Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Nonlinear Dyn (2023) 111:1971–1986

https://doi.org/10.1007/s11071-022-07969-4

ORIGINAL PAPER

Wave-based analysis of jointed elastic bars: stability of


nonlinear solutions
Nidish Narayanaa Balaji ·
Matthew R. W. Brake · Michael J. Leamy

Received: 3 March 2022 / Accepted: 30 September 2022 / Published online: 21 October 2022
© The Author(s), under exclusive licence to Springer Nature B.V. 2022

Abstract In this paper we develop two new approa- Keywords Wave-based vibration approach · Nonlin-
ches for directly assessing stability of nonlinear wave- ear wave propagation · Perturbation methods · Floquet
based solutions, with application to jointed elastic bars. theory
In the first stability approach, we strain a stiffness
parameter and construct analytical stability bound-
aries using a wave-based method. Not only does this 1 Introduction
accurately determine stability of the periodic solutions
found in the example case of two bars connected by In a companion paper [1], we developed two wave-
a nonlinear joint, but it directly governs the response based methods for predicting the nonlinear vibration
and stability of parametrically forced continuous sys- response of jointed, continuous elastic bars. These
tems without resorting to discretization, a new devel- methods consist of a nonlinear wave-based vibration
opment in of itself. In the second stability approach, approach (WBVA) informed by re-usable, perturbation-
we pose a perturbation eigenproblem residue (PER) derived scattering functions, and a numerical plane
and show that changes in the sign of the PER locate wave expansion (PWE) approach exploiting wave solu-
critical points where stability changes from stable to tions as expansion quantities. Both approaches were
unstable, and vice-versa. Lastly, we discuss follow-on applied to finding periodic solutions in example sys-
research using the developed stability approaches. In tems in which continuous bars, connected by nonlin-
particular, we identify an opportunity to study stabil- ear joints, are constrained at their ends and harmoni-
ity around internal resonance, and then identify a need cally excited. As documented in [1], the two approaches
to further develop and interpret the PER approach to exhibit very good accuracy while significantly reduc-
directly predict stability. ing the computation time required to obtain periodic
solutions. They are also notable for having fixed prob-
Supplementary Information The online version contains
supplementary material available at https://doi.org/10.1007/
lem size, independent of frequency. Arguably, stability
s11071-022-07969-4. of the periodic solutions predicted is as important as the
solutions themselves. In [1] we report stability results
N. N. Balaji · M. R. W. Brake for the examples studied using two new stability meth-
Department of Mechanical Engineering, Rice University,
Houston, TX, USA
ods, which we deferred discussion of until the present
paper.
M. J. Leamy (B)
School of Mechanical Engineering, Georgia Institute of
Direct methods for assessing stability of wave-based
Technology, Atlanta, GA, USA solutions have not been presented in the literature, the
e-mail: michael.leamy@me.gatech.edu necessity of which has been stated in recent studies (see,

123
1972 N. N. Balaji et al.

for instance [2,3]). When stability has been treated, 2 Stability via the strained parameter approach
previous studies, owing to a lack of alternatives, have
chosen to project the wave-based solution onto a finite We now formulate an analytical approach to determine
element mesh and then use time-domain monodromy stability of the multiple periodic solutions found in
matrix calculations to estimate stability. Although this [1] for the single-jointed system shown in Fig. 1. The
approach is very attractive owing to its simplicity, it approach shares many similarities with the “strained
leads to inaccuracies at larger amplitudes, which we parameter”, or Lindstedt-Poincaré, approach employed
document in Appendix A. In this paper, we assess sta- in the study of the damped Mathieu equation [4]. While
bility of the multi-harmonic periodic solutions found the Mathieu equation is a single ordinary differential
in [1] using two new approaches: a strained parameter, equation, the strained parameter approach developed
analytical method applied to the governing partial dif- herein assesses stability of continuous systems gov-
ferential equations based loosely on the technique as erned by partial differential equations. In addition, the
applied to parametrically excited ordinary differential motivating example studied consists of multiple con-
equations [4], and a computational method employ- tinuous domains connected by a nonlinear joint.
ing the residue of the perturbed system’s eigenvalue The equations governing vibration response in the
problem. Both are informed by Floquet theory. Flo- linearly elastic bars are given by,
quet theory (see [4]) provides the theoretical basis for
the fact that stability transitions of a system perturbed ∂ 2u L ∂ 2u L
ρA − Ey A = f L (X, t), 0 < X < (α1 + α2 )l
around a periodic solution occur when the perturbed ∂t 2 ∂ X2
(1)
system has a purely imaginary eigenvalue with mag-
nitude equaling an integer multiple of the frequency ∂ 2u R ∂ 2u R
ρA − Ey A = f R (X, t), (α1 + α2 ) < X < l
of the response (among other cases, see Sect. 3 below). ∂t 2 ∂ X2
(2)
Detection of such points along a forced response curve,
as opposed to a complete spectral decomposition of
the perturbed system for each response on the curve where the left and right bars have displacement fields
(such as the HBM perturbation approach in [5]), under- denoted by u L (X, t) and u R (X, t), while f L (X, t)
pins both proposed approaches. In the strained param- and f R (X, t) denote forcing per unit length in the left
eter approach we find stability tongues in the parame- and right subdomains, respectively. For simplicity of
ter space distinguishing stable and unstable solutions, resulting expressions, the two bars considered share
while in the perturbed eigenproblem residue approach the same Young’s modulus E y , density per unit vol-
we employ Floquet theory to find changes in the com- ume ρ, and cross-sectional area A. The three values for
puted residues which indicate a stability transition. αi , i = 1, 2, 3, form a partition of unity. Four boundary
The paper is organized as follows: we first develop conditions are required to completely specify the vibra-
an analytical, wave-based stability approach based on tion response of two jointed bars. The two clamped
straining a single stiffness parameter in Sect. 2. This ends provide two boundary conditions, u L (0, t) = 0
strained parameter approach yields stability curves and u R (l, t) = 0, while two others arise at the joint,
(i.e., Arnold tongues) and surfaces associated with 
∂u L
a parametrically excited continuous system. Next, in Ey A = K (u R −u L ) + ε (u R −u L )3
Sect. 3 we develop a computational approach termed ∂X X =X J
 
the perturbation eigenproblem residue (PER) for pre- ∂u R ∂u L
+C − , (3)
dicting stability changes in the forced problem. In ∂t ∂t X =X J
Sect. 4 we apply both methods to determine the stabil- 
∂u R
ity of the two example systems studied in [1]. The paper Ey A = K (u R −u L ) + ε (u R −u L )3
∂X X =X J
concludes with discussions of the results and avenues  
for future work. ∂u R ∂u L
+C − , (4)
∂t ∂t X =X J

where X J = (α1 + α2 )l. The joint considered has lin-


ear and cubic restoring stiffness characterized by coef-

123
Wave-based analysis of jointed elastic bars: stability 1973

Fig. 1 Two linearly elastic bars joined by a nonlinear spring conjugate—this is only shown for a+ , but assumed for all other
and a linear dashpot. Each bar has one fixed end. Single- coefficients. The global coordinate X denotes position starting
frequency plane waves are injected at the location of the forcing. from the left end. Periodic solutions to this problem have been
Note that each wave coefficient includes itself plus its complex found in [1]

ficients K and ε, respectively, and linear damping −δu L (X J , t)) + C (δ u̇ R (X J , t) − δ u̇ L (X J , t)) ,
characterized by C. Note that ε denotes a small book- (10)
keeping parameter later set to one [4]. δu L (0, t) = 0, (11)
We identify any periodic solution [1] with left and
δu R (0, t) = 0, (12)
right displacement fields denoted by u ∗L (X, t) and
u ∗R (X, t), respectively. Next, we assess the local stabil-
ity of these solutions by introducing small perturbations where u ∗J (t) ≡ u ∗R (X J , t)−u ∗L (X J , t) denotes the joint
δu L (X, t) and δu R (X, t) such that the total displace- displacement. Retaining only terms up to and includ-
ment fields are now given by, ing O(ε) in the stability analysis (i.e., only retaining the
fundamental harmonic in u ∗J (t)), we rewrite the brack-
u L (X, t) =u ∗L (X, t) + δu L (X, t), 0 < X < XJ, eted term on the right-hand side of Eqs. (9), and (10)
(5) as,
u R (X, t) =u ∗R (X, t) + δu R (X, t), X J < X < l.   
(6) K + 3εu ∗2 J (t) = K + ε P cos(2ωt), (13)

Substituting the above into the governing field equa- where we have used u ∗J (t) ≡ a ∗ cos (ωt + φ ∗ ), φ ∗
tions, Eqs. (1), and (2), the joint conditions, Eqs. (3), denotes the joint displacement phase relative to the
 ∗2 3a ∗2
and (4), and the zero displacement boundary condi- forcing, K ≡ K + 3εa 2 , and P ≡ 2 . We note that
tions, we retain only linear terms in δu L (X, t) and both the amplitude, a ∗ , and phase, φ ∗ , are known from
δu R (X, t), yielding the solutions obtained in [1], and these fully distinguish
one solution from another.
∂ 2 δu L ∂ 2 δu L Using Eq. (13) in Eqs. (7–12), we recover a para-
ρA − Ey A = 0, 0 < X < X J , (7)
∂t 2 ∂ X2 metrically excited problem as illustrated in Fig. 2(a).
∂ 2u R ∂ 2u R While this problem arises from studying the stability
ρ A 2 − Ey A = 0, X J < X < l, (8) of a directly excited system, it may be of general inter-
∂t  ∂ X2
∂δu L    est. For example, similar problems will arise when two
Ey A  = K + 3εu ∗2 J (t) (δu R (X J , t) domains are connected by time-varying stiffness, such
∂ X X =X J
as in parametrically driven micromechanical oscillators
−δu L (X J , t)) + C (δ u̇ R (X J , t) − δ u̇ L (X J , t)) , (9) [6]. For the forced problem, of interest herein, stability
  
∂δu R  ∗2 of the solution is guaranteed when the parametrically
Ey A = K + 3εu (t) (δu R (X J , t)
∂X  X =X J
J excited problem is stable.

123
1974 N. N. Balaji et al.

Fig. 2 a Damped,
parametrically excited
system arising from the
local stability analysis of the
corresponding forced
system; solution approach at
b zeroth-order, c first-order,
and d second-order

We next proceed to find wave-based solutions to the into Eqs. (7–13) results in three ordered problems, as
linear problem illustrated in Fig. 2(a). To do so, we illustrated in Fig. 2(b–d), governed by,
employ a strained parameter approach. We first expand
δu L and δu R , 
(0) (1)
O ε0 :
δu L (X, t) = u L (X, t) + εu L (X, t)
(2)
 ∂ 2 u (0) ∂ 2 u (0)
+ ε2 u L (X, t) + O ε3 , (14) ρA L
− Ey A L
= 0, 0 < X < X J , (17)
∂t 2 ∂ X2
(0) (1)
δu R (X, t) = u R (X, t) + εu R (X, t) ∂ 2 u (0) ∂ 2 u (0)
 ρA R
− E y A R
= 0, X J < X < l, (18)
(2) ∂t 2  ∂ X2
+ ε2 u R (X, t) + O ε3 , (15)
∂u (0)   
(0) 
Ey A L  = K 0 u (0) − u L  X =X , (19)
and subsequently expand the static component of the ∂X  R
J
X =X J
spring stiffness, (0)
u L (0, t) = 0, (20)

 
∂u R   
K = K 0 + εK 1 + ε2 K 2 + O ε3 , (16) (0)
(0) (0) 
Ey A  = K0 u R − u L  , (21)
which amounts to “straining” the stiffness in the prob- ∂X  X =X J
X =X J
lem. We also place the damping at first order to avoid (0)
u R (l, t) = 0, (22)
decay in the ensuing zeroth-order problem: C →
Ĉ. 
Substitution of the expanded and ordered quantities O ε1 :

123
Wave-based analysis of jointed elastic bars: stability 1975

∂ 2 u (1) ∂ 2 u (1) we assign wave coefficients a0+ , a0− , . . ., d0− and their
ρA L
− E y A L
= 0, 0 < X < X J , (23)
∂t 2 ∂ X2 complex conjugates. Note that the subscript notation
(1) (1) here is different from that used in [1] where subscripts
∂ 2u R ∂ 2u R
ρA − E y A = 0, X J < X < l, (24) were used to denote the harmonic index while they are
∂t 2  ∂ X2
used here to denote the ordered wave components of
∂u   
(1)
(1) (1) 
Ey A L  = K0 u R − u L  + F (1) (t), the perturbation analysis. These coefficients are related
∂X  X =X J by propagation relationships,
X =X J
(25)
(1) b0+ = a0+ eikβ1 l , a0− = b0− eikβ1 l , d0+
u L (0, t) = 0, (26)
 = c0+ eikβ2 l , c0− = d0− eikβ2 l , (37)
∂u R   
(1)
(1) (1)  (1)
Ey A  = K0 u R − u L  +F (t),
∂X  X =X J joint conditions,
X =X J
(27)
i E y Ak −2K 0
(1)
u R (l, t) = 0, (28) c0+ = c0− + b+ ,
 −2K 0 + i E y Ak −2K 0 + i E y Ak 0
O ε2 : i E y Ak −2K 0
b0− = b0+ + c− ,
(2) (2) −2K 0 + i E y Ak −2K 0 + i E y Ak 0
∂ 2u L ∂ 2u L
ρA − Ey A = 0, 0 < X < X J , (29) (38)
∂t 2 ∂ X2
(2) (2)
∂ 2u R ∂ 2u R and clamped boundary conditions,
ρA − Ey A = 0, X J < X < l, (30)
∂t 2  ∂ X2
∂u    a0+ = −a0− , d0+ = −d0− ,
(2) (39)
(2) (2) 
Ey A L  = K0 u R − u L  + F (2) (t),
∂X  X =X J
X =X J where k denotes the wavenumber associated with the
(31)
system’s j th natural frequency, ω j . We point the reader
(2)
u L (0, t) = 0, (32) to more details on wave-based approaches in [1], par-

∂u R   
(2) ticularly as concerns joint conditions for the wave coef-
(2) (2) 
Ey A  = K0 u R − u L  + F (2) (t) ficients. Note that the zeroth-order stability problem is
∂X  X =X J
an eigenvalue problem. Next, we assemble the coeffi-
X =X J
(33) cient relationships into the form,
(2)

u R (l, t) = 0, (34) A(0)
s ω j ; K 0 zs(0) = 0, (40)
(0)
where zs denotes an eigenvector holding the eight
where the forcing arising at the joints at O(ε1 ) and
zeroth-order

wave coefficients (see Fig. 2(b)) and
O(ε2 ), respectively, is given by:
A(0)
s ω j ; K 0 denotes the zeroth-order stability coeffi-
 cient matrix as a function of the j th natural frequency
F (1) (t) = (K 1 + P cos(2ωt)) u (0) (0)
R (X J , t) − u L (X J , t) ω j and parameterized by stiffness K 0 . Note that the
 (0)
(0) (0)
+ Ĉ u̇ R (X J , t) − u̇ L (X J , t) , (35) dependence of As on frequency is nonlinear, and as a
 result, there are an infinite number of natural frequen-
(1) (1)
F (2) (t) = (K 1 + P cos(2ωt)) u R (X J , t) − u L (X J , t) cies satisfying Eq. (40) consistent with the fact that the

(0) (0) jointed system is continuous. While it is not apparent
+ K 2 u R (X J , t) − u L (X J , t)
 until the next order analysis, we also note that K 0 will
+ Ĉ u̇ (1) (1)
R (X J , t) − u̇ L (X J , t) . (36) be chosen based on requiring the eigenfrequency ω j to
be equal to the forcing frequency ω. The zeroth-order
Similar to the wave-based solution procedure for the problem is completed by finding the eigenvalues and
nonlinear forced problem [1], we seek exact wave solu- eigenvectors associated with Eq. (40) [7].
tions for the three linear problems shown in Fig. 2(b– Following completion of the zeroth-order problem,
d). Starting with the zeroth-order problem in Fig. 2(b), we turn our attention to the first-order problem. We note

123
1976 N. N. Balaji et al.

that the zeroth-order spring stretch required to evaluate We note that the two values of K 1 in Eqs. (44) begin
Eq. (35) is now known and given by, to form two sides of a stability tongue in the P versus
(0) (0)

K plane, as seen later. With known values for K 1 , it is
u R (X J , t) − u L (X J , t) = a (0) cos ω j (K 0 )t

also possible to find the eigenvectors, but these are not
+b(0) sin ω j (K 0 )t , (41) needed from this point forward.
where explicit dependence of ω j on K 0 is noted. Since With secular terms removed, we must still find
the zeroth-order solution is an eigenfunction, we can the forced response associated with the non-secular
freely choose the harmonic amplitudes a (0) and b(0) . terms in Eq. (42). As is the usual practice in pertur-
Substitution of Eq. (41) into Eq. (35) yields an updated bation approaches [4], we neglect the homogeneous
first-order forcing, (or unforced) solution in all orders above the zeroth-
order without loss of generality. With the choice of
Pa (0)
K 0 established, the non-secular forcing occurs at fre-
F (1) (t) = cos (2ω − ω j (K 0 ))t quency 3ω j (K 0 ). We find the forced response of the
2
O(ε1 ) problem using another exact wave approach,
Pb(0)

− sin (2ω − ω j (K 0 ))t as depicted by the wave coefficients in Fig. 2(c) and
2 their complex conjugates. These first-order wave coef-
Pa (0)
ficients are related by propagation relationships,
+ cos (2ω + ω j (K 0 ))t
2
Pb(0)
b1+ = a1+ ei3kβ1 l , a1− = b1− ei3kβ1 l ,
+ sin (2ω + ω j (K 0 ))t
 2 d1+ = c1+ ei3kβ2 l , c1− = d1− ei3kβ2 l , (45)
(0) (0)
+ K 1 a + Ĉω j (K 0 )b cos(ω j (K 0 )t)
 joint conditions,
+ K 1 b(0) − Ĉω j (K 0 )a (0) sin(ω j (K 0 )t).
i3E y Ak −2K 0
(42) c1+ = c1− + b+
−2K 0 + i3E y Ak −2K 0 + i3E y Ak 3
P (0) i P (0)
We note secular, or resonant, terms on the right-hand 4a 4 b
+ + ,
side of Eq. (42) having frequency dependence equal −2K 0 + i3E y Ak −2K 0 + i3E y Ak
to one of the zeroth-order system’s eigenfrequencies; i3E y Ak −2K 0
b1− = b+ + c−
namely the underlined terms. These must be eliminated −2K 0 + i3E y Ak 1 −2K 0 + i3E y Ak 1
in order for the response to be bounded and to satisfy P (0) i P (0)
4a 4 b
the assumed ordering of the asymptotic expansions. In − − , (46)
−2K 0 + i3E y Ak −2K 0 + i3E y Ak
addition, for particular choices of K 0 such that 2ω −
ω j (K 0 ) = ω j (K 0 ), additional secular terms appear as and clamped boundary conditions,
indicated by the doubly underlined terms. In particular,
the latter condition holds for a certain value of K 0 close, a1+ = −a1− , d1+ = −d1− , (47)
but not equal, to K from the forced problem. When
this occurs, the excitation frequency ω equals the j th where the joint conditions have been developed similar
eigenfrequency ω j (K 0 ), as remarked during the zeroth- to the discussion in [1] with the exclusion of damping
order problem. (0)
and the inclusion of imposed forces Pa4 e−3ω j (K 0 )t +
Proceeding with the choice of K 0 such that ω =
i Pb(0) −3ω j (K 0 )t
ω j (K 0 ), we remove secular terms by requiring, 4 e + c.c.. Similar to the zeroth-order
    problem, we assemble the coefficient relationships into
K 1 + P2 Ĉω a (0) 0 matrix form,
= . (43)
−Ĉω K 1 − P2 b(0) 0

A(1)
s 3ω j (K 0 ) zs(1) = fs(1) , (48)
Zeroing the determinant of the coefficient matrix in Eq. (1)
(43) yields K 1 , where
zs holds the eight first-order wave coefficients,
 A(1)
s 3ω j (K 0 ) denotes the first-order stability coeffi-
1 (1)
K1 = ± P 2 − 4Ĉ 2 ω2 . (44) cient matrix, and fs holds forcing terms proportional
2

123
Wave-based analysis of jointed elastic bars: stability 1977

to P4 a (0) and i 4P b(0) . Unlike the eigenvalue problem at


 3εa ∗2
forced problem, we recall that K ≡ K + 2 and
3a ∗2
zeroth-order, Eq. (48) admits a single
(1)

solution
for the P≡ 2 , where a ∗ and φ ∗
denote the joint displace-
coefficients via inversion of As 3ω j (K 0 ) , when this ment’s amplitude and phase for the periodic solution
inverse exists. For cases where this is not possible (see under consideration. We also note that Ĉ and  always
discussion of one such case in Sect. 4), it is implied that appear with ε, and thus, no generality is lost in setting
the strained system is resonant at 3ω j (K 0 ) in addition ε to one. Inverting Eq. (52) yields two curves defining
to having ω j (K 0 ) as a resonance. We will, however, the stability tongue [8] separating stable and unstable
not dwell on this aspect presently and proceed with 
solutions in the P versus K plane (depicted in Fig. 7
the general formulation of the method assuming that
(1)
for the single-jointed case).
As 3ω j (K 0 ) is non-singular. We note that the strained parameter approach may
Following the solution of Eq. (48), the spring dis- also be developed for the two-jointed system studied in
placement required in the second-order problem can be [1], where it is only necessary to strain one of the two
written as, joint stiffnesses, regardless of the fact that these two


u (1) (1)
R (X J , t) − u L (X J , t) = a
(1)
cos 3ω j (K 0 )t joints have the same stiffness—i.e., if they differed, it

would still be only necessary to strain one or the other,
+b(1) sin 3ω j (K 0 )t ,
and not both. This is due to the fact that the stabil-
(49) ity problem has codimension-1, i.e., only one resonant
where we note that a (1) depends only on a (0) (and not condition must be met, and thus, only one parameter
b ), and b depends only on b (and not a (0) ). In
(0) (1) (0) must be strained [9].
(1) (1)
fact, aa (0) = bb(0) . This results from the similarity in how
a (0) and b(0) appear in the first-order joint relationships. 3 Stability via the perturbed eigenproblem residue
With the first-order problem complete, we turn final approach
attention to the second-order problem. Substitution of
ω = ω j (K 0 ) and Eqs. (41), (49) into Eq. (36) yields an The second stability approach is based on a PWE for-
updated second-order forcing, of which only the secular mulation of the perturbed system of the original prob-
portion is of interest, lem. We consider a generic single jointed-bar vibration
 
Pa (1) problem of the form
(2)
Fsec (t) = + K 2 a (0) cos(ω j (K 0 )t)
2
  ∂ 2u L ∂ 2u L
ρA − E A = f L (X, t), 0 < X < X J
Pb(1) (0) ∂t 2
y
∂ X2
+ + K2b sin(ω j (K 0 )t). (50)
2 (53)
Elimination of the two secular terms yields, ∂ 2u R ∂ 2u R
ρA − E y A = f R (X, t), XJ < X < l
P a (1) P b(1) P ∂t 2 ∂ X2
K2 = − = − = − r (1)/(0) , (51) (54)
2 a (0) 2 b(0) 2 
∂u L
where we introduce r (1)/(0) to represent the ratio of Ey A = f nl ( u, u̇, . . . ) (55)
∂X X =X J
first- to zeroth-order spring displacement components. 
Lastly, we reconstitute the strained stiffness, ∂u R
Ey A = f nl ( u, u̇, . . . ), (56)
  ∂X X =X J
 1 P
K = K0 ± ε P 2 − 4Ĉ 2 ω2 − ε2 r (1)/(0) +O ε3 , u L (0, t) = 0, (57)
2 2
(52) u R (l, t) = 0, (58)
recalling that K 0 is found by enforcing that the exci-
tation frequency ω equals the j th natural frequency of and formulate the approach using this case as an exam-
ple. Application to cases with multiple joints is straight-
the zeroth-order stability problem, ω = ω j (K 0 ). It is forward since the method does not make assumptions
apparent from Eq. (52) that non-zero damping lifts the on the number or types of nonlinearities present. In the
stability curves off of the K  axis, and thus has a stabi- above, f L (X, t) and f R (X, t) denote general forcing
lizing effect. To return fully to assessing stability of the functions (independent of the solution u L ,R ). We note

123
1978 N. N. Balaji et al.

that this approach does not demand the presence of a instance, around the joint as
small parameter, i.e., this approach is applicable for
arbitrarily strong nonlinearities and not just weak non- 
δu L (X, t) = d+
j e
jk X − jλt
e + d−
j e
− jk X − jλt
e
linearities (as was the case for the strained parameter
j
approach presented in Sect. 2). Following the notation
in Sect. 2, we use starred superscripts to denote the + c.c. α1l < X < (α1 + α2 )l, (67)
periodic solutions u ∗L ,R of the original forced vibration
problem and δu L ,R to denote infinitesimal perturba- where k in the above comes from the dispersion rela-
tions of the √
tionship λ = E/ρk. Here, λ denotes the unknown
complex frequency that the perturbed system responds
left solution ← u ∗L (X, t) + δu L (X, t), 0 < X < X J , at. Assembling this system and using the FFT to obtain
(59) the wave-component stiffness coefficient contributions
right solution ← u ∗R (X, t) + δu R (X, t), X J < X < l. for the periodic stiffness terms in Eqs. (63) and (64)
(60) above (which are 2π/ω-periodic), we obtain a linear
set of equations of the form
Substituting these expressions into Eqs. (53–58) and A(λ; P, u ∗ , ω)z = 0 (68)
dropping the higher-order terms yields a homogeneous where z ∈ C Nw Nh is a complex vector consisting of the
set of governing equations for δu L ,R . These can be Nh complex harmonics of the Nw wave components;
expressed as A ∈ C Nr Nh ×Nr Nh is a complex matrix that is a func-
tion of λ alone (but parameterized by the periodic solu-
∂ 2 δu L ∂ 2 δu L tion u ∗ , ω). Note that no AFT (Alternating Frequency
ρA − Ey A = 0, 0 < X < X J (61) -Time) procedure needs to be carried out in terms of
∂t 2 ∂ X2
the complex frequency λ here; we use the appropriate
∂ 2 δu R ∂ 2 δu R
ρA − E y A = 0, X J < X < l (62) entries of the frequency-domain Jacobian of the PWE
∂t 2 ∂ X
2
 residue function (see [1]) to get the periodic stiffness
∂δu L ∂ f nl
Ey A = δu components once. As expected from Floquet theory,
∂ X X =X J ∂u u ∗ ,... this is a Nonlinear EigenProblem (NEP) in λ. How-

∂ f nl ever, since A contains exponentials of λ, it is not, in
+ ˙
δu (63)
∂ u̇ u ∗ ,... general, possible to spectrally decompose this problem
  quite easily and, furthermore, there exist an infinite set
∂δu R ∂ f nl
Ey A = δu of eigenpairs. We unsuccessfully attempted the use of
∂ X X =X J ∂u u ∗ ,...
 the Padé approximant and interpolatory Krylov tech-
∂ f nl
+ ˙
δu. (64) niques1 for partial spectral decomposition of this NEP,
∂ u̇ u ∗ ,... but still feel that such approaches must be explored in
δu L (0, t) = 0, (65) detail in the future to enable quantitative stability anal-
δu R (l, t) = 0. (66) ysis.
Qualitative analysis of stability can, however, be
carried out merely by seeking out the points along a
Since u ∗L ,R (X, t) (and therefore u ∗ , u˙∗ , . . . ) is given forced response curve where transitions of sta-
fixed from the solution of Eqs. (53–58), the above is bility occur in a manner that completely avoids the chal-
a linear homogeneous PDE with time-varying linear lenges associated with spectral decomposition. Floquet
stiffness and damping acting at the joint, and the solu- theory asserts [4] that stability transitions occur when
tion can be obtained using a wave-based approach. An the perturbed system has as eigenvalue imξ , a purely
important aspect in the homogeneous case, however, imaginary quantity with integer m > 0, with three pos-
is that the frequency of the response is complex and sibilities for ξ :
unknown, unlike the inhomogeneous case where the
ξ = 0; Perturbation remains unchanging in time;
frequency is real and fixed by the excitation. Resort-
ing to a multi-harmonic wave-based representation of 1 As implemented in http://guettel.com/rktoolbox/examples/
the solution, the solution δu L can be represented, for html/example_nlep.html.

123
Wave-based analysis of jointed elastic bars: stability 1979


ξ = ω; Perturbation varies periodically with the det (Â) will be referred to as the Perturbation
λ=iξ
same fundamental period as the solution; Eigenproblem Residue (PER) henceforth. The sign-
and change property will be clear if elaborated upon in the
ξ = ω/2; Perturbation varies periodically with twice following manner: suppose the true eigenvalues of Eq.
the fundamental period as the solution. (70) are denoted by γ j with j being the index of the
Choosing ξ = 0 gives rise to the trivial solution and eigenvalue, we can write the PER as a characteristic
need not be considered in the context of continuum polynomial in the following manner:
dynamics with homogeneous essential boundary con- ∞

ditions. In a multi-harmonic approach such as the wave- P E R = pÂ(λ) = α (λ − γ j )
based strategies considered presently, m can be set to j=1

 ∞
1 without loss of generality since the higher harmonics λ − γj
will be intrinsically included for each ξ . In our con- =α |λ − γ j |ei j=1 (72)
j=1
text, therefore, the cases corresponding to ξ = ω and
ξ = ω/2 will have to be considered.
Here |(.)| and (.) denote the absolute value and angle
If ξ is an eigenvalue
 of the system in Eq. (68), the
of their complex arguments respectively, and α is some
determinant of A λ=iξ will be exactly zero. Checking
real constant. A sufficient condition for the existence
for points along any given forced response curve where
and validity of such a polynomial representation is the
this happens allows one to obtain the stability transi-
analyticity of the PER for all points λ on the complex
tion or critical points directly. Since A is complex, the
domain. The existence of a Taylor series representation
determinant is, in general (for the cases when ξ is not an
is guaranteed in this case, justifying the use of the poly-
eigenvalue), also complex, making the task of detecting
nomial form above. This is clearly the case here since
the zero point on a forced response curve challenging.
λ only appears as smooth functions in the matrix A(λ)
We first therefore create a fully real counterpart of the
(and therefore in its determinant). This analyticity con-
Jacobian matrix (this process was previously alluded to
dition could be violated in some problems with non-
in [1]). Denoting by {z} and
{z} the real and imagi-
smooth nonlinearities (e.g., Filippov dynamical sys-
nary parts of the vector of wave component harmonics
tems, see, for instance [10]). In these cases, classical
z, Eq. (68) can be rewritten as
Floquet theory is not strictly applicable and further con-
siderations become necessary. We will presently not
Az = A {z} + iA
{z}
  concern ourselves with such issues and proceed with
  {z} the discussion making note of the fact that this frame-
= A iA . (69)

{z} work is invalid for such cases. We also note here that
there are no bounded values of λ for which the PER
Symbols denoting the functional and parametric depen- becomes unbounded; i.e., the characteristic polynomial
dences of A are dropped in the above for brevity. does not have any poles on the complex plane. It, how-
Decomposing the complex quantity Az into its real and ever, has an infinite number of zeros (despite the finite
imaginary parts denoted by {Az} and
{Az}, respec- size of Â), referred interchangeably here as eigenvalues
tively, yields the following fully real form of Eq. (68): of the NEP, consistent with that expected in continuum
      
{Az} {A} {iA} {z} 0 dynamics.
= = . (70) Since the PER is always real, the effective phase

{Az}
{A}
{iA}
{z} 0 ∞ ◦ ◦
j=1 λ − γ j is either 0 or 180 . Consider two points
The Jacobian matrix of this system, right before and after a transition in stability such as
 
{A} {iA} shown in Fig. 3. Just before the critical point (point A
 = , (71) (A)

{A}
{iA} in Fig. 3), the zero γm closest to iξ will lie just to its
left (since the solution is known to be stable) such that
is fully real and its determinant will be real always.
the quantity iξ − γm(A) equals 0◦ . Equivalently, the
Along a forced response curve, a zero-crossing of (A)
vector in the complex plane joining γm to iξ points to
the determinant of A]λ=iξ will be the same as a sign
the right since this solution is known to be stable. The
change in the determinant of  . The quantity total angle of the PER at this point can be expressed as
λ=iξ

123
1980 N. N. Balaji et al.

ing response amplitude, the zeros away from iξ are


assumed to undergo only negligible changes and there-
(A) (B)
fore γ j  γ j is assumed for j = m. The first term
in Eq. (74) is approximated as P E R A (see Eq. (73)
above) and can be simplified as
P E R B = P E R A + 180◦ . (75)
This shows that the difference in complex angle
between the PER values across a critical point is exactly
180◦ . Since the PER is always a real number, this
amounts to a sign change as mentioned earlier. Sim-
ilar reasoning can be employed to observe that a sign
change will also happen at the other critical point in
Fig. 3 A schematic forced response (stable: blue, solid lines; Fig. 3 where the solutions transition from unstable to
unstable: red, dashed lines) with insets showing the complex stable.
plane including the point iξ , the corresponding closest zero One drawback of the method presented is the fact
(A,B) (A,B)
γm , and the complex vector (iξ − γm ) denoted using that it can only predict stability changes and it is not
a line segment and terminal arrow. A and B are two points on
the response curve across the stability transition possible, in general, to determine the stability of a solu-
tion from just the sign or the value of the PER. Exact
stability determination would instead require knowl-
the sum, edge of the location in the left or right half-plane for all

 infinite number of γ j , of which little can be said at this
(A)
P E RA = iξ − γ j + iξ − γm(A) point. An exploration of the necessary conditions for
j=1 the PER sign change (stability transition is just a suffi-
j =m
cient condition) is deferred for future research. But the
∞
(A) efficacy of the approach is empirically demonstrated
= iξ − γ j , (73)
by the fact that it provides satisfactory estimates of the
j=1
j =m stable and unstable branches on the response curve for
both the numerical examples considered in [1].
where the zero term is canceled in the final expression,
which itself has to be either 180◦ or 0◦ since the PER
is a real quantity. 4 Stability results for an example system
Looking at a point just after the transition (point B
(B)
in Fig. 3), the zero γm has crossed through iξ and In this section we present stability results for the
will now lie just to its right. Consequently, the quantity single-jointed example problem illustrated in Fig. 1
(B)
iξ − γm now takes a value of 180◦ , and we can and parameterized by quantities appearing in Table 1.
express the total angle of the PER at this point as the Periodic solutions to this problem are presented in [1],
sum, which also contains stability results for a second exam-
ple problem consisting of three bars connected by two

 nonlinear joints.
P E RB = iξ − γ j(B) + iξ − γm(B) The stability boundary predicted by the strained
j=1 parameter approach for a fixed excitation frequency
j =m

ω is a parabolic curve (known as a stability or Arnold
 (B) tongue), in the P − K  space, where P denotes the
= iξ − γ j + 180◦ . (74)
j=1
periodic stiffness and K  the constant stiffness of the
j =m parametrically forced system (recall
set to 1)—such a
curve will be presented later. If, in addition, we vary the
Around these transition points, due to the arbitrarily frequency, a stability surface results separating regions
small change in the excitation frequency ω and ensu- of stable and unstable parametrically forced systems.

123
Wave-based analysis of jointed elastic bars: stability 1981

Table 1 Mechanical and geometric parameters used for the solution surface, shaded blue in Fig. 4 and labeled as
single-jointed bar example (see Fig. 1 for accompanying the Solution Manifold. The solution curves and mani-
schematic)
fold occupy more space inside of the stability bound-
Parameter Value ary as the forcing increases, as depicted in the pro-
Ey 262 GPa
gression of sub-figures. For example, at 7.5 MN forc-
ing, the solution manifold never intersects the stability
ρ 1280 kg/m3
boundary, indicating all solutions are stable. In fact,
K 109 N/m
the solutions for 7.5 MN forcing closely resemble the
C 320 N/(ms−1 )
linear solutions such that no appreciable bending of
 108 N/m3 the frequency response curve occurs, and thus multi-
A 1.7145× 10−3 m2 ple solutions near resonance have yet to appear. At 15
l 1.0 m MN and 30 MN forcing, appreciable bending occurs
α1 0.28 together with multiple solutions and the solution man-
α1 + α2 1/3 ifolds intersect with the stability boundaries. For these
α3 2/3 forcing levels, the portions of the response curves asso-
ciated with mid-amplitude response are now unstable.
Lastly, we note that the intersection of the surface man-
ifold with the stability boundary marks the Transition
We present this surface in the subplots of Fig. 4 near the Boundary, as indicated by the solid magenta curves in
second natural frequency of the single-jointed system the sub-figures.
shown in Fig. 1, and label it as the Stability Boundary— We study the convergence of the stability tongues as
note that this boundary is the same surface in all six the analysis order increases in Fig. 5. The sub-figure on
sub-figures. We provide a narrated video in the sup- the left presents the frequency response curve for the
plementary material which includes revolving views single-jointed system studied, evaluated with a forc-
of the figures and further explanations. The stability ing of 45 MN. At such a large forcing level, the ratio
analysis is carried-out up to, and including, first-order of nonlinear to linear restoring force at resonance is
terms. When we return to the directly forced system, 0.25, well-above that which can be expected for solu-
specifying values for the joint linear and nonlinear stiff- tion accuracy. However, this forcing level yields three
ness, K and , respectively, together with the excita- solutions whose stability can be used to test the accu-
tion frequency ω and response amplitude a ∗ for a given racy of first-order versus second-order stability anal-
periodic solution, yields a point in the K  − P − ω yses. The three solutions are indicated in the left sub-
space; see Eq. (13). If this point falls inside of the sta- figure at 81.6 k rad/s using three markers. Subsequently,

bility boundary, the solution is unstable, whereas as if values for P and K for each are computed and indi-
it falls outside, the solution is stable. If in addition we cated with the same markers in the right sub-figure,
sweep frequency, response curves result which lie in together with the stability tongue corresponding to 81.6
the same space - these appear in Fig. 4 as curves whose k rad/s. We note from the right sub-figure that the sta-
stable portions are indicated by solid black lines, and bility tongues match closely at low values of P, and
whose unstable portions by dashed green lines. A first- thus, only first-order stability is necessary to assess
order WBVA approach is used to obtain the solutions the stability of solutions at low forcing, justifying the
(see [1] for details) and a first-order strained parameter use of first-order analysis in Fig. 4. At higher forc-
approach is used to obtain the stability boundaries. We ing amplitudes, the mid- and high-amplitude solutions
present solution curves at forcing levels of 7.5 MN, correspond to larger values of P. If first-order stabil-
15 MN, and 30 MN, and we provide two views for ity analysis is employed, the stability determination
each forcing level to aid visualization. It is important can be inaccurate, and thus second-order stability must

to note that while K is fixed along these curves, K is be employed. This is clearly documented in Fig. 5(b)

not, and thus the curves bend towards increasing K where the high-amplitude solution appears inside of
as resonance is approached and the response ampli- the first-order stability tongue, erroneously suggesting
tude a ∗ increases. If we also vary the linear stiffness unstable response. However, using second-order sta-
K in the directly forced problem, we can generate a bility analysis, this solution subsequently lies outside

123
1982 N. N. Balaji et al.

Fig. 4 Stability boundaries from the strained parameter amplitudes of a 7.5 MN; b 15 MN; and c 30 MN. Two views of
approach and the solution manifolds (first-order perturbation for the surface are presented in each sub-figure for clarity
both) of the single-jointed system corresponding to excitation

of the stability tongue as the tongue appears rotated diction might arise, we have found first-order analysis
away from the first-order tongue. In our investigations is always accurate for responses in which the ratio of
of solution response and stability using the WBVA and nonlinear to linear restoring force is less than 0.1.
the strained parameter approach, focusing specifically Next we investigate the PER stability approach in
on the resonance peak where issues in stability pre- the context of the single-jointed system. For all fre-

123
Wave-based analysis of jointed elastic bars: stability 1983

Fig. 5 Stability assessment of periodic solutions in the single- order stability analysis. Also shown in (b) are the P and K  pairs
jointed system around the second resonance, undergoing a 45 MN for the three periodic solutions where, in both sub-figures, a circle
amplitude periodic excitation: a first harmonic periodic response denotes the low-amplitude solution, squares the mid-amplitude
with three solutions indicated at 81.6 k rad/s, and b accompany- solution, and hexagrams the high-amplitude solution
ing stability boundaries at 81.6 k rad/s using first- and second-

quency responses considered, we set ξ equal to the present system, consistent with the strained parameter
excitation frequency ω in detecting critical points for approach. Figure 6 plots the frequency response and
which the sign of the PER changes. No critical points the perturbation eigenproblem residue corresponding
were found by setting ξ to ω/2, implying that stabil- to the PWE solution of the single-jointed system har-
ity loss occurs only through the ξ = ω case for the monically forced at an amplitude of 30 MN using both

Fig. 6 Frequency response (first harmonic amplitude) of the ear scale and b a log scale. Stability transition points are denoted
PWE solution of the single-jointed system undergoing harmonic by red points, while the stable and unstable solutions are plotted
excitation near the second resonance with amplitude 30 MN, in blue and red, respectively
along with the perturbation eigenproblem residue using a a lin-

123
1984 N. N. Balaji et al.

Fig. 7 First- and second-order stability boundaries from the strained parameter approach for the single-jointed system

Fig. 8 Comparison of the forced response stability predictions stability approach developed in Sect. 3. In all figures, the stable
for the single-jointed system a using the FE-HB approach with and unstable branches are plotted using solid and dashed lines,
frequency domain Hill’s coefficients [5] for stability determi- respectively
nation, and b using the PWE frequency response and the PER

a linearand log scale. Specifically, we plot the PER, The PER approach does not provide insight into the
det (Â) , evaluated using λ = iω. The left sub- mechanism for the log value exhibiting a local mini-
λ=iξ mum at approximately 79.8 k rad/s. Revisiting the sta-
figure shows the PER clearly changing signs at the two
bility boundaries predicted by the strained parameter
bifurcation points in the frequency response curve, cor-
approach, however, provides the requisite insight. Fig-
rectly indicating stability transitions in agreement with
ure 7 presents the stability boundaries when we carry-
the strained parameter approach. In the right sub-figure,
out the strained parameter approach to the first- and
the log value of the PER drops sharply at the bifurca-
second-orders. At the second order, which also neces-
tion points, but interestingly, shows a drop and local
sitates the inclusion of third harmonic terms, K 2 (and
minimum at approximately 79.8 k rad/s.

123
Wave-based analysis of jointed elastic bars: stability 1985

thus K  ) goes unbounded at the same point in frequency turbed eigenproblem. The strained parameter approach
where the PER exhibits a local minimum. Examining holds for only moderate nonlinear response due to its
the coefficient matrix A(3ω; K 0 ) evaluated near the reliance on an asymptotic expansion; the PER avoids
local minimum frequency, we find that it has a zero this limitation and in fact is used to determine stability
eigenvalue at 79.7636 k rad/s. Since the determinant of periodic solutions at large nonlinear response.
of a matrix equals the product of its eigenvalues, it is We now list some avenues for future research that
apparent that 3ω at this frequency satisfies an internal were identified during the course of this study. One
resonance condition. Further investigation of the inter- issue with the PER approach, as presented herein, is
nal resonance mechanism lies beyond the scope of this that it can only determine locations of stability change,
paper, but may offer an interesting direction for follow- and does not assess stability directly. Therefore, a PER
on efforts. approach that can assess the latter warrants inves-
Lastly, Fig. 8 presents the finite element validation tigation. Further, the PER method described in this
and PWE forced response solutions for the single- paper only involves a sufficient condition for the sign
jointed system, where we assess stability of the PWE change. An investigation of the necessary conditions
solutions using the PER approach. We note strong could potentially yield greater insights. Both stability
agreement between the two approaches in both the pre- approaches (strained parameter as well as PER) rely
dicted response and accompanying stability. Notably, on Floquet theory, whose applicability for non-smooth
at a forcing level of 45 MN, the ratio of nonlinear to problems is not very clear at this point. This will have
linear restoring force at resonance is over 0.7, which to be studied in detail (see [10], for instance) to under-
is well-beyond the validity region for the WBVA and stand the theoretical limitations of the methods more
the strained parameter stability approach. However, the clearly. The internal resonance observed in the single-
PWE combined with the PER stability approach accu- jointed problem (see Fig. 6 and related discussions)
rately captures both the frequency response and the should also be given additional attention.
multiple solution stability at all amplitudes considered.
Funding Authors N.N. Balaji and M.R.W. Brake acknowl-
This can be further contrasted with the poor perfor- edge support from the National Science Foundation under grant
mance of the finite-element substitution approach dis- No. 1847130. M.J. Leamy acknowledges support from the
cussed earlier and documented in Fig. 9(a). National Science Foundation under Grant No. 1929849.

Data availability The datasets generated during and/or analyzed


during the current study are not publicly available due to limi-
5 Concluding remarks tations on archiving data long-term, but are available from the
corresponding author on reasonable request.
As pointed out by previous researchers, a direct anal-
Declarations
ysis of stability has been lacking in the literature for
nonlinear wave-based solution approaches. We address Conflict of interest The authors have no relevant financial or
this deficiency by developing two direct approaches, non-financial interests to disclose.
and then apply the methods to determine solution sta-
bility of an harmonically forced, example two-bar sys-
tem connected by a nonlinear joint. The first stability Appendix A
approach expands (or strains) a stiffness parameter in
the problem, which allows one to construct analytical In this appendix, the PWE approach [1] is used to obtain
stability boundaries in the parameter space of the per- the frequency responses of the single jointed system
turbed problem using a wave-based method. Not only (Fig. 1) around its first resonance. We then project these
does this determine stability of periodic solutions aris- solutions onto a 90-element finite element mesh, for
ing from direct excitation of a continuous system, but use as initial conditions, and estimate its monodromy
it also directly governs the response and stability of matrix using transient simulations over a single period.
parametrically forced continuous systems, which is a The predicted stability using this hybrid approach is
new development in and of itself. The second stabil- indicated in Fig. 9(a). Comparing these stability pre-
ity approach is primarily a computational tool which dictions with pure finite element stability predictions
searches for sign changes in the residue of the per- (Fig. 9(b), which use frequency domain Hill’s coef-

123
1986 N. N. Balaji et al.

Fig. 9 Comparison of the forced response stability predictions HB-computed frequency responses using the frequency domain
for the single-jointed system a using the PWE-computed fre- Hill’s coefficient approach [5] (reproduced from Fig. 8). In all
quency responses with stability assessed using transient Flo- figures, the stable and unstable branches are plotted using solid
quet multipliers estimated from the FE model, and b the FE- and dashed lines, respectively

ficients [5]), shows that although the response curves 6. DeMartini, B.E., Rhoads, J.F., Turner, K.L., Shaw, S.W.,
are nearly the same, the stability predictions contradict, Moehlis, J.: Linear and nonlinear tuning of parametrically
excited mems oscillators. J. Microelectromech. Syst. 16(2),
specifically around the peaks at large forcing levels. 310–318 (2007)
This shows that the hybrid approach can lead to incor- 7. Lv, H., Leamy, M.J.: Damping frame vibrations using ane-
rect conclusions about solution stability, and motivates choic stubs: analysis using an exact wave-based approach.
the need for direct wave-based stability approaches. J. Vib. Acoust. (2021). https://doi.org/10.1115/1.4049388
8. Kovacic, I., Rand, R., Mohamed S.S.: Mathieu’s equation
and its generalizations: overview of stability charts and their
features. Appl. Mech. Rev., 70(2), (2018)
References 9. Luongo, A., Paolone, A.: On the reconstitution problem in
the multiple time-scale method. Nonlinear Dyn. 19(2), 135–
1. Balaji, N.N., Brake, M.R.W., Leamy, M.J.: Wave-based 158 (1999)
analysis of jointed elastic bars: nonlinear periodic response. 10. Leine, R.I., Nijmeijer, H.: Dynamics and Bifurcations of
Nonlinear Dynamics, in press, (2022) Non-Smooth Mechanical Systems. Springer, Berlin; Lon-
2. Chouvion, B.: Vibration analysis of beam structures with don, (2004). ISBN 978-3-642-06029-8
localized nonlinearities by a wave approach. J. Sound Vib.
439, 344–361 (2019). https://doi.org/10.1016/j.jsv.2018.09.
063 Publisher’s Note Springer Nature remains neutral with regard
3. Chouvion, B.: A wave approach to show the existence to jurisdictional claims in published maps and institutional affil-
of detached resonant curves in the frequency response of iations.
a beam with an attached nonlinear energy sink. Mech.
Res. Commun. 95, 16–22 (2019). https://doi.org/10.1016/ Springer Nature or its licensor (e.g. a society or other partner)
j.mechrescom.2018.11.006 holds exclusive rights to this article under a publishing agreement
4. Nayfeh, A.H., Mook, D.T.: Nonlinear Oscillations. Physics with the author(s) or other rightsholder(s); author self-archiving
Textbook. Wiley, Weinheim, wiley classics library ed. 1995, of the accepted manuscript version of this article is solely gov-
[nachdr.] edition, (2004) ISBN 978-0-471-12142-8 erned by the terms of such publishing agreement and applicable
5. Von Groll, G., Ewins, D.J.: The harmonic balance method law.
with arc-length continuation in rotor/stator contact prob-
lems. J. Sound Vib. 241(2), 223–233 (2001). https://doi.org/
10.1006/jsvi.2000.3298

123

You might also like