Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Mechanism and Machine Theory 160 (2021) 104288

Contents lists available at ScienceDirect

Mechanism and Machine Theory


journal homepage: www.elsevier.com/locate/mechmachtheory

A homotopy transformation method for interval-based model


updating of uncertain vibrating systems
Dario Richiedei∗, Iacopo Tamellin, Alberto Trevisani
Department of Management and Engineering, University of Padova, Italy

a r t i c l e i n f o a b s t r a c t

Article history: This paper proposes a novel indirect, two-stage approach for model updating in linear vi-
Received 7 October 2020 brating systems, exploiting measured natural frequencies, antiresonances and uncertain or
Revised 5 February 2021
incomplete data of the mode shapes. In the first stage, the technique relies on the partial
Accepted 9 February 2021
eigenstructure assignment paradigm and recasts model updating into a non-linear, non-
Available online 24 February 2021
convex minimization that simultaneously updates the mass and stiffness matrices. Uncer-
Keywords: tainty on mode shapes is formulated through interval (bounds) where they should be-
Model updating long. The inverse eigenvalue problem is solved through homotopy optimization, improved
Natural frequencies through variables lifting and McCormick’s constraints. The unknown parameters are nor-
Antiresonance frequencies malized introducing Jacobian matrices, computed through the complex step derivatives,
Mode shapes to improve the numerical conditioning of the problem and speed up the computation. In
Vibrating systems the second stage, the damping matrix is identified through the generalized formulation of
Multibody systems proportional damping provided by the Caughey’s series.
Two challenging experimental test-cases are solved and prove the method effectiveness:
the linearized multibody model of a flexible manipulator and a structure made by a can-
tilever beam plus a lumped spring-mass system.
© 2021 Elsevier Ltd. All rights reserved.

1. Introduction

1.1. Motivations and state of the art

Dynamic models of linear vibrating systems, expressed through their mass, damping and stiffness matrices, are widely
adopted in model-based approaches to vibration control [1], state estimation [2], synthesis of motion profiles [3], mechanical
design [4], or even “hybrid techniques” that solve simultaneously more of the mentioned problems [5,6]. Effective design is
obtained if accurate dynamic models are available. The method of estimating the correct model parameters matching some
suitable measurements of the system properties is known as model updating (MU) in the field of dynamics of structures
and machines. A comprehensive review of some of the proposed methods can be found, for example in [7] and [8].
MU techniques are usually formulated in the frequency domain by means of modal analysis, although time-domain iden-
tification is sometimes adopted too [9]. Among the different techniques exploiting modal analysis, several papers exploit
the measured natural frequencies to perform MU (see e.g. [10,11]). This choice has the paramount advantage that measur-
ing natural frequencies, especially in lightly damped systems, is very simple and can be done with negligible error and


Corresponding author.
E-mail addresses: dario.richiedei@unipd.it (D. Richiedei), iacopo.tamellin@unipd.it (I. Tamellin), alberto.trevisani@unipd.it (A. Trevisani).

https://doi.org/10.1016/j.mechmachtheory.2021.104288
0094-114X/© 2021 Elsevier Ltd. All rights reserved.
D. Richiedei, I. Tamellin and A. Trevisani Mechanism and Machine Theory 160 (2021) 104288

uncertainty. Either direct, i.e. based on a one-step analytical solution [10], or indirect, i.e. based on iterative computations
[11] methods have been proposed.
Mode shapes provide an important information to be exploited in MU too, and have been often included especially by
indirect methods [11]. Indeed, mode shapes are useful to avoid models that only apparently fit the results of experimental
modal analysis by just matching the natural frequencies. Additionally, mode shape-based methods are often less affected by
environmental conditions, such as temperature [12]. Nonetheless, mode shapes provided by experimental modal analysis are
intrinsically uncertain vectors. In contrast, natural frequencies are usually known with higher precision and accuracy.
The first uncertainty source on the mode shapes is related to the need to extract them from the measurements or more
frequently from the receptances obtained through signal processing. The extraction algorithms are usually very complicate
(see e.g. [13].) and might impose skilled operators for obtaining reliable and robust estimations that get rid of noise, mea-
surement error, numerical error, spurious dynamics and other perturbations.
A second uncertainty source is related to the fact that MU techniques, also including the one proposed in this work,
reasonably use the normal modes that allows studying the system and performing MU through real eigenvector. Indeed,
normal modes are modal vectors whose phase angles are either 0 or 180°. Such an assumption, that usually fits very well
with lightly damped systems, requires a further step to estimate the real, normal modes from the complex identified ones,
i.e. by correctly approximating the phase angle scatter. The literature and commercial software for modal analysis proposes
various algorithms to make effective transformation (see e.g. [14].), which is however a simplification of the true system.
Uncertainty also arises whenever some mode shape entries are difficult measure in experimental modal analysis, such
as for example rotational coordinates that several FE models employ. The problem is sometimes overcome through model
reduction, which might lead to loss of information (e.g. nodal inertias can vanish in the MU problem). A more effective
approach is extrapolating them through assumed shape functions or interpolating them through redundant translational
measurements and shape functions [11].
This discussion highlights that the experimental mode shapes should be carefully handled in any MU technique.
Antiresonances, i.e. the frequencies at which the system experiences no vibration at a particular coordinate when it is
excited by a harmonic force at a certain coordinate [15], are sometimes adopted in MU. Some Authors state that antireso-
nances can be identified from experimental receptance with less error than mode [16,17] especially if measured on point
receptances (i.e. when the exciter or the response transducer is located in the same coordinate) due to the alternating pat-
tern of resonances and antiresonances in the frequency spectrum [16]. However, other Authors do not fully agree with this
statement. Indeed, since antiresonance frequencies depend on the input and measurement coordinates, they are sensitive
to the excitation point [18,19]. For example, in the case of continuous system, antiresonance frequency might vary due to
unwanted small changes of the impact point of the instrumented hammer. Additionally, in [19] it is shown that the positive
contribution of antiresonances strongly depends on the geometry of the model and the choice of variables to be identified
through MU.
All these comments outline two important features that increase the reliability of a MU technique: the capability to
handle natural frequencies, antiresonance frequencies and antiresonance by merging them in the mathematical problem
and the capability to handle uncertainty.
Several works in the literature address uncertainty through statistical analysis, and hence modeling uncertainty as a
random variable with a given probability density (or distribution) function, leading to the so-called stochastic techniques
(see e.g. [20–22]), by contrast of deterministic approaches that assume the measured parameters as correct and hence do
not explicitly handle uncertainty (see e.g. [11]). For example, Bayesian approaches that assign variances to the data have been
often proposed. On the other hand, it is sometimes difficult and time-consuming to correctly represent uncertainty with a
stochastic model and therefore a true statistic can be impractical, since a correct probabilistic model require large volumes
of data [23]. Therefore, the uncertainty distributions are sometimes assumed a-priori, and the different choices affect the
solution. Indeed, uncertainty on measurements can be either random, with different probabilistic model, or systematic.
A third way of handling uncertainty has been also proposed: the use of interval representation of the uncertain pa-
rameters that has been therefore proved to be advantageous whenever accurate representation of the statistics cannot be
obtained [24], or when measurements and estimation are uncertain but with a non-probabilistic nature [25]. With this ap-
proach, just the lower and upper bounds on the magnitude of uncertain parameters and measurements are required. On the
other hand, intervals overcome the limitations of treating the measured quantities as exact values. The theory of interval
analysis has been sometimes adopted to MU to exploit the benefit of this representation of uncertainty. For example, in
[24] MU is applied to stiffness matrix by assuming that some stiffness parameters and the measured frequencies are uncer-
tain. Although interval analysis is a powerful tool, it poses computational issues if many parameters are assumed uncertain
or should be adopted and, as the authors of [24] themselves state, these algorithms are not yet very efficient from a com-
putational point of view and still need to be improved. In [25] a method to update the stiffness matrix has been proposed,
despite the presence of errors in the mass and damping matrix through interval analysis. Again, the issue is the great com-
putational effort. Another critical issue of the standard interval arithmetic is that the operations on intervals sometimes lead
to over-estimated uncertainty bounds [23].

2
D. Richiedei, I. Tamellin and A. Trevisani Mechanism and Machine Theory 160 (2021) 104288

1.2. Contribution of this paper

To overcome the limitations of interval analysis, while exploiting the advantages of a non-probabilistic description of
uncertainty, this paper casts the interval analysis MU problem as a deterministic MU by defining constraints on both the
updating parameters and on the mode shape and formulating an optimization method. Solution is then done through ho-
motopy transformation of the inverse eigenvalue problem and variable lifting, to handle the presence of bilinear terms in
the unknowns, that make non-convex the minimization problem obtained. This strategy has the advantage, over traditional
interval analysis technique, that it is easier to solve through standard solvers for optimization and allows considering among
the unknowns the entries of both the mass and stiffness matrices. Additionally, more general constraints, that are not just
interval, can be introduced, to exploit important properties of vibrating systems such as the pole-zero interlacing properties.
The proposed method is indirect (iterative) since is based on minimizing a function representing the eigenvalue problem.
Conversely with many direct methods, whose solutions can be without physical meaning and have weak robustness to mea-
surement noise, the use of iterative techniques preserve the physical meaning of the updated matrices and allows solving
more complicate problems.
Damping matrix is updated too, in a subsequent step of the method, by exploiting a more general formulation of pro-
portional damping. This damping model, together with the definition of bounds on the mode shapes, make the proposed
method suitable for system experiencing more complicate damping than the usual Rayleigh damping.
The method is also suitable for including antiresonances into MU, together with natural frequencies and mode shape.
Hence, the general frame here proposed is very versatile and can be adopted to update different linear models under various
set of modal data available.
The paper proposes the theoretical development of the method and of its numerical issue and validates it through two
challenging experimental test cases: a structure comprising a cantilever beam with an additional lumped spring-mass sys-
tem, and the linearized model of a multibody system.

2. Method description

2.1. Basic definitions

Let us consider a N-DOF (degree of freedom) linear-time-invariant vibrating system modelled through its mass, damp-
ing and stiffness matrices, denoted M, C, K ∈ RN×N respectively. The displacement vector is q ∈ RN , and q˙ , q̈ ∈ RN are the
velocity and acceleration vectors, and f ∈ RN is the external force vector:
Mq̈(t ) + Cq˙ (t ) + Kq(t ) = f (t ) (1)
The transfer function from the force applied at the c-th coordinate to the displacement√of the r-th coordinate, i.e. the
receptance hrc (s ), is computed as follows (where s = jω: ω is the radian frequency and j = −1):
 
det s2 Mr̄c̄ + sCr̄c̄ + Kr̄c̄
hrc (s ) = (−1 ) r+c
  (2)
det s2 M + sC + K

The minor matrices Mr̄c̄ , Cr̄c̄ , Kr̄c̄ ∈ R(N−1)×(N−1) are obtained by removing the c-th row and the r-th column from matri-
ces M, C, K. The roots of the denominator in Eq. (2) are the system poles that appear in complex-conjugate pairs: pi , p∗i =

−ξ p,i ω p,i ± jω p,i 1 − ξ p,i , where ω p,i is the i th undamped resonance frequency and ξ p,i its damping ratio (subscript p is
adopted to represent the pole features) . The roots of the numerator of Eq. (2) are the  n complex-conjugate pairs of zeros
of receptance hrc (jω ), 0 ≤ n ≤ N − 1, of the and are denoted by: zi , zi∗ = −ξz,i ωz,i ± jωz,i 1 − ξz,i , where ωz,i is the i th un-
damped antiresonance frequency and ξz,i its damping ratio (subscript z is adopted to represent the features of the zeros).
The number of antiresonances, n, depends upon r and c [26,27].

2.2. Problem statement and overview of the proposed method

Let us consider the proportional (or classical) damping assumption, i.e. matrix C is diagonalizable by the eigenvector of
the undamped eigenvalue problem (see e.g. [28]). Systems featuring proportional damping have real vibrational modes, the
normal modes, that are exactly those of the undamped system. Hence, mode shape can be computed through the first-
order eigenvalue problem of the undamped system with real eigenvalue and eigenvectors. In the case of lightly-damped
systems with non-classical damping, whenever damping produces just small phase differences between the system coor-
dinates it is acceptable to approximate the damping matrix through a classical one [29]. In contrast, for heavily damped
systems the damping matrix cannot be approximated through a classical one and mode shapes are highly complex. The
latter problem goes however beyond the scope of this paper, which is targeted to lightly damped systems or systems with
quasi-proportional damping.
Given this assumption, MU is solved in this paper through a two-stage procedure:
1. in the first stage the mass and stiffness matrices are simultaneously updated;
2. in the second stage, the damping matrix is separately identified.

3
D. Richiedei, I. Tamellin and A. Trevisani Mechanism and Machine Theory 160 (2021) 104288

Fig. 1. Sketch of the concept of interval representation of a mode shape in a sample cantilever beam.

The use of two-stage approaches has been often exploited in the literature, since it simplifies the updating of M and K,
leading to more reliable results [30].
Let us first focus on the first stage: the goal is to find the feasible updated matrices M and K such that the updated
system fits the experimental data or improves the original model.
Four problems are solved:

Problem 1. Given a set of experimental natural frequencies and intervals where the related mode shapes are expected to
belong, update M and K such that the natural frequencies of the updated system are as close as possible to the
measured ones.
Problem 2. Given a set of experimental natural frequencies, update M and K such that the natural frequencies of the
updated system are as close as possible to the measured ones.
Problem 3. Given a set of experimental antiresonance frequencies, update M and K such that the antiresonance frequencies
of the updated system are as close as possible to the measured ones.
Problem 4. Given a set of experimental natural frequencies (eventually with intervals where the related mode shapes are
expected to belong) and a set of experimental antiresonance frequencies, update M and K such that the natural
and the antiresonance frequencies of the updated system are as close as possible to the measured ones.

Problems 1 and 4 are those of greatest interest, since merge different information provided by natural frequencies, mode
shapes and antiresonances. Therefore, more attention will be paid to them.

2.3. Formulation of model updating as an inverse eigenproblem for natural frequencies

Let ω p,i denote the i th measured natural frequency and u p,i the i th measured (estimated) mode shape vector related to
it. The following eigenvalue problem should hold:
 
−ω p,i 2 M + K u p,i = 0 (3)

Let us represent the updating of the undamped model through additive matrices M, K ∈ RN×N with respect to the orig-
inal (nominal) ones M0 , K0 ∈ RN×N to be updated. This representation exactly models parameters that linearly affect the
system matrices. The mass and stiffness matrices of the updated model are related to those of the original (nominal) model
as:

M = M0 + M(x ), K = K 0 + K ( x ) (4)
The effect of the nx updating parameters on the updating matrices can be decomposed as the sum of the contribution of
each parameter:

nxm 
nxk
M = M x , K = K x (5)
x=1 x=1

where nx =nxm +nxk , with nxm and nxk denote the number of mass and stiffness unknowns, respectively. A normalized repre-
sentation is provided in Section 3.2. The updating matrices depend on the nx updating parameters x ∈ Rnx to be estimated,
that are the values ensuring that the following eigenvalue problem holds for all the np measured natural frequencies:
 
−ω p,i 2 (M0 + M(x ) ) + K0 + K(x ) u p,i = 0 (6)
The exact solvability of Eq. (6) is not always ensured, especially if the updating parameters are constrained to a feasible
domain. Additionally, errors on u p,i jeopardize solvability and might force the convergence to wrong solutions. Hence, it
could be advantageous replacing the “deterministic” value of the estimated mode shape vector u p,i , or at least of some of
its entries, with intervals where its entries are expected to belong; this idea is sketched in Fig. 1 through a sample cantilever

4
D. Richiedei, I. Tamellin and A. Trevisani Mechanism and Machine Theory 160 (2021) 104288

beam. In the following, all the equations are written as if the whole mode shape is uncertain, to get a simpler formalism;
whenever some entries are exactly known, the extension is obvious.
By replacing the deterministic mode shape u p,i with an interval U p,i = [U p,L , U U ] (U L and U U denote its lower and
i p,i p,i p,i
upper limits), the eigenvalue problem of the updated system in Eq. (6) should be formulated through interval arithmetic
with the matrix enclosures.
Solving this interval function with the methods of interval analysis is very cumbersome. Therefore, this paper transforms
the interval analysis MU problem into a deterministic MU, by defining constraints on both the updating parameters and on
the mode shapes. In contrast, it is assumed that natural frequency errors are negligible. By exploiting the idea of partial
eigenstructure assignment proposed in [31–33] in the frame of dynamic structural modification, Eq. (6) is recast into a
non-linear, least-square problem:
  
np
   x
min α p,i  −ω p,i 2 (M0 + M(x ) ) + K0 + K(x ) u p,i 22 , ∈ p (7)
x, u p,i u p,i
i=1

The unknowns include x and the uncertain mode shape u p,i (or just its uncertain entries in the case of partial assumption
of a deterministic mode shape). The feasible domain p collects the constraints on both x and u p,i . The simplest form of
constraints are lower and upper element-wise inequalities for x and u p,i , denoted respectively xL , xU , U p,
L and U U , such
i p,i
that xL ≤ x ≤ xU and U p, L ≤u
i
U
p,i ≤ U p,i . However other constraints can be defined. Eq. (7) is a non-linear and non-convex
minimization problem because of the presence of bilinear terms of the unknowns, i.e. products of two entries of x and
u p,i . The positive scalars α p,i are adopted to defines differing levels of importance on each natural frequency. The solution
Eq. (7) is here achieved through homotopy optimization.
It should be noted that Eq. (7) can be also exploited to update the model with just the natural frequencies by neglecting
the mode shapes. In this case, it suffices setting the constraint [−1, +1] to all the entries of the eigenvectors, by considering
that mode shapes are defined regardless of their normalization.

2.4. Formulation of model updating as an inverse eigenproblem for antiresonance frequencies

The proposed approach can be also adopted to use antiresonances in MU. Antiresonance frequencies of hrc (jω ), i.e. pairs
of complex-conjugate zeros, are also obtained solving the adjunct system eigenproblem:
 
−ωz,i 2 Mrc + Krc uz,i = 0 (8)

where ωz,i is the i th measured antiresonance frequency and uz,i ∈ R N−1


is the eigenvector, the so-called eigenvector of the
adjunct undamped system defined by Mr̄c̄ , Kr̄c̄ [32], that is related to the steady-state forced response when excited by a
harmonic force with frequency ωz,i , as shown in [33,34].
In this case, the interval MU problem is formulated starting from the eigenvalue problem of the adjunct system in
Eq. (8) and then recasting it as a least-square minimization, similar to Eq. (7). By assuming an arbitrary number nz of
measured antiresonances of receptance hrc (ω ), the following problem is formulated:
  

nz
    x
min αz,i  −ωz,i 2
Mr0c + Mrc (x ) + Kr0c + Krc (x ) uz,i 22 , ∈ z (9)
x, uz,i uz,i
i=1

where M (x ), K (x ) ∈ R(N−1 )×(N−1 ) are obtained by removing the c-th row and the r-th column from M, K. The
rc rc

positive scalars αz,i define differing levels of importance of each antiresonance frequency. Clearly, antiresonances of different
receptances can be included in Eq. (9), by suitably redefining the adjunct system.
The eigenvectors of the adjunct system are treated in the same way of the mode shape u p,i , and hence constrained to
belong to an interval defined through lower and upper bounds (Uz,L i and Uz, U respectively) that set the feasible domain  . In
i z
practice, the extraction of uz,i is highly uncertain and is not usually done. In this case, it suffices setting the constraint [−1,
+1] to all the entries of the eigenvectors, since eigenvectors are defined regardless of their normalization. In this way it is
possible performing MU through antiresonance frequencies by solving problem the eigenvalue problem cast as Eq. (9) that
is easier to solve than using the determinant, i.e. det(−ωz,i 2 (Mr0c + Mrc (x )) + Kr0c + Krc (x )) = 0 (see Eq. (1)). Indeed,
such an equation is a non-linear, non-convex, non-separable function with high-order products of more unknowns, whose
solution is cumbersome [32].

2.5. Formulation of model updating exploiting natural frequencies, antiresonances and their reciprocal location in the frequency
spectrum

The drawbacks of MU with natural frequencies and antiresonances alone can be overcome by the simultaneous use of
both the information, together with the interval definition of mode shapes through constraints. The simultaneous use of
Eqs.(7) and (9) leads to the following problem, where the feasible domain  collects all the constraints on the unknowns

5
D. Richiedei, I. Tamellin and A. Trevisani Mechanism and Machine Theory 160 (2021) 104288

Fig. 2. Sketch of the concept of interval representation of a mode shape in a sample cantilever beam, refined through the interlacing condition.

Fig. 3. Sketch of the concept of sign constraints due to the interlacing condition.

x
{u p,i }:
uz,i

n  2  rc   
αz,i  −ωz,i
i=1
z
M0 + Mrc (x ) + Kr0c + Krc (x ) uz,i 22 + x
min
n p   , u p,i ∈  (10)
x, u p,i ,uz,i
+ i=1 α p,i  −ω2p,i (M0 + M(x ) ) + K0 + K(x ) u p,i 22 uz,i

Besides than just merging Eqs.(7) and (9), an improved approach is proposed in this paper by exploiting the relation between
the reciprocal location of natural frequencies and antiresonances in the frequency spectrum and the mode shape. The same
relation is useful to MU without any estimation of mode shapes. A relation, that holds for low frequency modes, discloses
the pole-zero interlacing condition for receptance [33,35]
   
sign u p,i (r )u p,i (c ) = sign u p,i+1 (r )u p,i+1 (c ) ⇒ ω p,i < ωz,i < ω p,i+1 (11)

In practice, if the modal constants (i.e. the product of the two entries of the i th eigenvector at the response point u p,i (r )
and at the excitation point u p,i (c )) of two consecutive modes have the same sign, then there will be an antiresonance ωz,i
between the natural frequencies of the two modes considered. Such a condition can be adopted for low frequency modes
whose mode shape varies slowly in space.
The use of the condition in Eq. (11) in MU is twofold. On the one hand, it can be used to refine the bounds on the
eigenvectors by making them tighter, as sketched in Fig. 2 with reference to the cantilever beam proposed in Fig. 1.
On the other hand, the interlacing condition can be exploited in a different problem, that is MU through just the avail-
ability of some natural frequencies and antiresonance frequencies: the knowledge of the pattern of natural frequencies and
antiresonances in the frequency spectrum of an arbitrary hrc (ω ) enables to impose the modal constants of the mode shapes
as constraints on the sign of the mode shape, in lieu of their interval definition. It should be noted that the condition in
Eq. (11) is not convex when used in this way. Hence, a proper strategy to formulate such a constraint should be introduced
to boost the problem convergence towards a global optimal solution. An approach to convexify it, is proposed in Section 3.6.
Fig. 3.

6
D. Richiedei, I. Tamellin and A. Trevisani Mechanism and Machine Theory 160 (2021) 104288

3. Numerical solution of the updating of mass and stiffness matrices through homotopy optimization

3.1. Outline of the method

To overcome the issue of non-convexity of Eqs. (7), (9) and (10), updating of the mass and stiffness matrices is solved in
this paper by exploiting the theory of homotopy transformation and variables lifting. Homotopy is a well-established opti-
mization technique (see e.g. [36–39]), that boosts the attainment of the global optimum solution in non-convex optimization
problems by initially solving a convex approximation of the optimization and then morphing it back to the original problem
through a finite sequence of optimizations. Numerical solutions can be summarized through the following steps:

1. Normalization of the updating parameters (Section 3.2);


2. Variables lifting (Section 3.3);
3. Definition of the homotopy map (Section 3.4);
4. Introduction of McCormick’s constraints (Section 3.5);
5. Convexification of the pole-zero interlacing condition (Section 3.6).

3.2. Introduction of normalized modifications

To improve numerical conditioning of the proposed minimization problems, that include unknowns with significantly
different magnitudes, dimensionless normalized modifications m̄x and k̄x are introduced:
m x k x
m̄x = , k̄x = (12)
m0x k0x
where mx and kx denote updating parameters (i.e. the entry of x), whose original values are denoted by m0x and k0x . For
the parameters whose original value is zero, a reasonable arbitrary value within its lower and upper bound can be chosen.
The normalized updating parameters are collected in the vector of the normalized unknows x̄ = [m̄1 , ..., m̄nxm , k̄1 , ..., k̄nxk ] ∈
Rnx .
M and K are expressed through the constant Jacobian matrices:

nxm 
nxk
M = mx Jmx m0x , K = kx Jkx k0x (13)
x=1 x=1

where:
∂ M0 ∂ K0
Jmx = , Jkx = (14)
∂ m0x ∂ k0x
A fast and accurate way to compute the Jacobian matrices is exploiting the “complex step derivative” calculation, as sug-
gested in [11]:
     
Im K0 k0x + jε Im M0 m0x + jε
Jkx = Jmx = (15)
ε ε
where ε is a small scalar whose magnitude is close to the machine precision.
The MU problem expressed through normalized updating parameters, is the following one:

n z
  
nxm   
nxk rc  ⎫

⎪ α  −ω 2
M rc
+ m J m 0 rc
+ K rc
+ k J k 0
u  2
+ ⎪

⎨ i=1 z,i z,i 0 x=1 x m x x 0 x=1 x k x x

z ,i 2 ⎬
min
x, u p,i , uz,i ⎪

n p  
nxm 
nxk  x (16)
⎩+ i=1 α p,i  −ω p,i M0 + x=1 mx Jmx mx + K0 + x=1 kx Jkx kx up,i 2 ,
⎪ 2 0 0 2
up,i ∈ ⎪ ⎪

uz,i

where the feasible domain 


¯ is redefined to account for the normalized, dimensionless bounds.

3.3. Variables lifting

Non convexity in Eqs.(7), (9) and (10) arises due to the presence of bilinear terms xi u j (for some indexes i,j). Variables
lifting, i.e. the introduction of the auxiliary variables tw (with w an arbitrary index) defined as

tw = xi u j , (17)
allows writing Eqs.(7),(9) and (10) as convex functions of tw , i.e. writing a convex approximation (denoted fc ) of the non-
convex objective functions, named f . The introduction of new variables, together with the original ones, lifts the problem
to a higher dimension.

7
D. Richiedei, I. Tamellin and A. Trevisani Mechanism and Machine Theory 160 (2021) 104288

3.4. Homotopy maps

Homotopy map is introduced to morph back the convex approximation to the original non-convex ones (i.e. those in
Eqs.(7) and (10) depending on the problem solved). The homotopy map is here defined by replacing each bilinear term xi u j
with an affine combination of itself and its related variable tw :

λxi u j + (1 − λ )tw (18)


The scalar λ ∈ [0; 1] is the morphing parameter that increases discretely from 0 to 1, in a finite number nλ of iterations.
When λ = 0 the convex approximation fc of the original problem is solved while, when λ = 1 the original non-convex cost
problem f is handled. In the intermediate non-convex problems, i.e. when λ ∈ (0; 1 ), the solution of the previous step is set
as the initial guess for the following one, by defining a path of solutions that can converge to the global optimal solution,
whenever the convex approximation is correctly defined. Finally, the solution obtained at the last step leads to the optimal
updating mass and stiffness matrices, M(xλ=1 ) = Mopt and K(xλ=1 ) = Kopt . This approach have been shown to be
effective adopted in the recent years also for the inverse dynamic structural modification of vibrating systems [31–33].

3.5. McCormick’s constraints

McCormick’s constraints have been introduced in [40] to define the tightest convex approximation of a bilinear term.
Given two variables xi and u j , respectively belonging to the intervals [xLi , xUi ] and [U Lj , U Uj ], the McCormick’s constraints
approximate the variables lifting equality in Eq. (17) by means of four inequalities:

⎪tw ≥ xLi u j + U jL xi − xLi U jL



⎨tw ≥ xU u j + U U xi − xU U U
i j i j
(19)

⎪tw ≤ xUi u j + U jL xi − xUi U jL



tw ≤ xi u j + U j xi − xi U j
L U L U

3.6. Convexification of the pole-zero interlacing condition

The solution of the MU through natural and antiresonance frequencies in Problem 4 requires the convexification of the
interlacing condition in Eq. (11). The constraints on up,i imposed by Eq. (11) are recast as inequalities exploiting the fact that
eigenvectors are defined regardless of their normalization. Hence, the sign of the r-th and c-th entries of up,i are imposed
through negativity or positivity constraints of the form [UPL ; 0 ), with UPL < 0, or (0; UPU ]. Upper and lower bound are adopted,
rather than using half-plane constraints, to exploit them within the McCormick’s constraints.

4. Updating of the damping matrix

Once the mass and stiffness matrices are updated, damping matrix is updated separately. A more general formulation of
the classical damping, compared to the widespread one proposed by Rayleigh, is here adopted. Accordingly to the Caughey
model [28], the updated damping matrix C is formulated through the following series:
  h
C=M βh M−1 K (20)
h∈Y

where the real number h, either positive or negative, denotes the exponents of the series that belongs to a chosen, finite
nh -dimensional set Y = [hi , ..., h f ], while the coefficients βh are the series coefficients. Each coefficient β h is related to the
experimental modal damping ratio ξ p,i of the i th natural frequency ω p,i through the following equation:

1 1   2h
ξ p,i = βh ω p,i (21)
2 ω p,i
h∈Y

Collecting equations as Eq. (21) for all the measured modal damping ratio leads to a linear system of the following
form:

β = ξ (22)
where:
⎡  2hi  2h f ⎤
1
ω p,1 ··· 1
ω p,1 ⎧ ⎫ ⎧ ⎫
ω p,1 ω p,1 ⎨ β1 ⎬ ⎨ ξ p,1 ⎬
1⎢ ⎥
= ⎢ .. .. .. ⎥ ∈ Rn p ×nh , β = .. ∈ Rnh , ξ = .. ∈ Rn p (23)
2⎣ 
.
2hi
.

.
2h f ⎦ ⎩ . ⎭ ⎩ . ⎭
1
ω p,n p ··· 1
ω p,n p βnh ξ p,n p
ω p,n p ω p,n p

8
D. Richiedei, I. Tamellin and A. Trevisani Mechanism and Machine Theory 160 (2021) 104288

Fig. 4. Flow-chart of the model-updating procedure.

Whenever the number of identified modal damping ratios is greater than the cardinality of Y, least-square solution of the
linear system Eq. (22) should be computed. An effective way to represent uncertainty on some damping factors and to
improve the numerical conditioning of ,is weighing the least-square through a positive definite matrix W ∈ Rn p ×n p :

 −1
β = T W T Wξ (24)

Fig. 4.

5. Method assessment

5.1. General description and implementation details

The MU approach proposed in this paper has been implemented through MATLAB and the optimization toolbox Yalmip
[41]. The solution of the non-convex minimization can be performed through any solver for non-linear least-square prob-
lems; in this paper, the standard fmincon solver of MATLAB is employed. Two experimental test cases are proposed, as
summarized in Table 1. A flow-chart of the overall procedure is proposed in Fig. 5.

9
D. Richiedei, I. Tamellin and A. Trevisani Mechanism and Machine Theory 160 (2021) 104288

Table 1
Summary of the two test-cases.

Test case section 5.2 5.3

Problem statement Problem 1 Problem 4


Model of the system FEM + ERLS FEM
Number of DOFs 30 14
Number of system configurations 1 4
Number of measured natural frequencies 7 12
Number of measured antiresonances 0 4
Number of updated parameters 29 7
Number of unknown mode shape entries 210 168
Number of unknown entries of the adjunct system eigenvector 0 48
Size of t 868 596
Normalized/ non normalized formulation Normalized Non normalized

Fig. 5. Picture of the mechanism (left) and its finite-element model (right).

5.2. Test case 1 – planar flexible linkage manipulator

5.2.1. Model of the system


The first system consists of a planar, flexible-link multibody system, shown in Fig. 5. The mechanism is made by five
flexible links made of aluminum. Three motors are installed to drive links 1, 2 and 5 and two springs are installed to
balance gravity.
The motion of the mechanism is modeled as the large rigid-body motion of an equivalent rigid-link system (the ERLS)
plus the small elastic displacements, that are defined with respect to the ERLS. A set of nonlinear ordinary differential
equations is obtained [2,11,42]. Let us define θ as the vector of the ERLS generalized coordinates, and u the vector of the
elastic displacements of the nodes of the FE model (with respect to the ERLS), defined in a global reference.
A linearized model for MU is obtained by the first-order term of the Taylor’s expansion computed about the asymptotic
equilibrium configuration of the ERLS, θ e , adopted for the experimental modal analysis. Linearization about an asymptotic
equilibrium configuration allows applying the proposed method to update model of nonlinear multibody systems, whenever
nonlinearities are related to the system position.
The resulting model recalls the one usually adopted in structural dynamics with locally constant matrices:
             
Me Me S ü(t ) C 0 u˙ (t ) Ke 0 u(t ) I
+ e + = f (t ) (25)
ST Me ST Me S θ̈ (t ) 0 0 θ˙ (t ) 0 Kg θ (t ) ST
θ ≡θe θ ≡ θe θ ≡ θe θ ≡ θe

where Me , Ce and Ke are the mass, damping and stiffness matrices of all the finite elements and are computed about the
equilibrium configuration; Kg represents the linearized effects of gravity. S = S(θ ) is the matrix of the sensitivity coefficients
of all ERLS nodes with respect to the ERLS generalized coordinates, which relates the velocities of the ERLS generalized
coordinates to the ones of all the nodes of the ERLS. Finally, f collects the external nodal forces. Euler-Bernoulli beam
elements (each one with two nodes and six DOFs) are exploited to model the links. Nine elements are used to represent the
planar motion of the mechanism, as shown in Fig. 5, leading to 27 elastic DOFs collected in vector u and 3 ERLS-generalized
coordinates collected in vector θ =[ϑ1 ,ϑ2 ,ϑ3 ]T , that represent the joint absolute rotations of the motors.

10
D. Richiedei, I. Tamellin and A. Trevisani Mechanism and Machine Theory 160 (2021) 104288

Fig. 6. Experimental FRFs for different excitation points.

5.2.2. Results
The experimental data adopted are those obtained by the experimental modal analysis performed in [11], here adopted
as benchmark. The frequency response functions of the mechanism for different excitation point are shown in Fig. 6.
Natural frequencies, damping and mode shapes have been identified through impact hammer excitations and accelerom-
eter measurements. Post-processing has been done through the PolyMAX of LMS Test.Lab 11B tool to extract the mode shapes
for the translational coordinates with the “pole-residual model” [13]. Since u includes the beam elastic rotations, least-square
fitting has been employed to expand the translational measurements for estimating the unmeasured rotational coordinates,
by taking advantage of redundant measurements of the transverse displacements along the beams and of polynomial inter-
polation functions. It follows that mode shapes are affected by uncertainty.
In accordance with the test proposed in [11], both the experimental modal analysis and the model refer to the equilib-
rium configuration θ e =[1.082, 2.635, 4.712]T rad. The bandwidth of interest spans from 0 to 200 Hz. Hence, the first seven
natural frequencies ω p,1 , ..., ω p,7 are considered, while mode shapes are assumed as uncertain values about those identified.
It should be noted that natural frequencies and mode shape vary as the equilibrium configuration changes [2,11,43]. No an-
tiresonances are adopted and all the coefficients α p,i have been set to 1, to make a proper comparison with [11]. Normalized
modifications are adopted to improve numerical conditioning.
The system original parameters are reported in Table 2, with ρ denoting the link linear mass density, E the Young’s
modulus. The lumped stiffness of the balancing springs is kbs while the nodal inertias and masses are respectively J and M
(with proper subscripts to denote their position along the mechanism).
The natural frequencies obtained are compared in Table 3 with those of the nominal model and those obtained in [11].
Correctness is evaluated through the percentage frequency error εi :
 u 
 ω p,i − ω p,i 

εi = 100  (26)
ω p,i 
where ω p,i and ωup,i are the experimental and the updated natural frequencies.

11
D. Richiedei, I. Tamellin and A. Trevisani Mechanism and Machine Theory 160 (2021) 104288

Table 2
Test-case 1: updating parameters.

Parameter Nominal value Lower bound Upper bound Updated value

ρ [kgm ]−3
2710.0 2574.5 2845.5 2777.9
E1 [GPa] 69.00 65.55 72.45 65.55
E2 [GPa] 69.00 65.55 72.45 70.47
E3 [GPa] 69.00 65.55 72.45 72.45
E4 [GPa] 69.00 65.55 72.45 65.55
E5 [GPa] 69.00 65.55 72.45 65.55
kbs [Nm−1 ] 1520.0 1216.0 1824.0 1219.3
MA,1 [kg] 7.644 6.115 9.172 8.234
MB,1 [kg] 0.392 0.313 0.470 0.314
MA,2 [kg] 9.517 7.613 11.420 11.42
MH,2 [kg] 0.400 0.320 0.480 0.400
MD,2 [kg] 0.648 0.518 0.777 0.518
MB,3 [kg] 0.671 0.536 0.805 0.766
MC,3 [kg] 0.383 0.306 0.459 0.460
MC,4 [kg] 0.657 0.526 0.789 0.526
MD,4 [kg] 0.308 0.246 0.369 0.285
ME,4 [kg] 1.844 1.537 2.151 1.751
ME,5 [kg] 0.114 0.095 0.133 0.160
MF,5 [kg] 0.046 0.036 0.055 0.046
JA,1 [kgm2 ] 0.0130 0.0104 0.0156 0.0144
JA,2 [kgm2 ] 0.0229 0.0183 0.0275 0.0275
JE,5 [kgm2 ] 0.0015 0.0012 0.0018 0.0014
JB,1 [kgm2 ] 0 0 0.00050 0.00044
JB,3 [kgm2 ] 0 0 0.0005 0.00044
JC,3 [kgm2 ] 0 0 0.00050 0.00019
JC,4 [kgm2 ] 0 0 0.00050 0.00044
JD,4 [kgm2 ] 0 0 0.00050 0.00044
JD,2 [kgm2 ] 0 0 0.00050 0
JE,4 [kgm2 ] 0 0 0.00120 0.00106

Table 3
Comparison of the experimental and analytical natural frequencies.

Natural frequency Natural frequency εi (original Natural frequency εi (updated model) Natural frequency εi (updated model)
(experimental) [Hz] (original model) model) [%] (updated model) in in [11] [%] (updated model) proposed method
[Hz] [11] [Hz] proposed method [Hz] [%]

ωp,1 13.39 13.72 2.42 13.27 0.82 13.39 0.00


ωp,2 43.58 43.88 0.69 43.08 1.12 43.58 0.00
ωp,3 64.65 66.95 3.56 64.83 0.26 64.65 0.00
ωp,4 112.98 127.25 12.62 120.77 6.89 119.98 6.19
ωp,5 138.15 147.22 6.57 142.30 2.99 138.15 0.00
ωp,6 154.73 171.95 11.13 156.25 0.97 161.33 4.26
ωp,7 198.56 204.92 3.20 198.63 0.03 198.56 0.00

Table 4
Comparison of the experimental and analytical damping ratios.

Experimental modal Modal damping ηi (updated model) Modal damping with the ηi (updated model)
damping [-] estimated in [11] [-] in [11] [%] proposed method [-] proposed method [%]

ξ p,1 0.007 0.007 0.0 0.007 0.2


ξ p,2 0.012 0.013 8.3 0.011 4.8
ξ p,3 0.017 0.018 5.9 0.018 8.2
ξ p,4 0.036 0.035 2.8 0.036 0.9
ξ p,5 0.043 0.040 7.0 0.040 6.1
ξ p,6 0.039 0.044 12.8 0.044 13.2
ξ p,7 0.048 0.055 14.6 0.047 2.0

The method proposed in this work ensures negligible frequency errors, as corroborated by Table 3: the average
εi decreases from 5.7% of the original nominal model to 1.5% of the updated one. The error obtained in [11] was bigger,
i.e. 1.9%. It should be noted that five eigenfrequencies are, in practice, exactly obtained through the new method, and just
the 6th mode provides a worst estimation than the one in [11].
Damping matrix has been estimated through Eq. (20) by setting Y=[0, 0.25, 0.5, 0.75, 1]T and without weighing the
least-squares. The resulting damping ratios are listed in Table 4 and compared with those obtained through the traditional
Rayleigh damping model as proposed in [11].

12
D. Richiedei, I. Tamellin and A. Trevisani Mechanism and Machine Theory 160 (2021) 104288

Fig. 7. The experimental testbed (test-case 2).

Fig. 8. Detail of the beam, with the piezoelectric patch actuator (test-case 2).

The relative percentage error ηi for the i th updated damping ratio is computed as:
 u 
 ξ − ξ p,i 
ηi = 100 p,i  (27)
ξ p,i 
where ξ p,i and ξ p,i
u are the experimental and the updated damping ratios, respectively. The average η provided by the
i
proposed method is 5.1% while in [11] it is 7.3%.

5.3. Test case 2 – cantilever beam

5.3.1. System model


The system investigated in the second test case is a flexible cantilever beam attached to the slip table of an electrody-
namic shaker (Tira GmbH TV50350LS), through a stinger (that is rigid in the frequency range of interest), and a load cell
(see Fig. 7). A detail of the beam alone is shown in Fig. 8, to highlight the piezoelectric patch (that is not evident in Fig. 7).
The cantilever beam (whose length L = 0.493 m, its width is 40 mm and its height is 4 mm) is modelled through
seven Euler-Bernoulli finite elements. A piezoelectric patch actuator PI P-876.A12 (whose length is 61 mm, its width 35 mm
and height equals to 0.5 mm), here just considered through its contributions on the mass and stiffness matrices (i.e., no
voltage is applied), is glued on the beam in correspondence of the finite element whose coordinates are y5 , ϕ5 , y6 , ϕ6 . An
accelerometer (Endevco 27AM1–100 10,203), whose mass is macc , is placed at the tip-end of the beam. The translational and
rotational coordinates of the beam are collected in vector q = [y1 , ϕ1 , ..., y7 , ϕ7 ] ∈ R14 .
T

The stinger is attached at the free-end of a load cell modeled as a 1-DOF mass-spring system as often done in the
literature [33,44]: mlc and klc denote the mass and stiffness of the load cell.
Since the mass of the moving parts of the shaker, msh is such that msh
ρ L + mlc + ρ p L p + macc and the shaker accel-
eration ÿs (t ), measured through a Dytran 3136A accelerometer, is controlled by a feedback driving units, the motion of the
table is unaffected by the elastic dynamics of the cantilever beam and of the load cell. Therefore ys (t ) can be assumed as
the input for the system made by the load cell and the beam. Hence, the following 14-DOFs undamped elastic model of the

13
D. Richiedei, I. Tamellin and A. Trevisani Mechanism and Machine Theory 160 (2021) 104288

Fig. 9. Test-case 2: model of the cantilever-beam on the slip table with four different configurations.

Table 5
Comparison of the experimental and analytical natural and antiresonance frequencies of FRF hy1 ,ys (jω ) in the four configurations.

Frequency Frequency (original εi (original Frequency (updated εi (updated


(experimental) [Hz] model) [Hz] model) [%] model) [Hz] model) [%]

ωAp,1 73.4 66.1 9.9 73.4 0.0


ωAp,2 127.3 125.8 1.2 128.7 1.2
ωAp,3 229.5 215.4 6.2 228.4 0.5
ωz,A 3 101.3 94.8 6.4 101.2 0.1
ωBp,1 50.5 45.6 9.7 50.5 0.0
ωBp,2 121.8 121.2 0.5 124.8 2.5
ωBp,3 218.8 205.4 6.1 214.8 1.8
ωz,B 3 101.2 94.8 6.3 101.2 0.0
ωCp,1 72.4 66.1 8.7 73.3 1.3
ωCp,2 125 123.7 1.0 125.8 0.6
ωCp,3 216.1 203.9 5.7 216.1 0.0
ωCz,3 101.3 94.8 6.4 101.2 0.1
ωDp,1 64.9 59.0 9.1 65.2 0.5
ωDp,2 123.9 123.1 0.7 125.8 1.5
ωDp,3 227.5 215.2 5.5 228.3 0.4
ωz,D 3 106.6 100.2 6.0 106.7 0.1

system is obtained, as sketched in Fig. 9:


   
MB + MP + e3 mlc eT3 q̈(t ) + KB + KP + e3 klc eT3 q(t ) = klc ys (t )e3 (28)

MP , KP ∈ R14×14 are the piezoelectric patch mass and stiffness matrices while MB , KB ∈ R14×14 are the mass and stiffness
matrices of the cantilever beam. e3 ∈ R14 is the vector whose entries are null except for the one related to y3 (t ), which is
one.
The measured receptance adopted to apply the proposed method is the one from the force applied to the translational
coordinate y3 (t ) to the tip-end of the beam, i.e. hy1 , f3 (jω ). Since in the experimental analysis hy1 ,ys (jω ) it is easier to
consider, it is measured the FRF hy1 ,ys (jω ) = klc hy1 , f3 (jω ), that features the same natural and antiresonance frequencies as
hy1 , f 3 ( jω ).

5.3.2. Results
Four experimental receptances hy1 ,ys (jω ) have been measured by moving a known proof mass (mt = 26.2e-3 kg) along
the beam, leading to four system configurations denoted by letters A, B, C, D. Configuration A is the system without the
additional proof mass, while the other ones are sketched in Fig. 9. Both the acceleration of the tip-end of the beam and of
the slip-table are measured through accelerometers and processed through a LMS SCADAS Mobile.
The experimental natural and antiresonance frequencies are reported in Table 5. The interlacing between natural and
antiresonance imposes that the following relation between the mode shapes entries must hold:
   
sign uup,1 (y1 )uup,1 (y3 ) = sign uup,2 (y1 )uup,2 (y3 )
    (29)
sign uup,2 (y1 )uup,2 (y3 ) = sign uup,3 (y1 )uup,3 (y3 )

14
D. Richiedei, I. Tamellin and A. Trevisani Mechanism and Machine Theory 160 (2021) 104288

Fig. 10. Sketch of the bounds on the mode shapes to ensure the pole-zero interlacing pattern in test-case 2.

Table 6
Test-case 2: updating parameters.

Parameter Nominal value Lower bound Upper bound Updated value

Beam Young’s Modulus: E [GPa] 55.0 50.3 73.8 68.1


Beam mass density: ρ [kgm−3 ] 2400 1775 3025 2383.2
Accelerometer mass: macc [kg] 0.002 0.0015 0.003 0.0015
Load cell mass: mlc [kg] 0.1097 0.0853 0.4997 0.0946
Load cell stiffness: klc [Nm−1 ] 2.084e5 1.084e5 3.084e5 1.871e5
Piezo patch Young’s Modulus: Ep [GPa] 34.7 14.6 101.7 15.3
Piezo patch mass density: ρ p [kgm−3 ] 4215.5 644.0 7786.9 7786.3

Table 7
Comparison of the experimental and analytical damping ratios.

Experimental modal Modal damping ηi (updated Modal damping with ηi (updated model)
damping [-] estimated with model) with the proposed method proposed method [%]
Rayleigh model [-] Rayleigh [%] [-]

ξ p,A 1 0.044 0.004 90.8 0.022 49.5


ξ p,A 2 0.013 0.002 82.2 0.013 0.7
ξ p,A 3 0.006 0.001 79.2 0.007 18.1
ξ p,B 1 0.019 0.003 85.5 0.015 20.8
ξ p,B 2 0.015 0.002 84.3 0.013 11.2
ξ p,B 3 0.008 0.001 83.9 0.008 8.7
ξ p,C 1 0.035 0.004 88.6 0.022 37.7
ξ p,C 2 0.012 0.002 81.9 0.013 2.6
ξ p,C 3 0.007 0.001 81.9 0.007 2.6
ξ p,D 1 0.032 0.004 88.9 0.020 39
ξ p,D 2 0.014 0.002 83.9 0.013 8.7
ξ p,D 3 0.007 0.001 82.4 0.007 0.1

Eq. (29) is translated into MU through constraints on u p,i as suggested in Sections 2.5 and 3.6. Fig. 10 sketches the negativity
and positivity bounds imposed for uup,i (y1 ) and uup,i (y3 )in such a way that the prescribed pole-zero interlacing pattern is
obtained.
All the system parameters, except the proof mass, are unknown variables in MU leading to x=[E, ρ , macc , mlc , klc , Ep ,
ρ p ]T . The nominal values and the reasonable constraints assumed are reported in Table 6. The values of the parameters in
the original (nominal) model, have been taken from manufacturers data or as initial guess, in the case of unknown values.
Since neither the mode shape u p,i nor the eigenvector of the adjunct systems uz,i have been measured, the constraints
have been arbitrarily set as [−1;+1], except for the entries of u p,i involved in the interlacing conditions.
By applying the proposed method, a notable improvement in all the natural and antiresonance frequencies is obtained:
the average percentage error decreases from 5.6% in the nominal model to 0.6% in the updated model. The optimal system
parameters are reported in Table 6.
Damping matrix is identified by setting Y = [−1.5, −0.75, 0]T and by exploiting a weighted least-square that trusts more
the modal damping ratios of the first natural frequency, to enforce the model consistence in the low frequency range. The
experimental modal damping ratios have been obtained through the half-power bandwidth method. The Rayleigh model,
use as a benchmark, approximates the damping ratios with an average error of 84.5%; the improved method here proposed
drastically reduces the error whose average value is 16.6%, as shown in Table 7. The nominal, experimental and updated
FRFs hy1 ,ys (jω ) are shown in Fig. 11 in the sample case of configuration D.

Conclusions

In this paper a novel indirect method for interval model updating is proposed. The definition of intervals where the mode
shapes should belong accounts for the unavoidable uncertainty on such parameters. To cope with difficulties in formulating
stochastic models or in solving interval analysis problem, an effective solution based on homotopy transformation and norm
minimization is proposed. The method relies on a two-stage procedure that handles both the updating of the mass and

15
D. Richiedei, I. Tamellin and A. Trevisani Mechanism and Machine Theory 160 (2021) 104288

Fig. 11. Nominal, experimental and updated FRFs hy1 ,ys (jω ) for configuration D.

stiffness parameters and then the identification of the damping matrix. In the first stage, updating of mass and stiffness
matrices is performed through different possible scenarios that uses natural frequencies, antiresonances, mode shapes, that
can be also merged due to the flexibility of the proposed approach. Uncertainty on the mode shape is included through
bounds and such eigenvectors are treated as constrained unknowns in the optimization problem representing the inverse
eigenvalue problem. In the second stage, the identification of the damping matrix is performed solving a linear system that
arises from the Caughey’s proportional damping model. Uncertainty is handled through weighted projections.
To assess for the method effectiveness two experimental test-cases are carried out. In the first test case the system ma-
trices of the linearized model of a flexible multibody system are updated through the measured natural frequencies and the
uncertainty bounds on the mode shape. A normalized formulation is adopted too, to tackle the ill-numerical conditioning.
The results corroborate the effectiveness of both the stages of the proposed method leading to an improvement with re-
spect to the results proposed in the literature. The second test-case is a system made by a flexible cantilever-beam plus a
spring-mass system. Model updating is performed through the natural frequencies, the antiresonance frequencies of a re-
ceptance and the condition on the mode shapes dictated by the pole-zero interlacing pattern. A very precise estimation of
the measured properties is, again, provided by the proposed method.
To summarize, the following benefits are provided by the proposed method, as proved by the challenging test cases:
• it can handle simultaneously natural frequencies, mode shapes, antiresonance frequencies and exploits the condition on
the pole-zero interlacing;
• the method is versatile and can “mix” the experimental information to solve different problems;
• uncertainty on the mode shapes is accounted for through bounds;
• the method can cope with incomplete information on some entries of the mode shapes;
• the mode shape can have arbitrary normalization;
• the solution based on homotopy transformation is effective and simple and can include several kinds of constraints to
assure physical meaning of the updating parameters;
• complex step derivative is suggested to simplify the numerical implementation in a normalized formulation that handles
parameters with different magnitudes.
• mass and stiffness matrices are simultaneously updated;
• the separated updating of damping, through a more general formulation of proportional damping, simplifies the updating
of mass and stiffness matrices.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have
appeared to influence the work reported in this paper.

16
D. Richiedei, I. Tamellin and A. Trevisani Mechanism and Machine Theory 160 (2021) 104288

Acknowledgment

The second author acknowledges the financial support of the Cariparo foundation (“Fondazione Cassa di Risparmio di
Padova e Rovigo”) through a Ph.D. scholarship.

References

[1] R. Belotti, D. Richiedei, I. Tamellin, Pole assignment for active vibration control of linear vibrating systems through linear matrix inequalities, Appl. Sci.
10 (2020) 5494 https://doi.org/10.3390/app10165494.
[2] R. Caracciolo, D. Richiedei, A. Trevisani, Robust piecewise-linear state observers for flexible link mechanisms, J. Dyn. Syst. Meas. Control. Trans. ASME.
(2008), doi:10.1115/1.2909600.
[3] P. Boscariol, A. Gasparetto, Optimal trajectory planning for nonlinear systems: robust and constrained solution, Robotica (2016), doi:10.1017/
S0263574714002239.
[4] R. Caracciolo, D. Richiedei, A. Trevisani, G. Zanardo, Designing vibratory linear feeders through an inverse dynamic structural modification approach,
Int. J. Adv. Manuf. Technol. (2015), doi:10.10 07/s0 0170-015- 7096- 0.
[5] D. Richiedei, A. Trevisani, Simultaneous active and passive control for eigenstructure assignment in lightly damped systems, Mech. Syst. Signal Process.
85 (2017) 556–566, doi:10.1016/j.ymssp.2016.08.046.
[6] R. Belotti, D. Richiedei, A. Trevisani, Multi-domain optimization of the eigenstructure of controlled underactuated vibrating systems, Struct. Multidiscip.
Optim. (2020), doi:10.10 07/s0 0158- 020- 02709- x.
[7] J.E. Mottershead, M.I. Friswell, Model updating in structural dynamics: a survey, J. Sound Vib. (1993), doi:10.1006/jsvi.1993.1340.
[8] S. Sehgal, H. Kumar, Structural dynamic model updating techniques: a state of the art review, Arch. Comput. Methods Eng. (2016), doi:10.1007/
s11831- 015- 9150- 3.
[9] C.M. Pappalardo, D. Guida, A time-domain system identification numerical procedure for obtaining linear dynamical models of multibody mechanical
systems, Arch. Appl. Mech. (2018), doi:10.10 07/s0 0419- 018- 1374- x.
[10] J.E. Mottershead, M. Link, M.I. Friswell, The sensitivity method in finite element model updating: a tutorial, Mech. Syst. Signal Process. (2011), doi:10.
1016/j.ymssp.2010.10.012.
[11] R. Belotti, R. Caracciolo, I. Palomba, D. Richiedei, A. Trevisani, An updating method for finite element models of flexible-link mechanisms based on an
equivalent rigid-link system, Shock Vib (2018), doi:10.1155/2018/1797506.
[12] W. Fan, P. Qiao, Vibration-based damage identification methods: a review and comparative study, Struct. Heal. Monit. (2011), doi:10.1177/
1475921710365419.
[13] B. Peeters, G. Lowet, H. Van der Auweraer, J. Leuridan, A new procedure for modal parameter estimation, Sound Vib. (2004).
[14] S.R. Ibrahim, Computation of normal modes from identified complex modes, AIAA J (1983), doi:10.2514/3.60118.
[15] D. Richiedei, I. Tamellin, Active approaches to vibration absorption through antiresonance assignment: a comparative study, Appl. Sci. 11 (2021) 1091,
doi:10.3390/app11031091.
[16] W. D’Ambrogio, A. Fregolent, The use of antiresonances for robust model updating, J. Sound Vib. (20 0 0), doi:10.1006/jsvi.1999.2987.
[17] V. Meruane, Model updating using antiresonant frequencies identified from transmissibility functions, J. Sound Vib. 332 (2013) 807–820, doi:10.1016/
j.jsv.2012.10.021.
[18] W. D’Ambrogio, A. Fregolent, Results obtained by minimising natural frequency and antir esonance errors of a beam model, Mech. Syst. Signal Process.
17 (2003) 29–37, doi:10.1006/mssp.2002.1536.
[19] D. Hanson, T.P. Waters, D.J. Thompson, R.B. Randall, R.A.J. Ford, The role of anti-resonance frequencies from operational modal analysis in finite element
model updating, Mech. Syst. Signal Process. (2007), doi:10.1016/j.ymssp.2006.01.001.
[20] C. Mares, J.E. Mottershead, M.I. Friswell, Stochastic model updating: part 1-theory and simulated example, Mech. Syst. Signal Process. (2006), doi:10.
1016/j.ymssp.20 05.06.0 06.
[21] R. Arnaud, F. Poirion, Stochastic annealing optimization of uncertain aeroelastic system, Aerosp. Sci. Technol. (2014), doi:10.1016/j.ast.2014.06.008.
[22] I. Boulkaibet, L. Mthembu, T. Marwala, M.I. Friswell, S. Adhikari, Finite element model updating using Hamiltonian Monte Carlo techniques, Inverse
Probl. Sci. Eng. (2017), doi:10.1080/17415977.2016.1215446.
[23] H.H. Khodaparast, J.E. Mottershead, K.J. Badcock, Interval model updating with irreducible uncertainty using the Kriging predictor, Mech. Syst. Signal
Process. (2011), doi:10.1016/j.ymssp.2010.10.009.
[24] S. Gabriele, C. Valente, An interval-based technique for FE model updating, Int. J. Reliab. Saf. (2009), doi:10.1504/IJRS.2009.026836.
[25] M.Q. Zhang, M. Beer, C.G. Koh, Interval analysis for system identification of linear mdof structures in the presence of modeling errors, J. Eng. Mech.
(2012), doi:10.1061/(ASCE)EM.1943-7889.0 0 0 0433.
[26] D.K. Miu, Physical interpretation of transfer function zeros for simple control systems with mechanical flexibilities, J. Dyn. Syst. Meas. Control. Trans.
ASME. 113 (1991) 419–424, doi:10.1115/1.2896426.
[27] M.La Civita, A. Sestieri, On antiresonance interpretation and energy concentration along continuous one-dimensional systems, Proceeding of the Inter-
national Modal Analysis Conference - IMAC, 1997.
[28] T.K. Caughey, M.E.J. O’kelly, Classical normal modes in damped linear dynamic systems, J. Appl. Mech. Trans. ASME. (1964), doi:10.1115/1.3627262.
[29] J.M. Montalvao, E. Silva, An overview of the fundamentals of modal analysis, Proceeding of the NATO Advanced Study Institute Modal Analysis Test,
Kruwer Academic Publishers, 1999, doi:10.1007/978- 94- 011- 4503- 9.
[30] Z. Xu Yuan, K.Ping Yu, Finite element model updating of damped structures using vibration test data under base excitation, J. Sound Vib. (2015),
doi:10.1016/j.jsv.2014.11.041.
[31] R. Belotti, D. Richiedei, A. Trevisani, Optimal design of vibrating systems through partial eigenstructure assignment, J. Mech. Des. Trans. ASME. 138
(2016) 071402, doi:10.1115/1.4033505.
[32] R. Belotti, D. Richiedei, I. Tamellin, Antiresonance assignment in point and cross receptances for undamped vibrating systems, J. Mech. Des. 142 (2019)
022301, doi:10.1115/1.4044329.
[33] D. Richiedei, I. Tamellin, A. Trevisani, Simultaneous assignment of resonances and antiresonances in vibrating systems through inverse dynamic struc-
tural modification, J. Sound Vib. 485 (2020) 115552, doi:10.1016/j.jsv.2020.115552.
[34] B.P. Wang, Antiresonance and its sensitivity analysis in structural systems, Collect. Tech. Pap. - AIAA/ASME/ASCE/AHS/ASC Struct. Struct. Dynamic
Materials Conference, 1998, doi:10.2514/6.1998-1751.
[35] A. Preumont, Vibration control of active structures, Solid Mech. Its Appl. 179 (2011) 1–452, doi:10.1007/978- 94- 007- 2033- 6_1.
[36] L.T. Watson, R.T. Haftka, Modern homotopy methods in optimization, Comput. Methods Appl. Mech. Eng. (1989), doi:10.1016/0 045-7825(89)90 053-4.
[37] C.P. Vyasarayani, T. Uchida, A. Carvalho, J. McPhee, Parameter identification in dynamic systems using the homotopy optimization approach, Multibody
Syst. Dyn. (2011), doi:10.1007/s11044- 011- 9260- 0.
[38] A. Hall, T. Uchida, F. Loh, C. Schmitke, J. Mcphee, Reduction of a vehicle multibody dynamic model using homotopy optimization, Arch. Mech. Eng.
(2013), doi:10.2478/meceng-2013-0 0 02.
[39] B. Ghannadi, R. Sharif Razavian, J. McPhee, A modified homotopy optimization for parameter identification in dynamic systems with backlash discon-
tinuity, Nonlinear Dyn. (2019), doi:10.1007/s11071- 018- 4550- 1.
[40] G.P. McCormick, Computability of global solutions to factorable nonconvex programs: part I - Convex underestimating problems, Math. Program. (1976).

17
D. Richiedei, I. Tamellin and A. Trevisani Mechanism and Machine Theory 160 (2021) 104288

[41] J. Löfberg, YALMIP: a toolbox for modeling and optimization in MATLAB, Proceeding of the IEEE International Symposium Computer Control System
Design, 2004, doi:10.1109/cacsd.2004.1393890.
[42] R. Vidoni, A. Gasparetto, M. Giovagnoni, Design and implementation of an ERLS-based 3-D dynamic formulation for flexible-link robots, Robot. Comput.
Integr. Manuf. (2013), doi:10.1016/j.rcim.2012.07.008.
[43] I. Palomba, R. Vidoni, Flexible-link multibody system eigenvalue analysis parameterized with respect to rigid-body motion, Appl. Sci. (2019), doi:10.
3390/app9235156.
[44] D. Richiedei, A. Trevisani, Shaper-based filters for the compensation of the load cell response in dynamic mass measurement, Mech. Syst. Signal
Process. (2018), doi:10.1016/j.ymssp.2017.04.049.

18

You might also like