Processing math: 100%
Research article

Asymptotic finite-dimensional approximations for a class of extensible elastic systems

  • Received: 07 January 2021 Accepted: 21 June 2021 Published: 16 August 2021
  • We consider the equation

    utt+δut+A2u+Aθ/2u2Aθu=g

    where A2 is a diagonal, self-adjoint and positive-definite operator and θ[0,1] and we study some finite-dimensional approximations of the problem. First, we analyze the dynamics in the case when the forcing term g is a combination of a finite number of modes. Next, we estimate the error we commit by neglecting the modes larger than a given N. We then prove, for a particular class of forcing terms, a theoretical result allowing to study the distribution of the energy among the modes and, with this background, we refine the results. Some generalizations and applications to the study of the stability of suspension bridges are given.

    Citation: Matteo Fogato. Asymptotic finite-dimensional approximations for a class of extensible elastic systems[J]. Mathematics in Engineering, 2022, 4(4): 1-36. doi: 10.3934/mine.2022025

    Related Papers:

    [1] Jason Howell, Katelynn Huneycutt, Justin T. Webster, Spencer Wilder . A thorough look at the (in)stability of piston-theoretic beams. Mathematics in Engineering, 2019, 1(3): 614-647. doi: 10.3934/mine.2019.3.614
    [2] José Antonio Vélez-Pérez, Panayotis Panayotaros . Wannier functions and discrete NLS equations for nematicons. Mathematics in Engineering, 2019, 1(2): 309-326. doi: 10.3934/mine.2019.2.309
    [3] Alessandra De Luca, Veronica Felli . Unique continuation from the edge of a crack. Mathematics in Engineering, 2021, 3(3): 1-40. doi: 10.3934/mine.2021023
    [4] Yves Achdou, Ziad Kobeissi . Mean field games of controls: Finite difference approximations. Mathematics in Engineering, 2021, 3(3): 1-35. doi: 10.3934/mine.2021024
    [5] Michael Herrmann, Karsten Matthies . Solitary waves in atomic chains and peridynamical media. Mathematics in Engineering, 2019, 1(2): 281-308. doi: 10.3934/mine.2019.2.281
    [6] Zaffar Mehdi Dar, M. Arrutselvi, Chandru Muthusamy, Sundararajan Natarajan, Gianmarco Manzini . Virtual element approximations of the time-fractional nonlinear convection-diffusion equation on polygonal meshes. Mathematics in Engineering, 2025, 7(2): 96-129. doi: 10.3934/mine.2025005
    [7] Jana Kopfová, Petra Nábělková . Thermoelastoplastic oscillator with Prandtl-Ishlinskii operator. Mathematics in Engineering, 2025, 7(3): 264-280. doi: 10.3934/mine.2025012
    [8] Alena Kirsanova, Vladimir Sobolev . Padé approximants of canards and critical regimes of Darrieus wind turbine model. Mathematics in Engineering, 2025, 7(3): 194-207. doi: 10.3934/mine.2025009
    [9] Federico Cluni, Vittorio Gusella, Dimitri Mugnai, Edoardo Proietti Lippi, Patrizia Pucci . A mixed operator approach to peridynamics. Mathematics in Engineering, 2023, 5(5): 1-22. doi: 10.3934/mine.2023082
    [10] Serena Della Corte, Antonia Diana, Carlo Mantegazza . Global existence and stability for the modified Mullins–Sekerka and surface diffusion flow. Mathematics in Engineering, 2022, 4(6): 1-104. doi: 10.3934/mine.2022054
  • We consider the equation

    utt+δut+A2u+Aθ/2u2Aθu=g

    where A2 is a diagonal, self-adjoint and positive-definite operator and θ[0,1] and we study some finite-dimensional approximations of the problem. First, we analyze the dynamics in the case when the forcing term g is a combination of a finite number of modes. Next, we estimate the error we commit by neglecting the modes larger than a given N. We then prove, for a particular class of forcing terms, a theoretical result allowing to study the distribution of the energy among the modes and, with this background, we refine the results. Some generalizations and applications to the study of the stability of suspension bridges are given.



    Let A2 be a diagonal, self-adjoint, strictly positive operator, densely defined on a real Hilbert space (H,(,),) and we consider the following nonlinear nonlocal evolution equation

    utt+δut+A2u+Aθ/2u2Aθu=ginH×R+ (1.1)

    where θ[0,1], δ>0 and gC0(R+,H) is a given forcing term.

    The purpose of the present paper is to give a rigorous finite-dimensional approximation of (1.1). To be more precise, we introduce the projection PN onto the space generated by the first N modes, that is, by the first N eigenvectors of the operator A2 and we consider the approximated problem

    utt+δut+A2u+Aθ/2u2Aθu=PNginH×R+ (1.2)

    We remark that, by taking u(0) and ut(0) in PNH, Eq (1.2) can be interpreted as a system of N ODEs. Therefore, Eq (1.2) actually provides a finite-dimensional approximation of equation (1.1). We aim to prove that any solution of (1.2) is asymptotically finite-dimensional and to estimate, for any ε>0, the smallest N=N(ε) such that the asymptotic distance in the phase space between the solution of (1.1) and the corresponding solution of (1.2) is less than ε. An improvement of the result will be studied for a particular class of forcing terms.

    The reduction of infinite-dimensional dynamical systems to finite-dimensional systems of ODEs is a technique which has been widely used in the theoretical and numerical study of PDEs. The idea was first stated by Galerkin [28] and it has been used in many different applied frameworks as well as in the theory of finite-dimensional inertial manifolds (see [15,19,21,52,54,55] and the references therein). In particular, it is a fairly common procedure, which we aim to make rigorous, in the study of suspension bridges [3] to approximate the physical system with the dynamics finite number of modes in order to reduce the computational complexity of the model. This approach can be physically justified by observing that "the higher modes with their shorter waves involve sharper curvature in the truss and, therefore, grater bending moment at a given amplitude and accordingly reflect the influence of the truss stiffness to a greater degree than do the lower modes" [51,p.11], which means that the dynamics of the higher modes corresponds to a physically irrelevant phenomenon. We remark that our goal would not be achieved just by estimating the dimension of the inertial manifold of our system, since we are interested in providing a finite-dimensional approximation of its asymptotic behavior.

    The problem of finding a finite number of natural parameters of a system that uniquely determine its asymptotic behavior was first discussed for the 2D Navier-Stokes equation [24,43] and to tackle it the concepts of finite-dimensional inertial manifold, determining modes and, later, determining nodes and determining local volume averages were introduced (see [16,Ch. 5], [18] and the references therein). Regarding our problem, Chueshov in [16,Ch. 5,Thm. 7.2] proved that the dynamics of the first N modes of (1.1) completely determines the evolution of the system and Eden and Milani in [22] proved that if the forcing term is Ndimensional, then any solution is attracted to an Mdimensional manifold with MN.

    Some particular cases of the damped Eq (1.1) have been widely studied in mathematical literature. An ODE version of the problem was investigated by Loud in [44,45]. Fitouri and Haraux in [27] improved some of the previous results on the ODE case and in [26] they provided a close-to-optimal ultimate bound in the PDE version of the problem. More recently, some sharp stability criteria for the unimodal version of (1.1) and for a related evolution equation were obtained by Haraux in [37] in the case g=0. The case when θ=1 was studied in a slightly different framework by Holmes and others in [40,47] as an example of chaotic dynamics (see also [34]) and some undamped versions of (1.1) were studied in the case θ=0 by Cazenave, Weissler and Haraux in [11,12,13,14] in order to obtain a description of the qualitative behavior of more complicated nonlinearities and by Gazzola and Garrione in [29] to study the dynamics of suspension bridges with multiple intermediate piers.

    The considered abstract equation was analyzed by many other authors in an even more general framework. Biler [7] and de Brito [9] investigated the decay properties of the unforced problem with weak damping and a more general nonlinear nonlocal term. Da Silvia and Narciso [49,50] studied an extensible beam model subject to a nonlocal nonlinear parameter-dependent damping and a forcing term. A lot of different variations of (1.1) with a large variety of damping and nonlinear terms has been studied in mathematical literature (see [16,17,20] and the references therein).

    In addition to its mathematical relevance, our study also presents a certain physical and engineering interest. In fact, the considered model is suitable to describe both mono-dimensional and multi-dimensional physical systems. More precisely, some particular cases of (1.1) concerning the dynamics of beams and plates was considered by Holmes and Marsden [38,39] in order to study the problem of flow-induced oscillations (see also [41,42]) and in order to provide some more information about the nonlinear structural behavior of suspension bridges. In particular, we expect our results to allow some progress in the study of the structural and torsional instability of plates, to which a vast literature is devoted [1,2,4,5,31,32].

    If we set A2=Δ2, θ=1 and H=L2(Ω), where Ω is a bounded domain in RN (N1) with the smooth boundary Ω, we obtain the equation

    utt+δut+Δ2u+(Ω|u|2)Δu=g,inΩ×(0,T).

    This problem is a special case of the more general model

    utt+Δ2uϕ(u2L2(Ω))Δu=F(x,t,u,ut)

    that was introduced in 1955 by Berger [6] as a simplification of the von Karman plate equation which describes large deflection of plate. Some related models were later applied to the study of the torsional instability of suspension bridges. In particular, our results apply also to the partially-hinged plate problem discussed in [8,25]

    {utt+δut+Δ2u+(PSΩu2r(r,s,t)drds)uxx=ginΩ×(0,T)u=uxx=0on{0,π}×[l,l]uyy+σuxx=uyyy+(2σ)uxxy=0on[0,π]×{l,l}

    where S>0 depends on the elasticity of the material of the deck of the bridge, l>0 represents the width of the bridge and σ>0 is the Poisson's ratio of the structure, which is assumed to be, in the case of suspension bridges, between 0 and 0.5. The term P is called "prestressing constant" and it expresses the buckling loads on the plate. In the case of suspension bridges, the compressive forces along the edges are introduced in order to increase the stability of the structure. The abstract prestressed model reads

    utt+δut+A2uPAu+||Aθ/2u||2Aθu=ginH×R+. (1.3)

    The study of this equation will not be discussed in detail since, under the hypothesis P<α1/21 (weak prestressing), the prestressing term does not modify the qualitative behavior of the system and in the case when Pα1/21 (strong prestressing) our results do not hold. In fact, in a strongly prestressed suspension bridge the linear part of (1.3), which is given by A2PA, is not a strictly positive operator anymore.

    Concerning the case where the models describes the dynamics of a mono-dimensional structure, if we take H=L2(I) (with I=[π,π]) and A=xx, we can distinguish three different physically significant cases: θ=0, θ=1 and θ=2.

    In the first case, the considered model has been introduced by Garrione and Gazzola [29] in order to describe the behavior of the deck of suspension bridges with two intermediate piers. In the work of Garrione and Gazzola, the deck of the bridge is modeled by a degenerate plate consisting of a beam with a continuum of cross sections free to rotate around the beam. Therefore, the longitudinal dynamics of the bridge is modeled by a beam equation, whose nonlinear term can be interpreted as a representation of "a stiffened beam where the displacement behaves superquadratically and nonlocally: if the beam is displaced from its equilibrium position in some point, then this increases the resistance to further displacements in all the other points" [29]. The nonlocal nature of such term is due to the elastic behavior of the components of the bridge, the sustaining cables in particular. This choice of the nonlinear term follows from a comparison between the qualitative behavior of some possible models and the actual behavior of suspension bridges. If we consider D(A)={vH2(I)H10(I):v(π)=v(π)=v(aπ)=v(bπ)=0} for a,b(0,1), where a and b model the position of the piers along the deck of the bridge, the system reads

    {utt+δut+uxxxx+u2L2(I)u=g(x,t)t0,xIu(0)=u0H2(I)H10(I),ut(0)=u1L2(I)u(π,t)=u(πb,t)=u(πa,t)=u(π,t)=0,t0.

    An analogous equation, in a different functional framework, is involved in the study of the interaction between the cables and the deck of a suspension bridge in the case when the hangers are considered inextensible (see [29,46]).

    The second case (θ=1) was obtained by Woinowsky-Krieger [53] in 1950 and, independently, by Burgreen [10] in 1951. It models the physical phenomenon that "if the beam is stretched somewhere, then this increases the resistance to further stretching in all the other points" [29]. The system has been widely studied in both mathematical and engineering literature (see [22,33] and the references therein). If we choose D(A)={vH2(I)H10(I):v(π)=v(π)=vxx(π)=vxx(π)=0}, the model becomes

    {utt+δut+uxxxxux2L2(I)uxx=g(x,t)t0,xIu(0)=u0H2(I)H10(I),ut(0)=u1L2(I)u(π,t)=uxx(π,t)=uxx(π,t)=u(π,t)=0,t0.

    The case θ=2 was first introduced in [29]. If we consider H=L2(I) and A=xx as we did before, the nonlinear term u2θAθ/2u reads uxx2L2(I)uxxxx and the corresponding nonlinear equation can be interpreted as a model for "a stiffened beam with bending energy behaving superquadratically and nonlocally: this means that if the beam is bent somewhere, then this increases the resistance to further bending in all the other points" [29]. Despite the physical interest of the case θ=2, due to its technical difficulty, in this paper we decided to restrict ourselves to the cases where θ[0,1].

    The results of the paper are given in three main theorems. First, in Theorem 2.3, we prove that if the forcing term is finite-dimensional, i.e., if g is a combination of a finite number N of modes, then any solution is asymptotically finite-dimensional too in a sense that we specify in Definition 2.2. In the case of small oscillations or large damping, our result improves the one of Eden and Milani [22]. The proof is based on an application of a recent work of Haraux [37]. Next, in Theorem 2.4 we prove that, under suitable smallness conditions on the nonlinearity and on the forcing term, we are able to give an Mdimensional approximation of (1.1). More precisely, we prove that for any ε>0 there exists NN such that the asymptotic distance between a solution of (1.1) and a solution of (1.2) is controlled by ε in the phase space norm. The proof relies on a continuous dependence result and on Theorem 2.3. To conclude, in Theorem 2.5, fixed θ=0, we focus on a particular class of forcing terms and we refine the result of Theorem 2.4. In particular, under suitable smallness conditions on the solution, we improve the ultimate bounds previously given for general forcing terms in [8,26] and we estimate how much the dynamics changes as we eliminate a single mode from the dynamics. This latter result represents one of the main novelties of the paper since, to the author's knowledge, this is the first statement of this type present in literature.

    The paper is organized as follows. In Section 2 we give some definitions and we state the main results of the paper. In Section 3, some technical results are given. The proofs of the main results are contained in Section 4, Section 5 and Section 6, which are devoted to the proof of Theorem 2.3, Theorem 2.4 and Theorem 2.5 respectively. In Section 7, we present some physical conclusions concerning the application of our results to suspension bridges with multiple intermediate piers.

    Let (H,(,),) be a Hilbert space and consider a diagonal, self-adjoint and positive-definite operator A2:D(A2)HH, with eigenvalues 0<α1<<αj and eigenfunctions en, solutions of the problem

    (Aen,Av)=αn(en,v)vD(A).

    The sequence (en)n1 is a complete orthonormal system of H. For our convenience, we preferred to use A2 instead of A to build the functional framework of the problem. The operator A2 defines a family of Hilbert spaces Hσ=D(Aσ/2) with σ0, endowed with the norms σ induced by the scalar products

    u,vHσ(u,v)σ:=(Aσ2u,Aσ2v)=n=1ασ/2nunvn,uσ:=(u,u)σ (2.1)

    where un=(u,en) and vn=(v,en). In particular, 0=. In the context of this work, we consider the cases when σ[2,2], where for negative s the space Hs is defined as the dual of Hs. Throughout the paper, we denote by , the duality product of H2. It possible to verify that HρHσ densely whenever 0σρ and that

    uHρ,0σ<ρuραρσ41uσ. (2.2)

    In this framework, for any family of indices J={j1,,jn}, we define the projection

    PJ:Hej1,,ejnu=h=1uhehnr=1ujrejr.

    In particular, we denote by PN and QN:=IPN the orthogonal projections onto e1,eN and onto eN+1, respectively. In addition, for any kN we introduce the projection k onto the orthogonal complement of ek given by

    k:=IPkQk1:Hek.

    Since A is a diagonal operator, we remark that

    s[0,2],M={m1,,mn},AsPM=PMAsandAsQM=QMAs. (2.3)

    Moreover, if u=QNu for some NN, then the estimate (2.2) can be improved by

    uHρ,0σ<ρuραρσ4N+1uσ. (2.4)

    By using the notation in (2.1), problem (1.1) may be rewritten as

    utt+δut+A2u+u2θAθu=ginH×R+. (2.5)

    Let us make clear what is meant by weak solution of (2.5):

    Definition 2.1. Assume that

    gC0b(R+,H):=C0(R+,H)L(R+,H). (2.6)

    A weak solution of (2.5) is a function

    uC0(R+,H2)C1(R+,H)C2(R+,H2)

    such that

    utt,φ+δ(ut,φ)+(u,φ)2+u2θ(u,φ)θ=(g,φ)φH2.

    We remark that by this definition it follows that u(0)=u0H2 and ut(0)=u1H. Existence and uniqueness of weak solutions follows from an immediate adaptation of the result in [33,Theorem 2.1] (see Theorem 3.1).

    First, we prove that if the forcing term if finite-dimensional, i.e. if g=PNg for some NN, then any weak solution of (2.5) is asymptotically finite-dimensional. Actually, we guarantee the validity of the result for a more general family of forcing terms. We introduce the notion of exponentially Ndimensional forcing term.

    Definition 2.2. We say that gC0b(R+,H) is exponentially Ndimensional if there exists η>0 such that

    limt(QNg(t)+QNgt(t))eηt=0.

    In Section 4, we prove the following statement which describes the asymptotic behavior of the solution in the case when the forcing term is exponentially Ndimensional.

    Theorem 2.3. Assume (2.6) and let δ>0. If g is exponentially Ndimensional, there exists MN and ˜η>0, both depending on δ, lim suptg(t), θ, N, η and α1, i.e., the first eigenvalue of A2, such that

    limt(QMu(t)22+QMut(t)2)e˜ηt=0,

    where u is a weak solution of (2.5).

    Motivated by physical arguments (see Section 7), we now consider a "separated variables" forcing term such as g(t)=gf(t), where gH and fC0b(R+,R).

    Let us consider a weak solution u of (2.5). Numerical simulations show that for some j we have lim supt|(u(t),ej)|lim suptu(t), that is, we have that the asymptotic amplitude of some modes of u seems to be negligible with respect to the overall dynamics (see Figure 3). Hence, we expect to be able to neglect such modes both from the forcing term g and the solution u, thus reducing the numerical complexity of the model. Therefore, for any finite family of indices J={j1,,jm}, we consider the finite-dimensional approximation of (2.5) given by

    vtt+δvt+A2v+v2θAθv=PJg. (2.7)

    We remark that in virtue of Theorem 2.3, any solution of (2.7) is exponentially finite-dimensional. We prove that under suitable smallness conditions on the forcing term, for an appropriate choice of J, (2.7) is a good approximation of (2.5), i.e., for any weak solution u of (2.5), the weak solution v of (2.7) provides a good exponentially finite-dimensional approximation of u. More precisely, in Section 5 we prove the following theorem:

    Theorem 2.4. Assume δ>0 and g(t)=gf(t) with gH and fC0b(R+,R). There exists ˉg=ˉg(α1,δ,θ)>0 such that, if

    g:=lim suptg(t)<ˉg,

    then for every ε>0 there exists a finite family of indices J={j1,jm} depending on α1, δ, g and ε such that

    lim supt(u(t)v(t)22+ut(t)vt(t)2)ε

    where u is a weak solution of (2.5) and v is a weak solution of (2.7).

    Moreover, if g is exponentially Ndimensional, then there exist MN and ˜η>0, both depending on α1, δ, lim suptg(t), θ, N and η, such that, if J={1,,M}, then

    limt(PMu(t)v(t)22+PMut(t)vt(t)2)e˜ηt=0.

    In Section 6 we further restrict ourselves to the case when the forcing term is sinusoidal in time and, for the sake of simplicity, we focus on the case when θ=0, i.e., we study the problem

    utt+δut+A2u+u2u=gsin(ωt). (2.8)

    For g small enough, Theorem 2.4 states that if we replace g with PMg, we commit an error arbitrarily small as M grows. This suggests to consider the case when g=PMg for some MN. Let v be a solution of

    vtt+δvt+A2v+v2v=kgsin(ωt). (2.9)

    Let us now estimate the distance between u and v. The following theorem holds:

    Theorem 2.5. Assume δ>0 and let g(t)=gsin(ωt) with g=PMg for some MN. There exists ˉg>0 depending on δ, ω and αj with j=1,M, such that, if g<ˉg, then, for any k{1,,M} and for any u and v weak solutions of (2.5) and (2.9),

    lim supt(ku(t)v(t)22+kut(t)vt(t)2)C(g,ek)4((αkω2)2+δ2ω2)2,

    where C=C(α1,,αM,g,δ,ω)>0.

    The results involved in the proof of Theorem 2.5 are the most physically significant in the applications considered (see Section 7). In fact, Theorem 2.5 relies upon an estimate on the asymptotic amplitude of each mode, that allows us to study the distribution of the energy among the modes (see Figures 3 and 5) and to obtain a new bound on the asymptotic H2norm of u that improves the estimate given in [8,Lemma 22] (see Figure 2).

    Theorems 2.4 and 2.5 are not perturbation statements. Indeed, for any fixed δ>0, an explicit expression of the smallness conditions on g and g required by the statements of Theorems 2.4 and 2.5 is obtained in Sections 5 and 6 respectively. Since the term g models the action of the wind along the deck of the bridge, we physically interpret such smallness conditions on g as requirements on the aerodynamic load on the structure. In particular, the conditions of Theorems 2.4 and 2.5 are equivalent to require that the speed of the wind v is below a certain threshold ˉv. Moreover, we remark that such conditions can not be avoided since even in the ODE case large forcing terms lead to a chaotic dynamics [44,45] and the behavior of the solutions can be quite complicated, even where the forcing term is periodic in time [30,48].

    Our results are adaptable to more general frameworks. In particular, exploiting the abstract results of Haraux [37] and Chueshov [16], the cases with strong damping terms and with more general nonlinearities such as Aθut and M(u2θ)Aθ/2u with 0θ1 appear to be treatable. On the other hand, our results can not be immediately generalized to evolution equations with nonlinear nonlocal damping terms such as N(u21)g(ut), since the linear analysis on which the proof of Theorem 2.5 is based seems not to be easily extendable to such case.

    We notice that, if the initial states of (2.5) and (2.7) were close to each other, a uniform estimate on the distance in the phase space between the solutions of the approximated and the exact problem would be expected to hold for any t0. Unfortunately, we were not able to obtain such estimate and the techniques exploited in the proofs of Theorems 2.4 and 2.5 do not seem suitable to get this result.

    We start by recalling some basic properties concerning well-posedness and regularity of the solutions.

    Theorem 3.1. Let (2.6) hold. Then

    1). (Weak solutions) If u(0)=u0H2 and ut(0)=u1H, problem (2.5) admits a unique global weak solution such that

    uC(R+,H2)C1(R+,H)C2(R+,H2);

    2). (Regular solutions) If u(0)=u0H4 and ut(0)=u1H2, problem (2.5) admits a unique regular solution, that is, a unique global weak solution such that

    uC(R+,H4)C1(R+,H2)C2(R+,H);

    3). (Continuous dependence on initial data) Let (u0n,u1n) be any sequence with

    (u0n,u1n)(u0,u1)inH2×H,

    and let un(t) denote the weak solution of (2.5) with initial data un(0)=un and ut(0)=u1n. Then for every T>0 we have that

    (un(t),un,t(t))(u(t),ut(t))uniformlyinC0([0,T],H2×H).

    The proof follows from a standard applications of monotone operator theory with locally Lipschitz perturbations. We refer to [20,Theorem 1.5 and Proposition 1.15] and the references therein for a detailed discussion, that we decided to omit. For an alternative approach, see [33,Theorem 2.1] for the global existence and uniqueness of weak solutions and continuous dependence on initial data and [8,Theorem 5] for the global existence and uniqueness of regular solutions.

    We remark that in Theorem 3.1 we did not introduce the concept of strong or classical solution. This choice is motivated by the fact that in some applications such formulations are not possible, as in the case of the multiple intermediate piers model discussed in the introduction (see [29,Section 4] for a more detailed discussion).

    The following proposition gives some ultimate bounds on the Sobolev norms of u. Since the result comes from a straightforward generalization of the estimates proved in Section 7 of [8], we omit the proof.

    Proposition 3.2. Assume (2.6) and let u be a weak solution of (2.5). We introduce the quantities g:=lim suptg(t) and

    E:=g2max(2δ2,12α1),α:={δ/2ifδ2<4α1,δ/2δ2/4α1ifδ24α1.

    Then, the following estimates on u hold:

    lim suptu(t)24Eα21+4αθ1E+α1=:Φ0;lim suptu(t)2θ4E+2α2Φ0α2θ1+2(2E+α2Φ0)+α1θ/21=:Φθ;lim suptu(t)222E+α2Φ0=:Φ2;lim suptut(t)2minλ>01+λλ(2E+maxs[0,Φ0]((λ+1)α2α1s12s2))=:Φv.

    We now prove the continuous dependence of the solutions on the forcing term under suitable smallness conditions on the parameters of the problem.

    Proposition 3.3. Let u and v be weak solutions respectively of the problems

    utt+δut+A2u+u2θAθu=g1,vtt+δvt+A2v+v2θAθv=g2 (3.1)

    where g1,g2C0b(R+,H). Let Υμ:=lim supt(u(t)+v(t))/22μ with μ in [0,2]. There exists Fθ(α1,δ,Υθ,Υ2θ) such that, if Fθ<1 holds, then there exists C>0 depending on δ and g such that

    lim supt(u(t)v(t)22+ut(t)vt(t)2)Clim suptg1(t)g2(t). (3.2)

    Moreover, if there exists η>0 such that lim suptg1(t)g2(t)eηt=0, then there exists η1>0 such that

    limt(u(t)v(t)22+ut(t)vt(t)2)eη1t=0. (3.3)

    In particular, we can take

    Fθ:=2ΥθΥ2θαθ/41+Υθα(1θ)/21max(1δ,12α1). (3.4)

    Proof. The idea of the proof is standard but, for our purposes, it is mandatory to fully report it since we are interested in making the smallness conditions required from our results explicit.

    Let α>0. We define

    Λα:=12wt2+12w22+αδ2w2+116w4θ+α(wt,w)

    and let E be the quantity

    E:=12wt2+12w22+14w4θ.

    Remark that, by using the Cauchy-Schwarz inequality, the Young inequality and (2.2), we get

    Λα1+αε212wt2+αδ2w2+α1+α/ε212α1w22+116w4θC1E,
    Λα1αε222wt2+αδ2w2+α1α/ε222α1w22+116w4θC2E, (3.5)

    where C1 and C2 are positive numbers, obtainable for suitable choices of the values of α, ε1 and ε2. In particular, to get C2 we have to require

    1αε22>0,α1αε22>0.

    Hence, for every α such that α<α1 we can find ε2 such that (3.5) holds.

    We first consider u and v as regular solutions of the problems in (3.1). We define w:=vu and r:=g2g1. The function w is the regular solution of the problem

    wtt+δwt+A2w+v2θAθvu2θAθu=r. (3.6)

    We remark that, if ξ:=(u+v)/2, we have

    v(t)2θAθv(t)u(t)2θAθu(t)=2(ξ(t),w)θAθξ(t)+ξ(t)2θAθw+14w2θAθw. (3.7)

    From the definition of Λα, by using (3.6) and (3.7), since u and v are regular solutions we get

    ˙Λα+(δα)wt2+αw22+2(ξ,w)θ(Aθξ,wt)+ξ2θ(Aθw,wt)++2α|(ξ,w)θ|2+αξ2θw2θ+α4w4θ=(r,wt+αw). (3.8)

    Let Cμ=supt0ξ(t)2μ for any μ[0,2]. For a suitable choice of α, by using Cauchy-Schwarz and Young inequality we have that for some positive constants ˉα and ˜α

    (δα)wt2+αw22+2(ξ,w)θ(Aθξ,wt)+ξ2θ(Aθw,wt)+2α|(ξ,w)θ|2++αξ2θw2θ+α4w4θ(δα)wt2++αw222ξθwθξ2θwtξ2θw2θwt+α4w4θ(δα2CθC2θαθ/41+Cθ2α(1θ)/21)wt2+(α2CθC2θαθ/41+Cθ2α(1θ)/21)w22++α4w4θˉαE˜αΛα. (3.9)

    In particular, we choose the parameter α so that

    {δα2CθC2θαθ/41+Cθ2α(1θ)/21>0α2CθC2θαθ/41+Cθ2α(1θ)/21>0,{δ>α+2CθC2θαθ/41+Cθ2α(1θ)/21α>2CθC2θαθ/41+Cθ2α(1θ)/21.

    Hence, since α<α1, if

    {δ>2CθC2θαθ/41+Cθα(1θ)/21,α1>2CθC2θαθ/41+Cθ2α(1θ)/21

    we can find values of α such that (3.9) holds. Therefore we can find α such that (3.9) is satisfied if

    2CθC2θαθ/41+Cθα(1θ)/21max(1δ,12α1)<1. (3.10)

    Now, for some positive ˜α and ˜C we get, from (3.8) and (3.9),

    ˙Λα+˜αΛα(r,wt+αw)˜Cr=:˜f(t). (3.11)

    By defining

    Mα(t)=Λα(t)tt0˜f(s)e˜α(st)ds,

    from (3.11) we obtain

    ˙Mα(t)+˜αMα(t)0.

    Hence, from the Gronwall inequality and from the fact that for any ε>0 there exists t0>0 such that |˜f(s)|˜C(ε+lim suptr(t)) for any st0, we get

    Λα(t)Λα(t0)e˜α(tt0)+tt0˜f(s)e˜α(st)dsΛα(t0)e˜α(tt0)+˜C(ε+lim suptr(t))e˜αte˜αte˜αt0˜α,tt0. (3.12)

    Since we can take ε arbitrarily small as t0 goes to infinity, from (3.12) we infer that there exists C>0 such that

    lim suptΛα(t)Clim suptr(t). (3.13)

    Moreover, if there exists η>0 such that lim suptr(t)eηt=0, then (3.12) yields that there exists η1>0 such that

    limtΛα(t)eη1t=0. (3.14)

    From (3.5), there exists a positive constant C2 such that Λα(t)C2E(t). Therefore, (3.13) and (3.14) imply (3.2) and (3.3) respectively.

    We remark that

    lim suptξ(t)2μ=Υμ.

    Hence, we can take Cμ=Υμ. Therefore, from (3.10), we get that if

    2ΥθΥ2θαθ/41+Υθα(1θ)/21max(1δ,12α1)<1,

    then the thesis holds for regular solution u and v.

    The same conclusions hold for u and v weak solutions of the problems in (3.1) by using a standard density argument. Indeed, since H4 is dense in H2 and H2 is dense in H, setting (u(0)=u0,ut(0)=u1) and (v(0)=v0,vt(0)=v1), there exists two sequences (u0n,u1n) and (v0n,v1n) in H4×H2 such that

    (u0n,u1n)(u0,u1)and(v0n,v1n)(v0,v1)inH2×H.

    Hence, from Theorem 3.1 we have the two sequences of regular solutions un and vn with (un(0)=u0n,un,t(0)=u1n) and (vn(0)=v0n,vn,t(0)=v1n) such that, for any T>0,

    (un,un,t)(u,ut),(vn,vn,t)(v,vt)uniformlyinC([0,T],H2×H).

    Therefore, since all the calculations hold for un and vn (and the difference wn:=unvn), we get the thesis for the weak solutions u and v passing to the limit when n.

    In order to prove Theorem 2.3, we give a reformulation of Theorem 4.1 of [37] adapted to our framework.

    Proposition 3.4. Let (H,(,),||) be a Hilbert space and let A2 be a self-adjoint and strictly positive linear operator on H with dense domain D(A). We introduce the Hilbert space V:=D(A) endowed with the norm 2:=(A,A) and we identify the unbounded operator A2 with its extension in L(V,V). The duality pairing in V×V will be denoted in the same way as the inner product in H.

    We consider B(t)C1(R+,L(V,H)) such that for any vV

    0lim supt(B(t)v,v)λv2,lim supt(B(t)v,v)λv2

    for some positive numbers λ and λ.

    Let u be a bounded solution of

    utt+δut+(A2+B(t))u=g

    where δ>0, gC(R+,H) and limt|g(t)|ec0t=0 for some positive constant c0.

    If

    λδ<1

    then there exists c>0 such that

    limt(u(t)2+|ut(t)|2)ect=0.

    Proof. We proceed as in the proof of Theorem 4.1 of [37] and we define the quadratic form on V×H given by

    Φ(t)=12(|ut|2+u2)+δ2(u,ut)+δ24|u|2+12(B(t)u,u).

    For any fixed t0>0 we have, if tt0,

    Φt=12(B(t)u,u)δ2|ut|2δ2(B(t)u+A2u,u)+(g,ut+δ2u)12suptt0(B(t)u,u)δ2|ut|2δ2u2+Kec0t.

    for some positive constant K. Hence, for t0 large enough

    Φt(t)δ2|ut(t)|2δλ2u(t)2+Kec0t

    Therefore, if λ<δ we get, for some positive α,

    Φt(t)+αΦ(t)Kec0t

    for any tt0 and from Gronwall lemma we get the thesis.

    We recall a further stability result due to Haraux for an ODE related to our problem.

    Proposition 3.5. [Theorem 2.1 of [37]] Let λ,δ>0, aL(R+) with a(t)0 for any t0. Let xC2(R+) be a solution of

    ¨x+δ˙x+(λ+a(t))x=0. (3.15)

    Assume

    lim supta(t)<δmax(δ,2λ).

    There there are η1>0 and M>0 such that any bounded solution x of (3.15) satisfies

    x2(t)+˙x2(t)M[x2(s)+˙x2(s)]eη1(ts)

    for any st.

    With minimal effort, the same statement can be proven for x solving

    ¨x+δ˙x+(λ+a(t))x=˜g.

    where ˜gC(R+) satisfies limt˜g(t)eηt=0 for some η>0.

    Some preliminary results on the behavior of a damped and forced harmonic oscillator are useful in order to simplify the following study. In particular, we study the equation

    ¨y+δ˙y+λy=Ψ, (3.16)

    where we require Ψ to be antiperiodic. We recall that a function f:RR is said to be antiperiodic of antiperiod τ (i.e. τantiperiodic) if

    f(t+τ)=f(t),tR.

    Proposition 3.6. Let us consider ΨL2loc(R+) antiperiodic of anti-period π/ω. We suppose that λ>0 and δ>0. Then there exists an antiperiodic solution z of anti-period π/ω of (3.16) and we have that for some η>0, for any y(t) solution of (3.16),

    limt(|y(t)z(t)|+|˙y(t)˙z(t)|)eηt=0.

    Proof. Let us consider AωL2([0,π/ω]) the space of the locally square-integrable antiperiodic functions with anti-period π/ω, endowed with the standard L2 norm on the interval [0,π/ω]. The family {en=ω/πe(2n+1)iωt}nZ is an orthonormal basis of this space. Hence, we write

    Ψ(t)=ωπnZψje(2n+1)iωt.

    Setting

    z(t):=ωπnZψnω2(2n+1)2+λ+iδω(2n+1)e(2n+1)iωt,

    it is immediate to verify that z(t) is an antiperiodic solution of (3.16). The thesis now follows from the standard theory of ODEs. Indeed, any solution of (3.16) is given by the sum of z(t) with a general solution yg of the associated homogeneous equation

    ¨yg+δ˙yg+λyg=0,

    which is given by

    yg(t)=eδt/2f(t),

    with

    f(t):={Ssin(t24λδ2+φ),if4λ>δ2,St24λδ2cos(φ)+Ssin(φ),if4λ=δ2,Ssinh(t2δ24λ+φ),if4λ<δ2,

    where the arbitrary constants S and φ are dependent from the initial conditions. We notice that

    max(|f(t)|,|f(t)|)Ceμt,

    for some constants C>0 and 0μ<δ/2. Therefore, since y(t)=z(t)+yg(t), we get that for a suitable choice of η>0

    limt(|y(t)z(t)|+|˙y(t)˙z(t)|)eηt=limt(|f(t)|+|f(t)δ2f(t)|)e(ηδ/2)tδ+42Climte(η+μδ/2)t=0,

    which is the thesis.

    Proposition 3.7. Let us consider ΨL2loc(R+) antiperiodic of anti-period π/ω and let y(t) satisfy (3.16). We suppose λ,δ>0 and 2λδ. We introduce the quantities

    w±λ:=π2ω2(λδ22±δδ24λ),Ω2λ:=π42ω4(w+λwλ)(tan(w+λ2)w+λtan(wλ2)wλ)

    where, for any wC, w is the complex number z such that

    z2=wandz{ζ:(ζ)>0}{ζ:(ζ)=0and(ζ)0}.

    Then the following estimate holds

    lim supty(t)ΩλΨL([0,π/ω]). (3.17)

    Moreover, if ΨC2(R+), then

    lim supt˙y(t)Ωλ˙ΨL([0,π/ω]).

    Proof. From Proposition 3.6, Eq (3.16) admits an antiperiodic solution z(t) and any solution of y(t) of (3.16) converges exponentially to z(t), which yields that lim supty(t)=lim suptz(t). Hence, since from the antiperiodicity of z(t) we have that lim suptz(t)=z, in order to get the result it suffices to estimate the Lnorm of z(t). In the notation of Proposition 3.6, we have that

    z(t):=ωπnZψnω2(2n+1)2+λ+iδω(2n+1)e(2n+1)iωt,

    Then, if cn=(ω2(2n+1)2+λ)2+δ2ω2(2n+1)2, from Cauchy-Schwarz inequality we obtain

    |z(t)|ωπnZ|ψn|cnωπnZ|ψn|22n01c2n. (3.18)

    Moreover, if ΨC2(R+), we have

    |˙z(t)|ωπnZ|(2n+1)ωψn|cnωπnZ|(2n+1)ωψn|22n01c2n. (3.19)

    First, we remark that from Parseval's theorem

    nZ|ψn|2=ΨL2([0,π/ω])πωΨL([0,π/ω]),nZ|(2n+1)ωψn|2=˙ΨL2([0,π/ω])πω˙ΨL([0,π/ω]). (3.20)

    Then, to conclude the proof, we compute a closed form for the serie

    n01c2n=n01ω4(2n+1)4(2λδ2)(2n+1)2ω2+λ2. (3.21)

    We observe that (3.21) becomes

    n01c2n=n0π4(w+λwλ)ω4[1(2n+1)2π2w+λ1(2n+1)2π2wλ]. (3.22)

    We now recall that the Mittag-Leffler expansion for the cotangent function gives

    cot(w)=1w+n=12ww2π2n2.

    Some straightforward computations give

    12tan(w2)=12cot(w2)cot(w)=n=02w(2n+1)2π2w2.

    Thus, we can infer that

    n01(2n+1)2π2wλ=tan(wλ2)4wλ.

    Hence, from (3.22) we can conclude that

    n01c2n=π44ω4(w+λwλ)(tan(w+λ2)w+λtan(wλ2)wλ). (3.23)

    By using (3.20) and (3.23) in (3.18) and (3.19), we obtain the thesis.

    In [36,Theorem 2.1], a result similar to Proposition 3.7 is proven. In particular, the maximum value of lim supty(t) as the forcing term Ψ varies in the unitary ball of L(R) is determined. On the other hand, for any fixed antiperiodic forcing term Ψ in C2(R), in Proposition 3.7 we estimated lim supty(t) and lim supt˙y(t). As Figure 1 shows, Proposition 3.7 almost always gives a better estimate on lim supty(t).

    Figure 1.  Comparison between the estimates on the norm of y solution of (3.16) given by [36] (blue) and by (3.17) (black) with δ=1 and ω=3 as λ vary from 1 to 150 (left) and with δ=1 and λ=5 as ω vary from 1 to 15 (right). In red, we represented the norm of the antiperiodic solution of (3.16) with Ψ(t)=signum(sin(ωt)).

    The remainder of the paper is organized as follows. First, in Section 4 we apply the results of Subsection 3.2 in order to prove Theorem 2.3. In particular, we apply Proposition 3.4 to prove that for N large enough, if g is exponentially Ndimensional, then there exists ˉNN such that any solution u of (2.5) is exponentially ˉNdimensional (see Lemma 4.1). After that, fixed n>N, we study the asymptotic amplitude of un(t)=(u(t),en) for any u solution of (2.5) and in Lemma 4.2 we determine whether un(t) decays exponentially as t goes to infinity. In subsection 4.2 we exploit Lemma 4.1 and Lemma 4.2 in order to get Theorem 2.3. We remark that, even though the thesis of Theorem 2.3 follows from Lemma 4.1, Lemma 4.2 is necessary in order to improve the result of Lemma 4.1. More precisely, Lemma 4.2 provides an improvement of the smallest number MN obtained in Lemma 4.1 such that if g is exponentially Ndimensional then any solution u is exponentially Mdimensional.

    Next, by exploiting the continuous dependence of the solution from the forcing term, that is, Proposition 3.3, and Theorem 2.3, in Section 5 we give the proof of Theorem 2.4.

    In Section 5, by proceeding as in a result of Bonheure, Gazzola and Moreira dos Santos [8,Theorem 6], we show that (2.8) admits an antiperiodic solution p. In Lemma 6.2 we use Proposition 3.7 to estimate, for any nN, the asymptotic amplitude of pn(t):=(p(t),en). Such result yields an estimate on the Hsnorms of p (see Lemma 6.3) which we numerically verified to be better than the a-priori estimates obtained in [8] (see Figure 2). From Proposition 3.3, we have that under suitable smallness conditions on lim suptg(t), any solution u of (2.8) converges to p in the phase space norm. Hence, from Lemma 6.2 and Lemma 6.3, in Lemma 6.4 we get an estimate on the asymptotic amplitude of un(t)=(u(t),en) and on the Hsnorms of u for any u solution of (2.8). Finally, in Lemma 6.5, we exploit the previous results of Section 6 in order to get a results for finite-dimensional systems of ODEs and in Subsection 2.5 we apply Lemma 6.5 and Lemma 6.4 to get Theorem 2.5.

    Figure 2.  Comparison between the general estimate on lim suptu(t)2 (blue) and the one obtained by using the antiperiodicity of the forcing term (red).

    We now apply the results of the previous section to our framework in order to prepare the proof of Theorem 2.3.

    Lemma 4.1. Let u be a weak solution of (2.5). Let g be exponentially Ndimensional. If there exists ˉNN such that

    lim supt(1α1θ1u(t)22+ut(t)2)<2δα(2θ)/2ˉN+1

    then there exists ˜η>0 such that

    lim supt(QˉNu(t)22+QˉNut(t)2)e˜ηt=0.

    Proof. Fix ˉNN and, for any s[0,2], let Υs:=lim suptu(t)2s. We introduce the operator-valued function B(t):=u(t)2θAθ. By using (2.3), we get that w=QˉNu solves

    wtt+δwt+(A2+B(t))w=QˉNg. (4.1)

    By using (2.4) we remark that for any vH2 such that QˉNv=v

    0lim supt(B(t)v,v)=lim suptu(t)2θv2θΥθα(2θ)/2ˉN+1v22,lim supt(B(t)v,v)=lim supt(ut(t),Aθu(t))v2θ12α(2θ)/2ˉN+1lim supt(1α1θ1u(t)22+ut(t)2)v22. (4.2)

    We introduce

    φ(t)=12(ut(t)2+Au(t)2)+δ2(u(t),ut(t))+δ24u(t)2.

    By applying Proposition 3.4 to (4.1), from (4.2) we get that if

    lim supt(1α1θ1u(t)22+ut(t)2)<2δα(2θ)/2ˉN+1,

    then φ(t)0 exponentially as t goes to infinity. This yields that there exists ˜η>0 such that

    limt(Aw(t)2+wt(t)2)e˜ηt=0.

    Therefore, since Aw2=w22, we get the thesis.

    We now apply Proposition 3.5 to the projection of (2.5) on the nth mode. The following lemma holds.

    Lemma 4.2. Let g be exponentially Ndimensional. For any weak solution u of (2.5), if

    nN+1suchthatlim suptu(t)2θ<δmax(2θδ1θ,2α(1θ)/2n), (4.3)

    then for any Mn there exists ˜η>0 such that for any nˉNM

    limt(|(u(t),eˉN)|2+|(ut(t),eˉN)|2)e˜ηt=0.

    Proof. Fixed nN+1, we consider the projection of u on the nth mode, i.e., un:=(u,en). The function un satisfies

    ¨un+δ˙un+(αn+u(t)2θαθ/2n)un=(g,en).

    Since nN+1, for some η>0, limt(g(t),en)eηt=0. Let us suppose that lim suptu(t)2θ<δmax(2θδ1θ,2α(1θ)/2n). Since

    max(2θδ1θ,2α(1θ)/2n)max(δαθ/2n,2α(1θ)/2n),

    we have that

    αθ/2nlim suptu(t)2θ<δmax(δ,2αn),

    which yields that, from Proposition 3.5,

    limt(|un(t)|2+|˙un(t)|2)eη1t=0.

    Since (αj)j is strictly increasing, max(2θδ1θ,2α(1θ)/2n) is an increasing sequence. Hence, if (4.3) holds, then for any ˉNn

    lim suptu(t)2θδmax(2θδ1θ,2α(1θ)/2ˉN),

    that implies that for any Mn there exists ˜η>0 such that for any nˉNM

    limt(|uˉN(t)|2+|˙uˉN(t)|2)e˜ηt=0,

    that is the thesis.

    Let g be exponentially Ndimensional and let u be a weak solution of (2.5). We recall that, from Proposition 3.2, we have

    lim suptu(t)2θ4E+2α2Φ0α2θ1+2(2E+α2Φ0)+α1θ/21=:Φθ;lim suptu(t)222E+α2Φ0=:Φ2;lim suptut(t)2minλ>01+λλ(2E+maxs[0,Φ0]((λ+1)α2α1s12s2))=:Φv. (4.4)

    We introduce the quantity ˉN defined as the smallest integer number greater than N such that

    1α1θ1Φ2+Φv<2δα(2θ)/2ˉN+1. (4.5)

    From (4.4), (4.5) implies

    lim supt(1α1θ1u(t)22+ut(t)2)<2δα(2θ)/2ˉN+1.

    Hence, from Lemma 4.1, if (4.5) holds then there exists η1>0 such that

    limt(QˉNu(t)22+QˉNut(t)2)eη1t=0.

    We introduce the set

    B:={nN:n[N,ˉN]andΦθ<δmax(2θδ1θ,2α(1θ)/2n+1)}

    and we define

    N_:={minBifB+ifB=.

    From Proposition 3.2 we have that lim suptu(t)2θΦθ. Hence, from Lemma 4.2, if N_+, there exists η2>0 such that

    limt(|(u(t),en+1)|2+|(˙u(t),en+1)|2)eη2t=0

    for any n[N_,ˉN]N, which yields

    limt(QN_PˉNu(t)22+QN_PˉNut(t)2)eη2t.

    Hence, if we set P:=I, Q:=0 and M:=min{N_,ˉN}, for some ˜η>0

    limt(QMu(t)22+QMut(t)2)e˜ηt=limt(QN_PˉNu(t)22+QN_PˉNut(t)2)e˜ηt++limt(QˉNu(t)22+QˉNut(t)2)e˜ηt=0.

    This concludes the proof of Theorem 2.3.

    Let us suppose that

    2ΦθΦ2α(θ2)/41+Φθ2α(1θ)/21max(1δ,1α1)<1, (5.1)

    where Φθ and Φ2 are defined in Proposition 3.2. Since Φθ and Φ2 depend on g and δ, we get that, for any fixed δ, (4.5) translates into Fθ(α1,δ,g)<1 for some Fθ. Therefore, for any fixed δ>0, there exists ˉg>0 such that if g<ˉg, then (4.5) holds. We remark that, since the term g models the action of the wind along the deck of the bridge, we physically interpret (4.5) as a requirement on the load exerted on the structure by the wind. In particular, since ˉg in engineering applications (see [23]) is proportional to the speed of the wind v, the relation (4.5) is equivalent to require that v<ˉv for some ˉv>0.

    Let u be a weak solution of (2.5) and for any J={j1,,jm} let vJ be a weak solution of the problem

    vJtt+δvJt+A2vJ+vJ2θAθvJ=PJg.

    We introduce the quantities Υμ=lim supt(u(t)+vJ(t))/22μ, where μ[0,2]. From Proposition 3.3 with g1=PJg and g2=g=PJg+QJg, there exists a function Fθ=Fθ(α1,δ,Υθ,Υ2θ), given by (3.4), such that if Fθ<1 then there exists a constant C>0 such that

    lim supt(u(t)vJ(t)22+ut(t)vJt(t)2)Clim suptQJg(t). (5.2)

    Since g=gf(t), for a suitable choice of J, we have that Clim suptQJg(t)<ε. Hence we can conclude that, for a suitable choice of the family J, (5.2) gives

    lim supt(u(t)vJ(t)22+ut(t)vJt(t)2)ε. (5.3)

    From Proposition 3.2 and (2.2), we have that ΥθΦθ and Υ2θ<αθ11Φ2. Hence, Fθ<1 is implied by (4.5). Therefore, fixed δ, if g<ˉg for some positive constant ˉg, where ˉg does not depend by J, then (5.3) holds. This proves the first part of Theorem 2.4.

    Let now g be exponentially N-dimensional and let MN be obtained from Theorem 2.3, i.e., let MN be such that for some η>0

    limt(QMu(t)22+QMut(t)2)eηt=0. (5.4)

    Let u and v be, respectively, weak solutions of (2.5) and

    vtt+δvt+A2v+v2θAθv=PMg.

    We remark that u is solution of the following problem

    utt+δut+A2u+u2θAθu=g=PMg+QMg.

    Since we supposed g to be exponentially N-dimensional and MN, there exists η>0 such that

    limtPMg(t)+QMg(t)PMg(t)eηt=limtQMg(t)eηt=0.

    Therefore, from Proposition 3.3 with g1=PMg and g2=g=PMg+QMg we have that, fixed δ, if g is sufficiently small, then there exists η1>0 such that

    limt(u(t)v(t)22+ut(t)vt(t)2)eη1t=0.

    Since v=PMv, from (5.4) we get that for some ˜η>0

    limt(PMu(t)v(t)22+PMut(t)vt(t)2)e˜ηt=0.

    This concludes the proof of Theorem 2.4.

    In Theorem 2.5, we restrict ourselves to the case when the forcing term is antiperiodic in time due to the engineering interest of this case (see Section 7). Moreover, for the sake of simplicity, we consider the case θ=0. The antiperiodicity of the forcing term allows us to provide some more information about the solution of (2.8). In particular, proceeding as in Theorem 6 of [8], where the result was proven in the periodic framework, by using Proposition 3.6, we obtain the following statement:

    Proposition 6.1. If g(t) is a continuous antiperiodic function of anti-period τ, then there exists a solution of (2.5) antiperiodic of anti-period τ.

    Proof. The proof proceeds as in [8,Theorem 6]. First, we fix n1 and we prove the existence of a τantiperiodic solution for the problem

    utt+δut+A2u+u2u=Png. (6.1)

    Hence, we seek a τantiperiodic solution un in the form

    un(x,t):=nk=1hnk(t)ek(x).

    We consider the spaces C2τ(R) and C0τ(R) of C2 and C2 τantiperiodic functions and in the same notations of [8,Theorem 6] we have that (6.1) is equivalent to

    Ln(h(t))+Gn(h(t))=g(t),

    where h:=(hn1,,hnn), g:=(g1,,gn), Ln is a diagonal operator such that

    Lkn(h):=¨hk+δ˙hk+αkhk

    and

    Gn(h):=14nj,k=1h2jh2k.

    We observe that for any q(C0τ(R))n from Proposition 3.6 there exists a unique h(C2τ(R))n such that Ln(h)=q. Thanks to the compact embedding (C2τ(R))n(C0τ(R))n, we have that the nonlinear map Γn:(C0τ(R))n×[0,1](C0τ(R))n defined by

    Γn(h,ν)=L1n(gνGn(h)),(h,ν)(C0τ(R))n×[0,1]

    is compact. Moreover, from Proposition 3.2 we have that there exists Hn>0 (independent of ν) such that if h(C0τ(R))n solves h=Γn(h,ν), then

    h(C0τ(R))nHn.

    Hence, since the equation h=Γn(h,0) from Proposition 3.6 admits a unique τantiperiodic solution, the Leray-Schauder principle ensures the existence of a solution h(C0τ(R))n of h=Γn(h,1). This proves the existence of a τantiperiodic solution of (6.1). The proof the result follows from the existence of a τantiperiodic solution of (6.1) exactly as in [8,Theorem 6] by showing that the sequence (un) converges to a τantiperiodic solution u of (2.8).

    In this section we use the quantities

    w±j:=π2ω2(αjδ22±δδ24αj),Ω2j:=π42ω4(w+jwj)(tan(w+j2)w+jtan(wj2)wj) (6.2)

    obtained by replacing λ by αj in Proposition 3.7.

    We now apply Proposition 3.7 in order to get an estimate on the jth mode of the antiperiodic solution p of (2.8), which we proved to exist in Proposition 6.1. In the following, whenever a real-valued function f(t) will be antiperiodic, we will write interchangeably lim suptf(t) and f.

    Lemma 6.2. Let p be an antiperiodic solution of (2.8). If

    maxjΩjlim suptp(t)2<1 (6.3)

    where Ωj is defined in (6.2), then, if we set Υ0:=lim suptp(t)2 and Υv:=lim suptpt(t)2,

    gj(1+Υ0Ωj)(αjω2)2+δ2ω2lim supt|pj(t)|gj(1Υ0Ωj)(αjω2)2+δ2ω2,(ω(1Υ0Ωj)2Υ0ΥvΩj)gj(1(Υ0Ωj)2)(αjω2)2+δ2ω2lim supt|˙pj(t)|(ω(1Υ0Ωj)+2Υ0ΥvΩj)gj(1Υ0Ωj)2(αjω2)2+δ2ω2,

    where pj:=(p,ej) and gj:=lim supt(g(t),ej)=(g,ej).

    Proof. We study the jth component of the problem (2.8), namely

    ¨pj+δ˙pj+αjpj+p2pj=gjsin(ωt). (6.4)

    We consider the antiperiodic solution v of the problem

    ¨v+δ˙v+αjv=gjsin(ωt). (6.5)

    It is possible to verify that the general solution of (6.5) is given by

    v(t)=gj(αjω2)2+δ2ω2sin(ωt+arctanδωω2αj)+Seδt/2sin(t24αjδ2+φ),

    where the constants S and φ are determined by the initial data of (6.5). Hence, it follows that, for any choice of the initial data of (6.5),

    lim suptv(t)=gj(αjω2)2+δ2ω2,lim supt˙v(t)=ωgj(αjω2)2+δ2ω2. (6.6)

    If we subtract (6.5) from (6.4), if w:=pjv we get

    ¨w+δ˙w+αjw=p2pj.

    Hence, from Proposition 3.7 we get, if p(0)j:=lim suptpj(t), p(1)j:=lim supt˙pj(t), Υ0:= lim suptp(t)2 and Υv:=lim suptpt(t)2,

    lim supt|w(t)|Υ0Ωjp(0)j,lim supt|˙w(t)|Ωj2(p(t),pt(t))pj(t)+p(t)2˙pj(t)L(0,π/ω)2Υ0ΥvΩjp(0)j+Υ0Ωjp(1)j. (6.7)

    Since p and v are both antiperiodic, w is antiperiodic and (6.7) gives

    |vpj|wΥ0Ωjp(0)j,|˙v˙pj|˙w2Υ0ΥvΩjp(0)j+Υ0Ωjp(1)j.

    We get then

    lim suptv(t)Υ0Ωjp(0)jp(0)jlim suptv(t)+Υ0Ωjp(0)j,lim supt˙v(t)2Υ0ΥvΩjp(0)jΥ0Ωjp(1)jp(1)jlim supt˙v(t)+Υ0Ωjp(1)j+2Υ0ΥvΩjp(0)j.

    Hence, from (6.6) we get, since hypothesis (6.3) holds,

    gj(1+Υ0Ωj)(αjω2)2+δ2ω2p(0)jgj(1Υ0Ωj)(αjω2)2+δ2ω2,

    which yields

    (ω(1Υ0Ωj)2Υ0ΥvΩj)gj(1(Υ0Ωj)2)(αjω2)2+δ2ω2p(1)j(ω(1Υ0Ωj)+2Υ0ΥvΩj)gj(1Υ0Ωj)2(αjω2)2+δ2ω2

    that is the thesis.

    We now apply the results of Lemma 6.2 in order to get an estimate on the Hnorm and H2norm of an antiperiodic solution p of (2.8).

    Lemma 6.3. Let p be an antiperiodic solution of (2.8). Let us suppose that

    maxjΩjΦ0<1,

    where Φ0 is defined in Proposition 3.2. Then the following estimates hold:

    lim suptp(t)2j=1g2j(1Φ0Ωj)2((αjω2)2+δ2ω2)=:φ<, (6.8)
    lim suptpt(t)2j=1(ω(1Φ0Ωj)+2Φ0ΦvΩj)2g2j(1Φ0Ωj)4((αjω2)2+δ2ω2)=:φv<, (6.9)
    lim suptp(t)22j=1αjg2j(1Φ0Ωj)2((αjω2)2+δ2ω2)=:φ2<. (6.10)

    Proof. We prove (6.10) only, since the proofs of (6.8) and (6.9) are completely analogous. From Lemma 6.2, by using that from Proposition 3.2 Υ0:=lim suptp(t)2Φ0,

    lim suptp(t)22j=1αjpj2j=1αjg2j(1Φ0Ωj)2((αjω2)2+δ2ω2).

    We recall that the sequence (αj)j is divergent. Therefore, for j large enough, wj=¯w+j and |w+jwj|=2π2δαjδ2/4/ω2π2δαj/ω2. Hence

    |Ω2j|π2δω2αj|(tan(w+j2)w+j)|π2δω2αj|tan(w+j2)||w+j|.

    We remark that

    |tan(a+ib)|sin2(2a)+sinh2(2b)(cos(2a)+cosh(2b))2.

    Moreover, from the definition of w+j (see (6.2)), we have that (w+j)+. Hence, we conclude that limj|tan(w+j/2)|=1 and consequently

    limtΩj=0.

    Then, since limjαj=+ and maxjΩjΦ0<1, we have that, for some positive constant C, for any jN

    αj(1Φ0Ωj)2((αjω2)2+δ2ω2)<C.

    Therefore, by using that

    j=1g2j=g2<,

    we get that

    j=1αjg2j(1Φ0Ωj)2((αjω2)2+δ2ω2)j=1Cg2j=Cg2<,

    that is the thesis.

    We observe that, from Proposition 3.3, any solution u of (2.8) exponentially converges to p under suitable smallness conditions on g. Hence, Lemma 6.2 and Lemma 6.3 hold for any weak solution u of (2.8). More precisely, the following lemma holds.

    Lemma 6.4. Let u be a weak solution of (2.8). If

    maxjΩjΦ0<1,F(ξ)<1,

    where F(ξ)=3ξmax(1/δ,1/(2α1))/α1 and ξ:=((Φ0+φ)/2)2, then

    lim suptu(t)2φ,lim suptu(t)22φ2,lim suptut(t)2φv,

    and

    gj(1+φΩj)(αjω2)2+δ2ω2lim supt|(u(t),ej)|gj(1φΩj)(αjω2)2+δ2ω2,(ω(1φΩj)2φφvΩj)gj(1(φΩj)2)(αjω2)2+δ2ω2lim supt|(ut(t),ej)|(ω(1φΩj)+2φφvΩj)gj(1φΩj)2(αjω2)2+δ2ω2,

    where φ,φv and φ2 are defined in (6.8), (6.9) and (6.10) respectively.

    Proof. Let p be an antiperiodic solution of (2.8). We define w=pu. The function w solves

    wtt+δwt+A2w+p2pu2u=0.

    We proceed as in Proposition 3.3 and we get that if

    F(lim suptξ(t)2)<1

    where ξ=(u+p)/2, then

    limt(u(t)p(t)22+ut(t)pt(t)2)=0. (6.11)

    Since

    lim suptξ(t)lim suptu(t)+lim suptp(t)2Φ0+φ2,

    from the monotonicity of F we get that F(ξ)<1 implies (6.11). Hence, the thesis follows from Lemma 6.2 and Lemma 6.3.

    Let us consider the finite-dimensional problem

    ¨x_+δ˙x_+Λx_+x_2x_=g_(t) (6.12)

    where x_(t)=(x1(t),,xn(t))Rn, g_(t)=(g1(t),,gn(t)), Λ=diag(αj)nj=1 and is the Euclidean norm in Rn. This problem is a finite-dimensional approximation of (2.8).

    Here, we estimate how much the evolution of the system changes as we eliminate a single mode from the dynamics. For the sake of simplicity, in the following we consider the case when the higher mode is the one we choose to neglect. We observe that

    Pn1¨x_+δ˙x_+Λn1Pn1x_+Pn1x_2Pn1x_+x2nPn1x_=Pn1g_(t) (6.13)

    where Pn1(a1,,an)=(a1,,an1), Λn1=diag(αj)n1j=1. We consider now the function y_(t), solution of

    ¨y_+δ˙y_+Λn1y_+(y_,0)2y_=Pn1g_(t) (6.14)

    At this point, the question is reduced to estimate the (asymptotic) distance between the solution x_ of (6.12) and the solution y_ of (6.14). To this end, with a slight abuse of notations, we introduce the Rnnorms 1 and 2 defined by x_1=|x1|++|xn| and x_2=α1|x1|2++αn|xn|2. We remark that the result is completely independent of the choice of the mode neglected. The following lemma holds.

    Lemma 6.5. Let x_ and y_ be solutions of Eqs (6.12) and (6.14) respectively. Let g=g_sin(ωt) with g_Rn and we suppose that F(ξ)<1, where ξ is defined in Lemma 6.4 and F(ξ)=3ξmax(1/δ,1/(2α1))/α1. Moreover, we suppose that

    maxjΩjΦ0<1,maxjΩjφ<1.

    Then there exists a function S of the parameters of the problem such that if S<1 then we have that

    lim suptPn1x_(t)y_(t)2C(χ_)χ2n,lim suptPn1˙x_(t)˙y_(t)C1(χ_,χ_v)χ2n+C2(χ_,χ_v)χn,vχn

    where χ_=(χ1,χn), χj:=lim suptmax(|xj(t)|,|yj(t)|), χ_v=(χ1,v,χn,v) and χj,v:=lim suptmax(|˙xj(t)|,|˙yj(t)|).

    Proof. First, we remark that as in Lemma 6.4, since F(ξ)<1, we have that there exist two antiperiodic functions p1_C2(R+,Rn) and p2_C2(R+,Rn1) such that

    limtx_(t)p1_(t)22+˙x_(t)˙p1_(t)2=0,limty_(t)p2_(t)22+˙y_(t)˙p2_(t)2=0.

    Therefore, since we are interested in the asymptotic behavior of our system, we can restrict ourselves to the case when x_ and y_ are both antiperiodic without loss of generality.

    Let us consider the difference between Eqs (6.13) and (6.14). If we set w_:=Pn1x_ and z_:=w_y_, we get

    ¨z_+δ˙z_+Λn1z_=Ψ_

    where Ψ_=x2nw_(w_2y_2)y_w_2z_ and for the sake of simplicity, abusing the notations, we wrote w_ and y_ instead of (w_,0) and (y_,0) respectively.

    We focus on one component, say j, in order to treat only scalar quantities. Hence, we consider the equation

    ¨zj+δ˙zj+αjzj=Ψj (6.15)

    where Ψj=x2nxj(w_2y_2)yjw_2zj=x2nxj(w_y_,w_+y_)yjw_2zj. The fact that x_ and y_ are antiperiodic implies that Ψ_ is antiperiodic too. Hence, we can apply Proposition 3.7 to (6.15) and, if we introduce the quantities

    φ:=maxt0max(x_(t)2,y_(t)2),φv:=maxt0max(˙x_(t)2,˙y_(t)2),χj:=max(xj,yj),χj,v:=max(˙xj,˙yj)forj=1,,n,

    then, set Z:=maxt0z_(t), we have

    zjΩjΨjΩj(χ2nχj+2φχjZ+φzj).

    Therefore, set Zj:=zj and Cj:=Ωjφ, by requiring that Cj<1 for any j=1,n we get

    ZjCjχj1Cj(χ2n+2φZφ). (6.16)

    We define the quantity

    S:=n1j=12Cjχj(1Cj)φ

    and we suppose S<1.

    We remark that for any x_Rn, x_x_1:=|x1|+|xn| and, for any bounded function f_:RRn, suptf_(t)1f1++fn. Hence we have that Zn1j=1Zj. Therefore, by summing (6.16) over j and solving in Z we get

    ZS1Sχ2n2φ. (6.17)

    Next, we remark that for any bounded function f_:RRn we have that suptf_(t)2α1f1++αnfn. Hence Z2:=maxt0z_(t)2n1j=1αjZj and from (6.16) and (6.17) it follows that

    Z2n1j=1αjZjn1j=1Cjχjαj1Cj(χ2n+2φZφ)1φ(1S)n1j=1Cjχjαj1Cjχ2n. (6.18)

    In particular, from (6.17) and (6.18) we conclude that there exist two positive constants b and c such that

    Zbχ2n,Z2cχ2n. (6.19)

    Moreover, from (6.19) and (6.16), there exist constants aj such that

    Zjajχ2nforanyj=1,,n1. (6.20)

    We now define Z(1)j:=˙zj and Z(1):=maxt0˙z_(t). By applying Proposition 3.7 to (6.15) we get

    Z(1)j=lim supt|˙zj(t)|Ωjlim supt|˙Ψj(t)|. (6.21)

    Since w_2y_2=(w_+y_,w_y_)=(w_+y_,z_), we have

    ˙Ψj=2xn˙xnxjx2n˙xj(˙w_+˙y_,z_)yj+(w_+y_,˙z_)yj(w_+y_,z_)˙yj2(w_,˙w_)zjw_2˙zj. (6.22)

    Therefore from (6.22) and (6.21) we get

    Z(1)jΩj(2χnχn,vχj+χ2nχj,v+2φvχjZ+2φχjZ(1)+2φχj,vZ+2φvφZj+φZ(1)j).

    Hence, by using (6.19) and (6.20), if Lj:=χj,v+2φvφaj+2(φχj,v+φvχj)b and Cj is defined as before, then

    Z(1)jCj1Cj2χnχn,vχj+Ljχ2n+2φχjZ(1)φ.

    By reasoning as before we conclude that, if S<1, then

    Z(1)11S(Sφχnχn,v+Lχ2n)

    where L is a suitable constant.

    We are now able to estimate the asymptotic distance between x_ and y_, since

    lim suptnx_(t)y_(t)2cχ2n,lim suptn˙x_(t)˙y_(t)S(1S)φχnχn,v+L1Sχ2n. (6.23)

    We remark that, since we can estimate φ and φv in function of χ_ and χ_v, S and L are dependent by χ1,χn and χv,1,χv,n only. Therefore, from (6.23) we get the thesis.

    Since g=PMg, from Lemma 6.4 we get that, if F(ξ)<1,

    limt|(u(t),ej)|=0,limt|(ut(t),ej)|=0forj>M.

    Therefore, we can rewrite (2.8) and (2.9) as finite-dimensional dynamical systems of the form (6.12) and (6.14) respectively.

    We introduce the quantities

    χj:=lim supt|(u(t),ej)|,χj,v:=lim supt|(ut(t),ej)|forjM.

    From Lemma 6.5, we have that if ΩjΦ0<1, Cj=Ωjφ<1 for any jM and

    S=Mj=12Cjχj(1Cj)φ<1

    where Φ0 and φ are defined in Proposition 3.2 and in Lemma 6.3, then

    lim suptku(t)v(t)21φ(1S)Mj=1Cjχjαj1Cjχ2k,lim suptkut(t)vt(t)S(1S)φχkχk,v+L1Sχ2k, (6.24)

    where L is obtained in the proof of Lemma 6.5. Fixed δ, we recall that S and L are constants depending on χ1,χn and χv,1,χv,n. Hence, since from Lemma 6.4 we have that

    χjgj(1φΩj)(αjω2)2+δ2ω2,χv,j(ω(1φΩj)+2φφvΩj)gj(1φΩj)2(αjω2)2+δ2ω2,

    from (6.24) we obtain that

    lim supt(ku(t)v(t)22+ku(t)v(t)2)Cg4k((αkω2)2+δ2ω2)2,

    where C is a constant depending on A2, g and ω, that is the thesis.

    In this section we show how the analysis performed in this paper can be useful in order to get some more information about the stability of real world structures such as suspension bridges.

    While in the first part of the paper (Theorem 2.3 and Theorem 2.4) we study the general case given by (2.5), in the second part (Theorem 2.5) we focus in particular on the case when θ=0 and

    g=gsin(ωt).

    In particular, taking H=L2(I) with I=[π,π], A=xx and D(A)={vH2(I)H10(I):v(π)=v(π)=v(aπ)=v(bπ)=0} for a,b(0,1), the results of Section 6 apply to the system

    {utt+δut+uxxxx+u2L2(I)u=g(x)sin(ωt)t0,xIu(0)=u0H2(I)H10(I),ut(0)=u1L2(I)u(π,t)=u(πb,t)=u(πa,t)=u(π,t)=0,t0. (7.1)

    This choice of the forcing term comes from the fact that, in engineering literature (see [35]), the load due to the vortex shedding of the wind along the structure of the bridge is usually modeled in this way with g(x)gR. The coefficient g depends on the wind speed and on the geometry of the structure and ω is the frequency at which vortex shedding occurs. More precisely, we have that in engineering applications g(x,t)=W2sin(ωt), where W is the scalar velocity of the wind blowing on the deck of the bridge and ω can be expressed in terms of the structural constants of the bridge and the aerodynamic parameters of the air. We refer to the European Eurocode [23] (see also [8]) for a more detailed discussion.

    The peculiar expression of the forcing term allows us to improve the estimate on the asymptotic H2norm of the solution of (7.1) that one is able to obtain with no other information on g than the value of lim suptg(t). A comparison between the general estimate on lim suptu2 (see Proposition 3.2) obtained by using the methods of [8,Lemma 22] and the one obtained by using the antiperiodicity of the forcing term (see Lemma 6.4) is given in Figure 2. The data considered are a=b=14/25, δ=1.5, and ω=20. The maximum value of g considered represents the largest value of g such that Lemma 6.4 can be applied.

    The improvement in the estimates on the asymptotic H2norm is obtained by using also ultimate bounds of the asymptotic amplitude of each mode. We represent in Figure 3 a comparison between these estimates, obtained in Lemma 6.4, and a numerical estimate on the asymptotic amplitude of each of the first 20 modes. Fixed δ=1.5 and g=1.5, we considered the cases when ω=5 (left) and ω=10 (right). We considered different positions of the piers, namely we chose a=b=14/25 (up) and (a,b)=(0.51,0.67) (down). Each of these choices respect the hypothesis of Lemma 6.4. We remark that the mode with largest amplitude is such that αj/ω1.

    Figure 3.  Comparison between the asymptotic estimate on the amplitude of the first 20 modes for different values of ω and for different configurations of the piers.

    The estimates on each single mode of u allow us to study more precisely how the asymptotic H2norm of u varies as the position of the piers vary, i.e., as a and b varies (see Lemma 6.4). Since most suspension bridges have symmetrical piers with a=b[1/2,2/3], we restrict ourselves to the case where (a,b)[1/2,2/3]×[1/2,2/3]. We represent in Figure 4 the estimate on the asymptotic H2norm given by Lemma 6.4 in function of a and b, with δ=1.5, g=1.5 and ω=10 fixed. We remark that this figure does not give any information about the stability of the bridge as a and b vary. In fact, the stability of a bridge is more endangered by the concentration of the energy on a single mode than by the generalized oscillation of the structure.

    Figure 4.  Plot of a theoretical estimate of the asymptotic H2norm in function of a and b.

    In order to study the distribution of the H2norm among the modes, we introduce the concept of family of asymptotic ηprevailing modes.

    Definition 7.1. Let 0<η<1. We say that a weak solution of (2.5) has a family S={j1,jn} of asymptotic ηprevailing modes if

    lim suptQSu22<η4lim suptPSu22. (7.2)

    In Figure 5 we plot the number of ηprevailing modes for η=0.1. The value of the parameters is the same as in Figure 4, namely δ=1.5, g=1.5 and ω=10. We can observe that the asymptotic H2norm concentrates on few modes as a=b. Moreover, we notice how the energy turns out to be more dispersed among the modes when ab.

    Figure 5.  Number of 0.1prevailing modes in function of a and b.

    In conclusion, we are able to assert that under suitable smallness conditions on the asymptotic amplitude of the forcing term and on the nonlinearity, we are able to perform a rather accurate modal analysis for the nonlinear nonlocal beam equations considered. In particular, Figure 5, allows us to conclude that the more stable configurations are achieved when ab. This suggests that, according to the model considered, asymmetric suspension bridges are more stable than suspension bridges where the piers are symmetric with respect to the center of the deck.

    The author would like to express his sincere gratitude to an anonymous referee for the useful comments, remarks and recommendations which definitely helped to improve the readability and the quality of the paper.

    The author declares no conflict of interest.



    [1] U. Battisti, E. Berchio, A. Ferrero, F. Gazzola, Energy transfer between modes in a nonlinear beam equation, J. Math. Pure. Appl., 108 (2017), 885-917. doi: 10.1016/j.matpur.2017.05.010
    [2] E. Berchio, D. Buoso, F. Gazzola, On the variation of longitudinal and torsional frequencies in a partially hinged rectangular plate, ESAIM COCV, 24 (2018), 63-87. doi: 10.1051/cocv/2016076
    [3] F. Bleich, C. B. McCullough, R. Rosecrans, G. S. Vincent, The mathematical theory of vibration in suspension bridges, U. S. Dept. of Commerce, Bureau of Public Roads, Washington D. C., 1950.
    [4] E. Berchio, A. Ferrero, F. Gazzola, Structural instability of nonlinear plates modeling suspension bridges: mathematical answers to some long-standing questions, Nonlinear Anal. Real, 28 (2016), 91-125. doi: 10.1016/j.nonrwa.2015.09.005
    [5] E. Berchio, F. Gazzola, A qualitative explanation of the origin of torsional instability in suspension bridges, Nonlinear Analysis Theor., 121 (2015), 54-72. doi: 10.1016/j.na.2014.10.026
    [6] M. Berger, A new approach to the large deflection of plate, J. Appl. Mech., 22 (1955), 465-472. doi: 10.1115/1.4011138
    [7] P. Biler, Remark on the decay for damped string and beam equations, Nonlinear Analysis Theor., 10 (1986), 836-842.
    [8] D. Bonheure, F. Gazzola, E. Moreira dos Santos, Periodic solutions and torsional instability in a nonlinear nonlocal plate equation, SIAM J. Math. Anal., 51 (2019), 3052-3091. doi: 10.1137/18M1221242
    [9] E. H. De Brito, The damped elastic stretched string equation generalized: Existence, uniqueness, regularity and stability, Appl. Anal., 13 (1982), 219-233. doi: 10.1080/00036818208839392
    [10] D. Burgreen, Free vibrations of a pin-ended column with constant distance between pin ends, J. Appl. Mech., 18 (1951), 135-139. doi: 10.1115/1.4010266
    [11] T. Cazenave, A. Haraux, F. B. Weissler, A class of nonlinear completely integrable abstract wave equations, J. Dyn. Differ. Equ., 5 (1993), 129-154. doi: 10.1007/BF01063738
    [12] T. Cazenave, A. Haraux, F. B. Weissler, Detailed asymptotics for a convex Hamiltonian system with two degrees of freedom, J. Dyn. Differ. Equ., 5 (1993), 155-187. doi: 10.1007/BF01063739
    [13] T. Cazenave, F. B. Weissler, Asymptotically periodic solutions for a class of nonlinear coupled oscillators, Portugalie Math., 52 (1995), 109-123.
    [14] T. Cazenave, F. B. Weissler, Unstable simple modes of the nonlinear string, Quart. Appl. Math., 54 (1996), 287-305. doi: 10.1090/qam/1388017
    [15] P. Constantin, C. Foias, R. Temam, On the large time Galerkin approximation of the Navier-Stokes equations, SIAM J. Numer. Anal., 21 (1984), 615-634. doi: 10.1137/0721043
    [16] I. D. Chueshov, Introduction to the theory of onfinite-dimensional dissipative systems, Kharkiv, Ukraine: Acta Scientific Publishing House, 2002.
    [17] I. D. Chueshov, Long-time dynamics of Kirchhoff wave models with strong nonlinear damping, J. Differ. Equations, 252 (2012), 1129-1262.
    [18] I. D. Chueshov, Theory of functionals that uniquely determine the asymptotic dynamics of infinite-dimensional dissipative systems, Russ. Math. Surv., 53 (1998), 731-776. doi: 10.1070/RM1998v053n04ABEH000057
    [19] I. D. Chueshov, Finite-dimensionality of the attractor in some problems of the nonlinear theory shells, Math. USSR Sbornik, 61 (1988), 411-420. doi: 10.1070/SM1988v061n02ABEH003215
    [20] I. D. Chueshov, I. Lasiecka, Long-time behavior of second order evolution equations with nonlinear damping, Memoirs of AMS, 195 (2008), viii+183.
    [21] A. Eden, C. Foias, B. Nicolaenko, R. Temam, Exponential attractors for dissipative evolution equations, New York: John-Wiley, 1994.
    [22] A. Eden, A. J. Milani, Exponential attractors for extensible beam equations, Nonlinearity, 6 (1993), 457-479. doi: 10.1088/0951-7715/6/3/007
    [23] Eurocode 1: Actions on structures. Parts 1-4: General actions - Wind actions, The European Union Per Regulation 305/2011, Directive 98/34/EC & 2004/18/EC. Avaiable from: http://www.phd.eng.br/wp-content/uploads/2015/12/en.1991.1.4.2005.pdf.
    [24] C. Foias, G. Prodi, Sur le comportement global des solutions non stationnaires des equations de Navier-Stokes en dimension deux, Rend. Sem. Mat. Univ. Padova., 39 (1967), 1-34.
    [25] V. Ferreira, F. Gazzola, E. Moreira dos Santos, Instability of modes in a partially hinged rectangular plate, J. Differ. Equations., 261 (2016), 6302-6304. doi: 10.1016/j.jde.2016.08.037
    [26] C. Fitouri, A. Haraux, Sharp estimates of bounded solutions to some semilinear second order dissipative equations, J. Math. Pure. Appl., 92 (2009), 313-321. doi: 10.1016/j.matpur.2009.04.010
    [27] C. Fitouri, A. Haraux, Boundedness and stability for the damped and forced single well Duffing equation, Disc. Cont. Dyn. Sys., 33 (2013), 211-223. doi: 10.3934/dcds.2013.33.211
    [28] B. G. Galerkin, Rods and plates. Series occurring in various questions concerning the elastic equilibrium of rods and plates, Eng. Bull. (Vestn Inszh Tech), 19 (1915), 897-908.
    [29] M. Garrione, F. Gazzola, Nonlinear equations for beams and degenerate plates with piers, Cham: Springer, 2019.
    [30] S. Gasmi, A. Haraux, N-cyclic functions and multiple subharmonic solutions of Duffing's equation, J. Math. Pure. Appl., 97 (2012), 411-423. doi: 10.1016/j.matpur.2009.08.005
    [31] C. Gasparetto, F. Gazzola, Resonance tongues for the Hill equation with Duffing coefficients and instabilities in a nonlinear beam equation, Commun. Contemp. Math., 20 (2018), 1-22.
    [32] F. Gazzola, Mathematical models for suspension bridges. Nonlinear structural instability, Cham: Springer, 2015.
    [33] M. Ghisi, M. Gobbino, A. Haraux, An infinite dimensional Duffing-like evolution equation with linear dissipation and an asymptotically small source term, Nonlinear Anal. Real, 43 (2018), 167-191. doi: 10.1016/j.nonrwa.2018.02.007
    [34] M. Ghisi, M. Gobbino, A. Haraux, Small perturbations for a Duffing-like evolution equation involving non-commuting operators, Nonlinear Differ. Equ. Appl., 28 (2021), 14. doi: 10.1007/s00030-021-00679-7
    [35] I. Giosan, P. Eng, Structural Vortex Shedding Response Estimation Methodology and Finite Element Simulation - Vortex Shedding Induced Loads on Free Standing Structures. Available from: https://bit.ly/3zEHNOy.
    [36] A. Haraux, On the double well Duffing equation with a small bounded forcing term, Rend. Accad. Naz. Sci. XL Mem. Mat. Appl., 29 (2005), 207-230.
    [37] A. Haraux, A sharp stability criterion for single well Duffing and Duffing-like equations, Nonlinear Anal., 190 (2020), 111600. doi: 10.1016/j.na.2019.111600
    [38] P. J. Holmes, J. E. Marsden, Bifurcation to divergence and flutter in flow-induced oscillations: an infinite dimensional analysis, Automatica, 14 (1978), 367-384. doi: 10.1016/0005-1098(78)90036-5
    [39] P. J. Holmes, J. E. Marsden, Bifurcations of dynamical systems and nonlinear oscillations in engineering systems, In: Nonlinear Partial Differential Equations and Applications, Berlin, Heidelberg: Springer, 1978,163-206.
    [40] P. J. Holmes, J. E. Marsden, A partial differential equation with infinitely many periodic orbits: chaotic oscillations of a forced beam, Arch. Rational Mech. Anal. 76 (1981), 135-165.
    [41] J. Howell, K. Huneycutt, J. T. Webster, S. Wilder, A thorough look at the (in)stability of piston-theoretic beams, Mathematics in Engineering, 1 (2019), 614-647. doi: 10.3934/mine.2019.3.614
    [42] J. S. Howell, D. Toundykov, J. T. Webster, A cantilevered extensible beam in axial flow: semigroup well-posedness and postflutter regimes, SIAM J. Math. Anal., 50 (2018), 2048-2085. doi: 10.1137/17M1140261
    [43] O. Ladyzhenskaya, A dynamical system generated by Navier-Stokes equation, J. Soviet Math., 3 (1975), 458-479. doi: 10.1007/BF01084684
    [44] W. S. Loud, On periodic solutions of Duffing's equation with damping, J. Math. Phys., 34 (1955), 173-178. doi: 10.1002/sapm1955341173
    [45] W. S. Loud, Boundedness and convergence of solutions of x"+cx+g(x)=e(t), Duke Math. J., 24 (1957), 63-72.
    [46] J. L. Luco, J. Turmo, Effect of hanger flexibility on dynamic response of suspension bridges, J. Eng. Mech., 136 (2010), 1444-1459. doi: 10.1061/(ASCE)EM.1943-7889.0000185
    [47] F. Moon, P. Holmes, A magnetoelastic strange attractor, J. Sound Vib., 65 (1979), 275-296. doi: 10.1016/0022-460X(79)90520-0
    [48] F. Nakajima, G. Seifert, The number of periodic solutions of 2-dimensional periodic systems, J. Differ. Equations, 49 (1983), 430-440. doi: 10.1016/0022-0396(83)90005-0
    [49] M. A. Jorge da Silva, V. Narciso, Attractors and their properties for a class of nonlocal nonextensible beams, Disc. Cont. Dyn. Sys., 35 (2015), 985-1008. doi: 10.3934/dcds.2015.35.985
    [50] M. A. Jorge da Silva, V. Narciso, Long-time dynamics for a class of extensible beam with nonlocal nonlinear damping, Evol. Equ. Control The., 6 (2017), 437-470. doi: 10.3934/eect.2017023
    [51] F. C. Smith, G. S. Vincent, Aerodynamic stability of suspension bridges: with special reference to the Tacoma Narrows Bridge, Part Ⅱ: Mathematical analysis, Investigation conducted by the Structural Research Laboratory, University of Washington - Seattle: University of Washington Press, 1950.
    [52] E. S. Titi, On approximate inertial manifolds to the Navier-Stokes Equations, J. Math. Anal. Appl., 149 (1990), 540-557. doi: 10.1016/0022-247X(90)90061-J
    [53] S. Woionosky-Krieger, The effect of an axial force on the vibration of hinged bars, J. Appl. Mech., 17 (1950), 35-36. doi: 10.1115/1.4010053
    [54] Z. Yang, Y. Li, F. Da, Robust attractors for a perturbed nonautonomous extensible beam equation with nonlinear nonlocal damping, Disc. Cont. Dyn. Sys., 39 (2019), 5975-6000. doi: 10.3934/dcds.2019261
    [55] Z. Yang, Y. Li, Criteria on the existence and stability of pullback exponential attractors and their application to non-autonomous Kirchhoff wave models, Disc. Cont. Dyn. Sys., 38 (2018), 2629-2653. doi: 10.3934/dcds.2018111
  • Reader Comments
  • © 2022 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1929) PDF downloads(111) Cited by(0)

Figures and Tables

Figures(5)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog