are operated under the presence of harmonic forcing that induces a periodic response. Some common examples of harmonically forced nonlinear systems include rotor systems with bearing nonlinearities [1-3], turbomachinery [4, 5], wind turbines [6], and nonlinear electrical circuits [7-9]. The response of the human musculoskeletal system is also highly nonlinear [10-13], and researchers often study neural network control systems as a form of oscillatory control force [14-17]. Even if a system does not naturally operate periodically under harmonic forcing, it is typical to perform experimental vibration testing using harmonic forcing (i.e. sine sweep testing) to better understand its behavior. Whenever a nonlinear system is harmonically forced, the specific forcing configuration can greatly affect the response of the system. For example, a nonlinear system that is excited with a sinusoid may respond harmonically at a number of
T. Proulx (ed.), Civil Engineering Topics, Volume 4, Conference Proceedings of the Society for Experimental Mechanics Series 7, DOI 10.1007/978-1-4419-9316-8_5, © The Society for Experimental Mechanics, Inc. 2011
51
52
frequencies, and the dominant frequency could be different from the forcing frequency. In another case, there may be several possible periodic solutions (stable or unstable) for a single forcing function. For certain forcing configurations, the response of a highly nonlinear system may be approximately linear, while a weakly nonlinear system may exhibit strong nonlinearity. In order to fully understand the dynamics of a nonlinear structure, especially one that will be harmonically forced, it is important to predict and characterize the periodic responses that are produced by each forcing configuration. Many techniques have been developed to calculate the periodic solutions of nonlinear systems, and those techniques usually take one of two basic approaches. The first approach is to approximate the analytical solution to the differential equations. The harmonic balance technique, which is similar to what is done for linear forced oscillations, is a popular approach where one assumes a harmonic series for the solution and usually truncates the series after one harmonic term [18, 19]. This approximate method tends to exhibit large error when the response contains strong higher harmonics; so additional terms can be included in the series, but this makes the mathematical formulation more difficult. The method of multiple scales and the method of averaging are commonly used perturbation techniques where one attempts to find an asymptotic linearization of the nonlinear solution, again using an approximating functional form for an analytical solution. A number of applications of these methods can be found in the following works [18, 20, 21]. All of these methods can provide accurate solutions in certain situations, but they can be difficult to implement for higher order systems that are strongly coupled, they are not straightforward to automate for a large range of forcing configurations, and they involve approximations so considerable expertise may be needed to be sure that those approximations are warranted. The second approach to calculating periodic solutions for nonlinear systems employs numerical methods. Essentially, one deals with the exact nonlinear equations for a system and uses direct time integration to search for the solution. This direct approach is powerful because closed form equations of motion are not required. If a system is damped then one can always calculate the periodic orbit conditions by integrating the equations of motion from a given set of initial conditions for a long enough time that the system settles into a steady periodic orbit. Typically, an adaptive time step control, fourth-order Runge-Kutta method [22, 23] or Newmark method [24] is used for time integration of structural dynamics systems. Although this approach is straightforward, it can be computationally expensive, and methods have been developed over the years to streamline such calculations. In fact, methods that find the periodic solutions of general differential equations are somewhat mature [25]. Two typical approaches are to evaluate and update a cost function with optimization [26, 27] or to solve a boundary value problem using a NewtonRaphson correction method [25]. The latter approach is widely used because the Newton-Raphson method converges quadratically. These approaches are usually started with an initial guess, which is subsequently corrected to satisfy convergence criteria, and this overall procedure is generally termed a shooting method. A parameter of interest can then be varied and the shooting method applied multiple times to compute a branch of solutions, and this is termed a continuation method. A general overview of shooting and continuation techniques is provided in [28]. Generally speaking, the methods for forced continuation can be categorized as corrector techniques, predictor-corrector techniques, or adaptive predictor-corrector techniques. Correction techniques use the general shooting technique to calculate solutions. Continuation is commonly performed with these methods by incrementing the forcing frequency (perhaps sequentially), using the previous solution as the initial conditions for the shooting method calculation of the solution. This procedure is improved by using a predictor-corrector algorithm, where each solution is used to calculate a prediction for the next solution on the curve. A prediction can be made for both the new forcing frequency and the new periodic response values. Then, the prediction is used as the initial guess for some form of shooting method that calculates the actual solution, and the convergence is usually much faster than without a prediction. Predictions can be made with a number of different geometric calculations, for example by calculating the tangent to the branch of solutions. More advanced methods will use an adaptive predictor-corrector algorithm, where the number of convergence calculations are recorded and used to adapt the prediction increment size between successive solutions, which further improves the efficiency of the calculations. A few well developed packages exist that use various forms of adaptive predictor-corrector methods for nonlinear system calculations. AUTO [29] is a code that is available for download. It has many sophisticated computational methods for nonlinear systems, but it does not contain a straightforward algorithm for calculating the periodic solutions of harmonically forced nonlinear systems. MATCONT [30] is another well developed package. It can be run in MATLAB and has a graphical user interface, but it is more suited towards calculating periodic solutions of autonomous systems. These codes are freely available for use but are not suited for all problems, so many researchers are still developing their own algorithms. A few of these use a simple corrector approach. In [31], a shooting method was applied to find periodic solutions of autonomous nonlinear systems. A single initial guess was used to calculate an entire branch of solutions for one parameter of interest, as opposed to performing some form of continuation from one solution to the next. In [23], a continuation method was used to calculate the forced-response periodic displacements of the center of a low-order model of a beam with clamped ends. A secant predictor step was used to vary the forcing frequency, but no predictions were calculated for the response solutions. The method showed promising results, but the algorithm could not follow sharp turns of the periodic solution branch. In [32], the harmonic balance method was used to calculate the periodic solutions of a reduced order torsional subsystem. Then, an
53
adaptive predictor continuation algorithm was used to find additional solutions as the operating frequency was varied. This method produced very detailed periodic solution curves but still requires one to use the approximations required by the harmonic balance method. If too few terms are used in the expansion then the algorithm may not predict important nonlinear features, as noted in [32]. A form of predictor-corrector continuation was also used in [22, 33-35]. They incremented the forcing frequency sequentially and then calculated a prediction for only the state vector with a first order Euler expansion from the previous solution. The algorithm provides quite robust performance. For example, in regions where a periodic solution branch has a turning point (i.e. a multiple solutions region), the algorithm included a re-parameterization of the solution curve in order to follow the turning point. The algorithm was applied to a few low order systems and showed promising results. More recently, a similar form of predictor-corrector continuation was presented in [24]. This method employs a pseudo-arc length continuation procedure that has a tangent predictor step and orthogonal corrector steps. Furthermore, the algorithm automatically controls the step size and exploits symmetry in the response to increase computational efficiency. So far, the authors of this method have computed the periodic solutions of a number of unforced, undamped nonlinear systems of high complexity and high order [24, 36-42] with exceptional detail. However, they have not extended their methods to the calculations of periodic solution branches of harmonically forced nonlinear systems. The goal of this work is to create a continuation method for efficiently calculating the periodic solutions of harmonically forced nonlinear systems. This work will utilize techniques primarily from [24] in order to do this. A Newton-Raphson correction technique will be used to calculate the first solution from an initial guess that the user must supply. In order to calculate the subsequent solutions, an adaptive predictor step is used with a mode-switching correction step that employs two additional Newton-Raphson correction methods. The forcing frequency is the parameter that is varied. The most important difference of this work from the previous methods is that the prediction calculation and the Newton-Raphson correction methods includes extra terms that quantify how the periodic orbit varies as the forcing frequency is varied. Additionally, the forced periodic orbit is uniquely determined by the phase of the forcing, so the periodic orbits are calculated with respect to zero phase on the force, and since the frequency of the forcing is varied, the equations of motion change with each periodic orbit calculation; this is accounted for by the proposed method. The procedure of the proposed continuation method will first be outlined. Then, the shooting method and the Newton-Raphson correction method will be reviewed, and the necessary changes of those procedures due to the harmonic forcing will be derived. The continuation procedure will then be explained in detail. In Sections 3 and 4, the algorithm will be applied to calculate the periodic solutions of one and two degree of freedom systems, respectively. The results will be used to better understand the nonlinear systems, and discussion will be provided regarding the performance of the algorithm. Conclusions will be provided in Section 5. 2.
Theory
A nonlinear system with input u(t) can be described by the following state space differential equation z
f z (t ), u (t )
(1)
where z \ n is the time dependent state vector, u \ p is a time dependent input vector, and t is the time variable. For T structural dynamic systems the state variable z is the vector valued function of time z ª¬ x T x T º¼ composed of the N displacement degrees of freedom (DOF) x \ N and the N time derivatives of the displacement DOF, x , where n=2*N. The superscript ‘T’ denotes the matrix transpose operator. It is assumed that f is a C1 (at least one-time continuously n
differentiable) nonlinear function defined on an open subset of \ that contains all the possible positions and velocities of the system of interest. For a specific input, the vector field f generates the flow Zt \ n that contains families of solution curves of the differential equation [43]. A single solution curve is defined by its initial state vector z0 and input vector u0 and by tracing the flow Zt(z0) for a finite time. This work considers systems with harmonic forcing so u (t T )
u (t ) and hence z
f z , T , where the dependence of the
system on the fundamental period T is explicitly shown. Under these conditions and when the system reaches steady state, it is assumed that the flow contains at least one periodic orbit J with period T, and z (t T ) z (t ) for any state z J . (From here forward, the overbar will be dropped from the state vectors when talking about states from a periodic orbit in order to simplify the notation.) Depending on the configuration of the system and its parameters, including the forcing frequency and amplitude, the flow may contain multiple periodic orbits [43, 44]. Tracking the periodic orbits in a nonlinear system can be difficult because a small change in the parameters of the system, such as the forcing frequency, may drastically change the dynamics of the flow. For example, a stiffening system has a frequency response peak that bends to higher frequencies, and
54
the response can jump from a high amplitude response to a very low amplitude response as the frequency is increased beyond the critical value. In order to characterize all of the periodic orbits that exist in the flow of the differential equation, one must calculate each pairing of initial state vector and period (z0, T) of the orbits. This work presents a continuation method that is used to calculate many of the periodic orbits of the differential equation, thereby producing a frequency response function. The general procedure is outlined below and shows the different steps of the algorithm. 1. Provide a starting guess: (z0,(0),T(0)) o The method begins after the user supplies (z0,(0), T(0)) where the subscript zero in parentheses designates that these conditions are the starting initial conditions. 2. Calculate a starting solution: (z0,(1),T(1)) o The proposed continuation method must be started from an actual periodic orbit of the system. To calculate this first solution, a Newton-Raphson correction method (NRCM 1), which accounts for changes in both the state vector 'z0 and the period 'T, can be employed to calculated the first periodic orbit solution (z0,(1),T(1)). 3. Initiate the predictor-corrector procedure to calculate a branch of solutions: (z0,(j),T(j)), j=1,2,3,4,… a. Prediction Step: (z0,pr,(j+1),Tpr(j+1)) o The current solution is used to calculate a prediction (z0,pr,(j+1), Tpr(j+1)) for the initial conditions of the next periodic solution. The prediction is then corrected to find a true solution near the predicted one. (k ) (k ) b. Correction Step: ( z0,( j 1) , T( j 1) ) , k =1,2,3,… o
(0) (0) The prediction step is used as the initial guess, ( z0,( j 1) , T( j 1) ) , for the second Newton-Raphson procedure.
The second Newton-Raphson correction method (NRCM 2) includes all the equations from NRCM 1, but it has an additional equation that constrains the corrections to be orthogonal to the prediction. NRCM 2 is (k ) (k ) th used to calculate corrections to the initial conditions, ( z0,( j 1) , T( j 1) ) , until the (j+1) periodic solution is o o
obtained. If a solution is obtained: After each (j+1)th solution is calculated, Step 3 is repeated. If a solution is not obtained: If the convergence criteria are not met within some pre-set number of iterations: 1. The procedure calculates a new prediction, (z0,pr,(j+1),Tpr(j+1)), using a smaller step and the correction portion of Step 3 is restarted. 2. The procedure switches to a third form of the Newton-Raphson correction method (NRCM 3). This form accounts for changes in the state vector 'z0 only, so that the period is held fixed (i.e. 'T=0), and there are no geometric constraints placed on the corrections. The prediction initial condition is used for an initial guess, and the NRCM 3 is used to calculate the periodic solution. When the solution is calculated, Step 3 above is repeated.
The details of this procedure will be provided below, but first the general Newton-Raphson correction method is derived. Some parts follow the derivation from [24], but there are several differences that arise due to the forcing, which causes the function f in Eq. (1) to depend explicitly on time, t, and on the period, T, of the forcing frequency. 2.1
Shooting Function
The proposed method is an iterative process that seeks to find the periodic solutions of a system for different forcing frequencies. The shooting function is created to quantify whether a certain set of initial conditions, (z0, T) produce a periodic response. First, the forcing frequency is set to :=2S/T and Eq. (1) is integrated over the interval (0, T) from an initial state vector z0. Then, the shooting function H \ n is defined as H z0 , t , T { zT ( z0 , t , T ) z0
(2)
where zT=z(z0, t=T,T) is the state vector found by integrating with initial condition z0 until t=T. This equation represents a two-point boundary value problem for the periodic orbit of the system where the forcing has zero phase at t=0 and t=T. Eq. (2) will be satisfied when H z0 , T ^0` . In practical numerical computation, the shooting function will be satisfied when H ( z0 , T ) | ^0` for some chosen convergence criterion. In this work, the convergence criterion defined in [24] is used
55
H z0 , T
zT z0
z0
z0
H
(3)
where H =1e-6 is the convergence value. Each time the equations of motion are integrated over the assumed period, the shooting function is subsequently calculated and the convergence is evaluated. 2.2
Newton-Raphson Correction Methods
When the given calculation of the shooting function does not satisfy the convergence criterion, a Newton-Raphson [25] method is used to calculate updates 'z0 and 'T to the initial conditions. Following the approach used in [24], the shooting function can be expanded in a Taylor series about z0 'z0 , T 'T ,
H z0 , t
T ,T
wH wH 'z0 'T wz0 z ,T wT z ,T 0 0
^0` ,
(4)
where only the linear terms have been retained. The partial derivatives in the Taylor expansion must be carefully calculated. The first partial derivative, > wH wz0 @ \ nxn , has the same form as in [24], and can be expanded as wz ( z0 , t , T ) >I @ wz0 t T
wH wz0 z ,T 0
(5)
where [I] is the nxn identity matrix. The time varying Jacobian matrix > wz wz0 @ \ nxn , which is the variation of the state vector at time t due to a small change in the initial conditions, can be calculated using one of two methods described in detail in [24]. The first method, which uses a finite differences approach, can be used even when the equations of motion are not available in closed form, but it is more computationally expensive so it is not used in this work. The second approach is to form the following system of time varying differential equations by differentiating Eq. (1) with respect to z0, switching the order of differentiation on the left hand side, and applying the chain rule on the right hand side. d § wz ( z0 , t , T ) · ¨ ¸ wz0 dt © ¹
ª wf ( z , t , T ) º ª wz ( z0 , t , T ) º « »« » wz wz0 z z (t ) ¼» ¬ ¼ ¬«
(6)
In this equation, the Jacobian matrix > wf wz @ \ nxn is simply the linearization of f about the state z(t) for each time instant t in the interval (0, T). Since, the state response was calculated on the interval (0, T) in order to evaluate the shooting function, it can be easily stored in order to integrate the system of Eq. (6) from initial conditions > wz ( z0 , t 0, T ) wz0 @ > I @ . The solution matrix > wz wz0 @ can then be evaluated at t = T and used to calculate > wH wz0 @ . The vector ^wH wT ` \ n in Eq. (4) differs from the result presented in [24], since the function f depends explicitly on t and
T. This partial derivative can be expanded using the chain rule as wH wT
wH wt
t T
wH wT
(7) t T
where the first term represents the change in the shooting function due to a change in the instant that is taken to be the end of the period and can be calculated from Eq. (1) as in [24]. wH wt
f ( zT , t t T
T ,T )
(8)
56
The second term in Eq. (7) was not present in [24], but arises here due to the forcing since f ( z , t , T ) has explicit dependence on the period of the forcing, T. The term represents the change in the shooting function due to a change in the period of the input.
wH wT
t T
wz ( z0 , t , T ) wT t
(9) T
Note that the second term of the shooting function does not appear in either of these expressions because ^wz0 wT ` {0} . Using the same approach that was used to derive Eq. (6), one can derive the following system of time varying differential equations that can be solved to calculate ^wz wT ` . d § wz ( z0 , t , T ) · ¸ dt ¨© wT ¹
ª wf ( z , t , T ) º wz ( z0 , t , T ) ½ wf « »® ¾ wz wT ¿ wT «¬ z z (t ) »¼ ¯
(10)
The term ^wf wT ` \ n has arisen because f ( z , t , T ) and z ( z0 , t , T ) both have explicit dependence on the period of the forcing, T. Hence, this second set of differential equations must be integrated from initial conditions ^wz ( z0 , t T , T ) wT ` f ( zT , t T , T ) over the interval (0,T) in order to compute ^wz ( z0 , t , T ) wT ` . After calculating partial derivatives in Eq. (4), an algebraic system of equations can be formed to solve for an update ('z0, 'T) to the initial conditions. These partial derivatives are used in slightly different ways at different steps in the continuation algorithm, as explained below. 2.3
Continuation Procedure
2.3.1 Provide a starting guess: (z0,(0),T(0)) In order to start the continuation procedure, the user must supply a starting guess. A good starting point guess (z0,(0),T(0)) consists of a low frequency (i.e. a long period) solution for the underlying linear system (i.e. for T(0) the underlying linear system can be solved for the periodic conditions z0,(0)). This is often called a Homotopy approach [18, 25]. 2.3.2
Calculate a starting solution: (z0,(1),T(1))
The starting guess is used to calculate a set of initial conditions (z0, T) that lead to a shooting function, H z0 , T , that satisfies convergence. To calculate the starting solution, the system is integrated from the starting guess, the shooting function is calculated, and the convergence is evaluated. If the initial guess does not satisfy convergence, then the partial derivatives are calculated and the following system is formed to calculate corrections to both the state vector 'z0 and the period 'T. This system is the NRCM 1 mentioned above. ª « wH « wz0 z ,T 0 ¬
º 'z wH » ® 0 ½¾ wT z ,T » ¯ 'T ¿ 0 ¼
H ( z0 , t T , T ) ½ ® ¾ 0 ¯ ¿
(11)
The right hand side vector contains the vector value of the shooting equation. The new initial conditions are (z0,(1)= z0,(1)+'z0, T(1)= T(0)+ 'T). Since the period has been updated the harmonic forcing frequency must also be updated in the equations of motion (i.e. :update=2S/( T(1)) such that u=u(:updatet). Then, these new conditions are used to integrate the equations of motion and check the shooting function for convergence. One then repeats this two stage process until the desired convergence is achieved at which point the first periodic orbit solution (z0,(1),T(1)) has been calculated. The Newton-Raphson corrector method is a local technique, so the starting guess, (z0,(0),T(0)), must be close to the actual solution or the correction scheme will diverge. The first periodic orbit solution and the Jacobian matrices that were used to produce it can be used to initiate the next step.
57
2.3.3
Initiate the predictor-corrector procedure to calculate a branch of solutions: (z0,(j),T(j)), j=1,2,3,4,…
In order to calculate an entire branch of solutions, an adaptive pseudo arc-length continuation [28] method is employed that uses a prediction step which is followed by a correction step. Predictor-corrector methods can follow complex solution braches because the prediction step follows the direction in which the branch of solutions actually evolves. The prediction step, which is calculated using information from the current periodic solution, is used to calculate new initial conditions for the next periodic orbit. The prediction step can be further adapted to evaluate the convergence of the previous correction steps and adjust the step size accordingly. After the prediction step, the shooting method is employed, but now with a new Newton-Raphson correction method that varies the state vector and the forcing frequency in a constrained direction. 2.3.3.1 Prediction Step: (z0,pr,(j+1),Tpr,(j+1)) The basic process is to use the previous periodic orbit solution, which is denoted as the jth solution, to calculate a tangent predictor step for the (j+1)th periodic orbit conditions, including a prediction for the initial state vector and the forcing period. The prediction ^ P` \ n 1 , which has components ^ Pz ` \ n and PT \ corresponding to the states and period respectively, is the tangent vector to the solution branch and can be calculated with the following system of equations. ª « wH « wz0 § · ¨ z0,( j ) ,T( j ) ¸ «¬ © ¹
^P ` ( j)
º wH » P ^ ( j)` wT §¨ z ,T ·¸ » © 0,( j ) ( j ) ¹ » ¼
ª¬ PzT,( j ) PT ,( j ) º¼
^0`½ ® ¾ ¯ 0 ¿
(12)
T
The components of the left hand side matrix come from the Taylor series expansion of the shooting function, and were previously defined. They characterize the vector space that is normal to the periodic solution curve, and once they are found, the desired tangent vector ^ P` in Eq. (12) is simply the vector that is in the null space of the matrix on the left in Eq. (12). Once the tangent vector {P} has been calculated it can be normalized to unit length, and the predictions for the next periodic solution can be calculated with z0, pr ,( j 1)
Tpr ,( j 1)
z0( j ) s( j ) Pz ,( j )
(13)
T( j ) s( j ) PT ,( j )
(14)
2S Tpr ,( j 1)
(15)
: pr ,( j 1)
where z0,pr,(j+1) and Tpr,(j+1) are the prediction initial conditions for the (j+1)th solution and s(j) is the jth step size. One must be sure to again update the differential equations so that the external forcing reflects the change in frequency (i.e. u=u(:pr,(j+1)t)). The step size is critical to the success and efficiency of the computation. One can fix the step size to a very small value and ensure that very small increments are made between successive solutions, but this can be impractical because computation times may be long. A better approach is to provide an automated step size control algorithm. In this work, the step size control algorithm from [24] is employed. The first step size is provided by the user, while all subsequent step sizes are determined by the following equation s( j)
sign s( j 1) ^ P( j ) `
T
^P ` ¨¨ KK ( j 1)
§
*
©
( j 1)
· ¸¸ s( j 1) ¹
(16)
where K(j-1) is the integer number of iterates that were required to update the previous shooting function solution and K* is an optimal number of shooting iterates which is supplied by the user. The signum function sign(