1 Introduction

Quasi-periodic oscillations are vibrations containing multiple incommensurate frequencies. They have been reported in physics, chemistry and engineering. For example, Ditto et al. [21] study quasi-periodic oscillations of a magneto-elastic ribbon and Hauck and Schneider [30] observe quasi-periodic oscillations in an oxidation reaction. Engineering applications include Gendelman et al. [26] who utilize the emergence of quasi-periodic oscillations for vibration absorption in an electric circuit. Moreover, Coudeyras et al. [16] observe quasi-periodic vibrations impacting break squeal and Kim and Noah [37] examine quasi-periodic responses of a nonlinear Jeffcott rotor with bearing clearance.

It seems naturally to expect quasi-periodic vibrations in mechanical systems when the applied external loads are quasi-periodic. This expectation, however, is generally not true even for the periodic case, as we have investigated in depth for mechanical systems [8]. Consequently, the question arises when we can rigorously expect a quasi-periodic steady-state response of a forced nonlinear mechanical system.

Numerical time integration seems to be a natural candidate to compute the response of nonlinear mechanical systems. Due to the low damping in modern engineering structures, however, long integration times are required to observe a steady-state response. Moreover, quasi-periodic orbits lack a fixed period after which the steady-state response repeats itself. Thus, additional signal processing tools, such as Fourier transformations are required to even recognize a quasi-periodic steady state. For example Kreider and Nayfeh [40] observed “unexplained sidebands” in the measured forced response of a clamped–clamped beam. Later, Emam et al. [22] explained these sidebands with the occurrence of quasi-periodic motions. Moreover, Held and Jefferies [32], Cumming and Linsay[17] and He et al. [31] use Poincare sections (cf. Guckenheimer and Holmes [28]) to recognize quasi-periodicity. A periodic motion occurs as single point inside a Poincare section, whereas a quasi-periodic orbit with two incommensurate frequencies leaves a densely filled orbit inside the trap. Orbits with more incommensurate frequencies generally generate more complicated patterns inside a Poincare section (cf. Cumming and Linsay [17] Fig. 2b for an example). Other techniques have been used by, e.g., Walden et al. [60] who assess the number of incommensurate basis frequencies to decompose their measurements of a Rayleigh–Bérnard convection. Moreover, Laskar [42] presents a mathematically basis to decompose a signal into its quasi-periodic components including a convergence proof.

Motivated by the computational success of the harmonic balance procedure in computing periodic orbits (e.g., Mickens [46]), Chua and Ushida [14] extended the harmonic balance method to the quasi-periodic case. In this procedure, an assumed quasi-periodic solution is expanded in a multi-frequency Fourier series and the governing ordinary differential equations of the nonlinear system are projected onto a finite set of Fourier modes. The arising nonlinear algebraic system is then solved with, e.g., the Newton–Raphson iteration. Recently, the harmonic balance procedure has been generalized by Schilder et al. [57], who also proposed a finite difference approximation. Their version is also implemented in the continuation package coco (cf. Dankowicz and Schilder [19]). The error due to the truncation of the generally infinite number of Fourier modes of quasi-periodic solutions, however, is unknown and the convergence of this method is a priori unclear. Furthermore, if the existence of a quasi-periodic is not shown, the expansion of the solution in a multi-frequency Fourier series is unjustified. Thus, the harmonic balance method is generally unreliable in those cases.

If the mechanical system is close to a solvable limit, perturbation methods can rigorously compute quasi-periodic orbits. Examples of general perturbation methods are the method of averaging (cf. Sanders et al. [55]), normal forms (cf. Murdock [47]) or the method of multiple time scales (cf. Nayfeh [48]). Moreover, Samoilenko [54] establishes a perturbation approach of quasi-periodic tori relying on a Green’s function. Similarly, the KAM-Theorem (cf. Guckenheimer and Holmes [28]) can guarantee the existence of quasi-periodic orbits in Hamiltonian systems under suitable conditions. The validity of all these methods, however, is restricted to a small neighborhood of the solvable, i.e., unphysical, limit. Whether the actual physical system is contained within this neighborhood is generally uncertain.

Fixed point theorems such as Brouwer’s or Schauder’s fixed point theorem (cf. Bobylev et al. [5] or Precup [51]) are powerful tools to prove the existence of solutions in the absence of small parameters. While the existence of periodic solutions is extensively treated in the literature (cf. Farkas [23], Rouche and Mawhin [52] or Bobylev [5]) the quasi-periodic or almost-periodicFootnote 1 case is less prominent (cf. Fink [24] and Corduneanu [15]).

Fink [24] guarantees the existence of an almost-periodic solution for one-dimensional first-order systems, if a monotonicity condition is met. Still relying on a monotonicity condition, Dafermos [18] extends the scalar results to abstract evolution equations in Hilbert spaces. His results can guarantee the existence of quasi-periodic responses in a fairly general setting for multi-dimensional first-order systems. The monotonicity condition, however, is not satisfied for simple mechanical systems such as the linear damped oscillator as we examine in Appendix A. Hence, the results of Dafermos [18] are inapplicable in the structural dynamics context.

Berger and Chen [3] give conditions for the existence of almost- and quasi-periodic solutions for second-order systems. Their results have been extended by, e.g., Carminati [11], Bolt et al. [4] and Mahwin [45], who also cover the case of bounded and almost-periodic solutions. Moreover, their theorems can guarantee the uniqueness of the bounded solution. These results, however, require again a monotonicity condition on the geometric nonlinearities (i.e., position dependent terms). In structural engineering this condition requires the stiffness to be negative definite (cf. Appendix A). This condition, however, is generally not met for oscillatory systems.

Alonso and Ortega [1], notably, derive an existence criterion for bounded solutions of second-order equations, which can be directly applied to oscillatory systems having a positive definite stiffness matrix. Their result guarantees the existence of a unique, asymptotically stable and bounded solution. The condition of their theorem, however, requires the Hessian of the nonlinear stiffness terms to the globally positive definite. This requirement excludes softening nonlinearities which have been observed experimentally by, e.g., Cho et al. [13] in a micro-mechanical system and by Yuan [61] in energy harvesting applications. Furthermore, the theorem by Alonso and Ortega [1] suffers from a restrictive linear global upper bound on the nonlinear terms.

In summary, the existence criteria for quasi-periodic solutions from the literature are inapplicable even for simple nonlinear oscillators. For the periodic case, however, various results can establish the existence of periodic solutions even for high amplitude oscillations. In our previous work [8], we identified a strengthened version of a theorem by Rouche and Mawhin [52] to be the most relevant for periodically driven mechanical systems. Here, we extend our result to the quasi-periodic case.

2 Set-up

We consider the general N-dimensional mechanical system of the form

$$\begin{aligned} \mathbf {M}\ddot{\mathbf {q}}+\mathbf {C}\dot{\mathbf {q}}+\mathbf {S}(\mathbf {q})= \mathbf {f}(t), \qquad \mathbf {q}\in \mathbb {R}^N, \end{aligned}$$
(1)

where the mass matrix \(\mathbf {M}\in \mathbb {R}^{N\times N}\) is symmetric and positive definite and the damping matrix \(\mathbf {C}\in \mathbb {R}^{N\times N}\) is symmetric. The vector \(\mathbf {S}(\mathbf {q})\) collects all position dependent forces, such a linear and nonlinear stiffness forces or non-potential forces (cf. Bolotin [6] for examples). The external forcing \(\mathbf {f}(t)\) is assumed to be quasi-periodic according to the following definition

Definition 2.1

The function \(\mathbf {f}(\varvec{\phi }(t))\) is quasi-periodic, if it maps the K-dimensional torus\(\mathbb {T}^K:=\left[ 0,2\pi \right] \times \left[ 0,2\pi \right] \times \ldots \times \left[ 0,2\pi \right] \) to \(\mathbb {R}^N\)

$$\begin{aligned} \mathbf {f}: \mathbb {T}^K \mapsto \mathbb {R}^N,\qquad \varvec{\phi }\mapsto \mathbf {f}(\varvec{\phi }), \end{aligned}$$
(2)

where the time evolution of the angles \(\varvec{\phi }\) is generated by K incommensurate frequencies

$$\begin{aligned}&\dot{\varvec{\phi }}=\varvec{\Omega },\qquad \varvec{\Omega }\in \mathbb {R}^K,\qquad \langle \varvec{\kappa } , \varvec{\Omega } \rangle \ne 0, \nonumber \\&\quad \text {for all}~~ \varvec{\kappa }\in \mathbb {Z}^K-\{0\}. \end{aligned}$$
(3)

A forced response of system (1) is most commonly computed by evaluating Eq. (1) along an assumed quasi-periodic solution \(\mathbf {q}^*(\varvec{\Omega }t)\) to E. (1). Projecting \(\mathbf {q}^*(\varvec{\Omega }t)\) on the torus \(\mathbb {T}^K\) Eq. (1) yields a partial differential equation on the bounded domain \(\mathbb {T}^K\), which can be discretized by, e.g., Fourier modes to yield the harmonic balance method (cf. Chua and Ushida [14]). Alternatively, Schilder et al. [57] propose a finite difference approximation. Since these methods evaluate Eq. (1) along a quasi-periodic orbit, they are only justifiable if a quasi-periodic solution to Eq. (1) actually exists. Applying the harmonic balance to a system without clarifying the existence of a steady-state response, can lead to erroneous conclusions as we have demonstrated on an explicit, periodically driven mechanical system [8]. To aid the existing numerical tools, it is thus of fundamental importance to guarantee the existence of quasi-periodic solutions to Eq. (1).

For periodically excited nonlinear mechanical systems numerous existence criteria are readily available (e.g., Farkas [23], Rouche and Mawhin [52] or Bobylev [5]). For the quasi-periodic case, however, the literature lacks applicable existence results for realistic nonlinear mechanical systems as previously detailed (see also Appendix A). Due to their resemblance, one could intuitively expect that the quasi-periodic case does not significantly differ from the periodic case. Unfortunately, this expectation is false and existence criteria from the periodic case do not straightforwardly extend to the quasi-periodic case. To demonstrate a delicate differences between periodic and quasi-periodic oscillations, we consider the linear oscillator

$$\begin{aligned}&\ddot{q}+\omega _0^2 q=f(\varvec{\phi }),\qquad \dot{\varvec{\phi }}=\varvec{\Omega }, \qquad q\in \mathbb {R},\quad \varvec{\phi } \in \mathbb {T}^K,\nonumber \\&\langle \varvec{\Omega },\varvec{\kappa }\rangle \ne 0, \quad \text {for all}~~\varvec{\kappa }\in \mathbb {Z}^K-\{0\}. \end{aligned}$$
(4)

For the periodic case (\(K=1\) in Eq. (4)), a periodic response exists if the forcing is non-resonant, i.e., \(n \Omega \ne \omega _0\) holds for all integers n (cf. Burd [10] Remark 2.1). Extending these results to the more general quasi-periodic case (\(K>1\)) one would intuitively expect that as long as the frequencies of the quasi-periodic forcing \(\langle \varvec{\Omega },\varvec{\kappa }\rangle \) are not equal to the natural frequency \(\omega _0\) (i.e., \(|\langle \varvec{\Omega },\varvec{\kappa }\rangle |\ne \omega _0 \)), then a quasi-periodic solution to system (4) exists. This expectation, however, is false. Even worse, for any incommensurate frequency basis \(\varvec{\Omega }\) there exists a quasi-periodic forcing such that the linear oscillator has no quasi-periodic solution (4). More specifically, in Appendix B we show the following:

Fact 2.1

For any incommensurate frequency basis \(\varvec{\Omega }\in \mathbb {R}^K\) with at least two frequencies (\(K\ge 2\)), there exists a continuous and bounded forcing \(f(\varvec{\phi })\) such that system (4) has no quasi-periodic solution.

Proof

We rely on the occurrence of small denominators, which arise in system (4) for any incommensurate frequency basis. We detail the derivations leading to the above fact in Appendix B. \(\square \)

In summary, applying a quasi-periodic excitation to the undamped linear oscillator (4) does not guarantee a quasi-periodic response (cf. Fact 2.1), while for almost all periodic forcings (i.e., except forcing at exact resonance) a periodic steady-state response arises. This observation demonstrates that results from the periodic system do not always straightforwardly extend to the quasi-periodic case.

3 Existence of quasi-periodic solutions

We prove the following general result:

Theorem 3.1

Assume that the quasi-periodic forcing \(\mathbf {f}(t)\) (cf. Definition 2.1) is continuous and the following conditions hold:

  1. (C1)

    The damping matrix \(\mathbf {C}\) is positive or negative definite, which implies that there exists a constant \(C_0>0\) such that

    $$\begin{aligned} |\mathbf {x}^{\mathrm{T}}\mathbf {C}\mathbf {x}|>C_{0}|\mathbf {x}|^2, \qquad \text { for all ~} \mathbf {x}\in \mathbb {R}^N, \end{aligned}$$
    (5)

    holds.

  2. (C2)

    The stiffness terms \(\mathbf {S}(\mathbf {q})\) derive from a potential, i.e., there exists a continuously differentiable scalar function \(V(\mathbf {q})\) such that

    $$\begin{aligned} \mathbf {S}(\mathbf {q})=\frac{\partial V(\mathbf {q})}{\partial \mathbf {q}}, \end{aligned}$$
    (6)

    holds.

  3. (C3)

    The stiffness terms \(\mathbf {S}(\mathbf {q})\) are continuous and globally Lipschitz, i.e., there exists a constant \(L>0\) such that

    $$\begin{aligned}&|\mathbf {S}(\mathbf {q}_1)-\mathbf {S}(\mathbf {q}_2)|\le L|\mathbf {q}_1-\mathbf {q}_2|,\qquad \nonumber \\&\qquad \text {for all}~~\mathbf {q}_1,\mathbf {q_2}\in \mathbb {R}^N, \end{aligned}$$
    (7)

    holds.

  4. (C4)

    The inner product of the stiffness terms \(\mathbf {S}(\mathbf {q})\) with the coordinates \(\mathbf {q}\) grows sufficient far away from the origin. Specifically, there exists a radius r and a minimal growths \(\alpha >0\), such that

    $$\begin{aligned}&\mathbf {q}^{\top }\mathbf {S}(\mathbf {q})>\alpha |\mathbf {q}|^2,~~\text {or}~~-\mathbf {q}^{\top }\mathbf {S}(\mathbf {q})>\alpha |\mathbf {q}|^2,\nonumber \\&\qquad \text { for all ~} |\mathbf {q}|>r, \end{aligned}$$
    (8)

    holds.

Then, system (1) has a differentiable quasi-periodic solution \(\mathbf {q}(\varvec{\Omega }t)\).

Proof

We base our proof of Theorem 3.1 on a fixed point theorem by Schäfer [56] (also see Smart [59]). We establish a homotopy between system (1) and the equation \(\mathbf {M}\ddot{\mathbf {q}}+\mathbf {C}\dot{\mathbf {q}}\pm \mathbf {M} \mathbf {q}=0\). Conditions (C1)–(C4) guarantee that the solutions remains bounded for all homotopy parameters. We detail our derivations in Appendix C and give a concise proof referring to relevant preliminary results and definitions in Sect. C.5. \(\square \)

Remark 3.1

Since periodic orbits are a special case of quasi-periodic solutions (i.e., \(K=1\) in Definition 2.1), one can also establish the existence of periodic solutions via Theorem 3.1. Comparing with our previous result [8] conditions (C1) and (C2) are identical. In the quasi-periodic case, however, we require inner product of the coordinates \(\mathbf {q}\) and the stiffness terms \(\mathbf {S}(\mathbf {q})\) to grow sufficient far from the origin (cf. condition (C4)), while in the periodic case it is sufficient that the stiffness terms are element-wise above or below the mean forcing value. Furthermore, we require the existence of a global Lipschitz constant in condition (C3). These deviations arise from the different fixed point theorems employed in the proofs of both results.

Remark 3.2

We emphasize that our Theorem 3.1 guarantees the existence of a quasi-periodic solution to Eq. (1) with the frequency base \(\varvec{\Omega }\) arising from the external excitation. This enables to approximate the steady-state solutions by Fourier series with known frequencies basis \(\varvec{\Omega }\). In practice this is the most common approach to compute quasi-periodic orbits (e.g., Schilder et al. [57] or Chua and Ushida [14]). Competing results in a similar setting of our Theorem 3.1, e.g., by Alonso and Ortega [1] Lemma 3.2, can only guarantee the existence of a bounded solution. Hence, an approximation by Fourier series is generally not justified by such a result.

Generally, results guaranteeing the existence of a bounded solution, such as Alonso and Ortega’s Lemma, can be strengthened to establish that an existing bounded solution is quasi-periodic with the same frequency base (or module) as the external loads. Thereby, Fourier series approximations with the frequency basis \(\varvec{\Omega }\) are again justified. In practice, however, the necessary arguments rely on very restrictive assumptions. For example, if the bounded solution is unique and stable, then general results on module containment (cf. Fink [24]) guarantee that the bounded solution is quasi-periodic with the same frequencies as the forcing. For nonlinear mechanical systems of the form (1), however, already in the periodic case multiple co-existing steady-state solutions and unstable periodic orbits are frequently observed (e.g., Nayfeh and Mook [49]).

In the case of multiple coexisting solutions, it is non-trivial whether these are actually even almost-periodic. For such a result one would need to show that only finitely many, semi-separated solutions exist (cf. Fink [24]). However, the periodically forced damped Duffing oscillator, a prominent example of system (1), can have infinite number of periodic orbits if a homoclinic tangle occurs (cf. Guckenheimer and Holmes [28]). Even if one achieves to establish that only finitely many semi-separated solutions exist, then the frequency basis of these almost-periodic solutions is generally not the same as the module of the time-varying terms (see Fink [24] for an explicit example). Thereby, an expansion of the solution in a Fourier series with frequency basis \(\varvec{\Omega }\) is generally unjustified.

Remark 3.3

If the stiffness terms in system (1) are polynomials such as in the classical Duffing oscillator, then condition (C3) is not satisfied. However, the Lipschitz constant L in Eq. (7) can be chosen arbitrarily large, which allows to conclude that condition (C4) is generally uncritical for realistic structures. Since any physical system will not allow for arbitrarily large displacements and velocities, equations of the form (1) modeling a realistic system are generally only valid in a bounded domain \(U\subset \mathbb {R}^N\). Outside this domain of validity, Eq. (1) does not represent the reality. Thus, it is justified to modify Eq. (1) outside U such that the stiffness terms are globally Lipschitz. Such a truncation can always be carried out if \(\mathbf {S}(\mathbf {q})\) is continuous. Then, \(\mathbf {S}(\mathbf {q})\) is uniformly continuous on the compact domain U. Hence, \(\mathbf {S}(\mathbf {q})\) is locally Lipschitz and there exists a Lipschitz constant L such that condition (C3) holds for all \(\mathbf {q}_1\) and \(\mathbf {q}_2\) in U.

We explicitly construct an example of such an envisioned truncation. We select a ball of radius r containing U (\(U\subset \mathcal {B}_r \subset \mathbb {R}^N\)) and denote the point on the boundary of \(\mathcal {B}_r \) with minimal distance to \(\mathbf {q}\) by \(\mathbf {q}^*(\mathbf {q})\). With this notation, we define the nonlinearities

$$\begin{aligned} \mathbf {S}_1(\mathbf {q})= & {} {\left\{ \begin{array}{ll} \mathbf {S}(\mathbf {q}) &{} |\mathbf {q}|<r,\\ \mathbf {S}(\mathbf {q}^*(\mathbf {q})), &{} |\mathbf {q}|\ge r, \end{array}\right. }\nonumber \\ \mathbf {S}_2(\mathbf {q})= & {} {\left\{ \begin{array}{ll} 0 &{} |\mathbf {q}|<r,\\ \mathbf {K}^*(\mathbf {q}-r\frac{\mathbf {q}}{|\mathbf {q}|}), &{} |\mathbf {q}|\ge r, \end{array}\right. }, \end{aligned}$$
(9)

where \(\mathbf {K}^*\) is a positive or negative definite matrix. We note that \(\mathbf {S}_2\) satisfies condition (C3) and hence if \(\mathbf {S}\) is continuous, then the system

$$\begin{aligned} \mathbf {M}\ddot{\mathbf {q}}+\mathbf {C}\dot{\mathbf {q}}+\mathbf {S}_1(\mathbf {q})+\mathbf {S}_2(\mathbf {q})= \mathbf {f}(t), \qquad \mathbf {q}\in \mathbb {R}^N, \end{aligned}$$
(10)

is Lipschitz continuous and satisfies condition (C3). Thus, if the remaining conditions (C1), (C2) and (C4) of our Theorem 3.1 are met, a quasi-periodic response exists.

Remark 3.4

We have restricted our Theorem 3.1 to time independent nonlinearities, which are customary in structural dynamics. However, our results can be extended to include time dependent nonlinearities \(\mathbf {S}(\mathbf {q},\varvec{\Omega } t)\), if the time dependence is quasi-periodic. Then, if conditions (C2)–(C4) are satisfied for all times, then the assertion of Theorem 3.1 still holds.

Remark 3.5

The existence of a quasi-periodic orbit, e.g., established by our Theorem 3.1 does not rule out the existence of other invariant objects such as periodic orbits, fixed points or chaotic attractors. Indeed, the coexistence of multiple stable or unstable attractors is a fundamental characteristic of nonlinear dynamical systems (e.g., Guckenheimer and Holmes [28]). Moreover, a general connection between quasi-periodic motion and chaos has been established by Ruelle and Takens [53] and Newhouse et al. [50]. A quasi-periodic route to chaos has been experimentally observed by, e.g., Held and Jeffries [32], Cumming and Linsay [17] and He et al. [31]. Since Theorem 3.1 can guarantee the existence of quasi-periodic motions, it can be applied in numerical investigation of the quasi-periodic route to chaos in the nonlinear mechanical system (1).

Remark 3.6

From the existence Theorem 3.1, we can conclude, that one can observe a quasi-periodic motion if the dynamical system (1) is initialized on one of the quasi-periodic orbits. Whether other initial conditions are attracted to this quasi-periodic orbit or whether it is stable with a basin of attraction cannot be inferred from our Theorem 3.1. Moreover, to the best of our knowledge, a robust and universal tool to conclude about the stability of a quasi-periodic orbit is not available. Although only stable orbits are observable, it is also of interest to compute unstable orbits. For example, Denis et al. [20] utilize unstable periodic orbits for system identification.

Condition (C4) specifies the sign of the inner product of the nonlinearity \(\mathbf {S}(\mathbf {q})\) and the coordinate \(\mathbf {q}\). Geometrically this implies that far away from the origin the vector \(\mathbf {S}(\mathbf {q})\) points either toward the origin (\(\mathbf {q}^{\top }\mathbf {S}(\mathbf {q})<0\)) or away from the origin (\(\mathbf {q}^{\top }\mathbf {S}(\mathbf {q})>0\)). These considerations play a major role in the construction of guiding functions and also index theory (cf. Krasonosel’skii et al. [39], Krasnosel’skii and Zabreĭko [38] or Bobylev et al.  [5]). The application reported, however, focus on either periodic or bounded solutions (see also Remark 3.2).

If the nonlinearities are continuously differentiable, then condition (C4) of Theorem 3.1 can be replaced by a more intuitive condition.

Theorem 3.2

If the nonlinearities \(\mathbf {S}(\mathbf {q})\) are continuously differentiable, condition (C4) holds if

  1. (C4*)

    The Hessian of \(V(\mathbf {q})\) is positive or negative definite for \(|\mathbf {q}|>\tilde{r}\), i.e., for some constant \(C_S>0\)

    $$\begin{aligned}&\!\!\!\mathbf {x}^{\top } \frac{\partial ^2 V(\mathbf {q})}{\partial \mathbf {q}^2} \mathbf {x}> C_S |\mathbf {x}|^2,\quad \text {or}\nonumber \\&\!\!\!\quad -\mathbf {x}^{\top } \frac{\partial ^2 V(\mathbf {q})}{\partial \mathbf {q}^2} \mathbf {x}> C_S |\mathbf {x}|^2 ,\nonumber \\&\!\!\!\quad \qquad \text {for all ~} |\mathbf {q}|>\tilde{r},~~ \mathbf {x}\in \mathbb {R}^N, \end{aligned}$$
    (11)

    holds.

Proof

We proof the above theorem in Appendix D. \(\square \)

Remark 3.7

We note that condition (C4*) requires differentiable nonlinearities, which is not required by condition (C4). Condition (C4*) is, however, generally easier to verify, since positive or negative definiteness can be checked through direct eigenvalue computation, leading minor criteria (or Sylvester’s criterion) or Cholesky decomposition (cf. Horn and Johnson [33]).

On our way to Theorem 3.1, we have established an analogous result for non-smooth but \(L_2\)-integrable external forcing, i.e.,

$$\begin{aligned} ||\mathbf {f}(t)||_{L_2}:= & {} \left( \frac{1}{(2\pi )^K}\int _{\mathbb {T}^K} \mathbf {f}^{\top }(\varvec{\phi })\mathbf {f}(\varvec{\phi })~{\mathrm{d}}\varvec{\phi } \right) ^{\frac{1}{2}}\nonumber \\= & {} \left( \lim _{T\rightarrow \infty }\frac{1}{2T}\int _{-T}^{T} \mathbf {f}^{\top }(t)\mathbf {f}(t)~{\mathrm{d}}t \right) ^{\frac{1}{2}} < \infty .\nonumber \\ \end{aligned}$$
(12)

Equation (12) is satisfied by, e.g., saw-tooth or square wave functions. For such non-smooth forcing functions, we have the following result:

Theorem 3.3

Assume that that the conditions (C1)–(C4), or equivalently condition (C4*) instead of condition (C4), are satisfied and the quasi-periodic forcing \(\mathbf {f}(t)\) (cf. Definition 2.1) satisfies Eq. (12). Then, system (1) has a quasi-periodic solution \(\mathbf {q}(\varvec{\Omega }t)\) with finite \(L_2\)-norm. Moreover, the time derivative of this solution is bounded in the \(L_2\)-norm.

Proof

Our proof of the above theorem is also based on a fixed point theorem by Schäfer [56] (also see Smart [59]). We detail our derivations in Appendix C and give an overview of the proof in Sect. C.4. \(\square \)

Moreover, basic results on the smoothness of solutions to ordinary differential equation (e.g., Hale [29] or Arnol’d [2]) for finite time guarantee the smoothness of the quasi-periodic solution

Corollary 3.1

(Smoothness) Assume the stiffness terms \(\mathbf {S}(\mathbf {q})\) and the forcing \(\mathbf {f}(t)\) of system (1) are s-times continuously differentiable (\(s>0\)) and the conditions of Theorem 3.1 hold. Then, the quasi-periodic solution guaranteed by Theorem 3.1s-times continuously differentiable.

Proof

Assume the contrary, i.e., the quasi-periodic solution \(\mathbf {q}^*\) of system (1) is non-smooth at some time instance \(t^*\). This statement is a direct contradiction to the existence of s-times continuous differentiable solutions for finite times proven by, e.g., Hale [29] or Arnol’d [2]. Hence, the quasi-periodic solutions are smooth as claimed in corollary (3.1). \(\square \)

4 Numerical examples

We demonstrate the applicability of our Theorem 3.1 on two oscillatory systems. First, we investigate a chain of oscillators proposed by Breunung and Haller [7]. This system is a natural extension of the classical Duffing oscillator or the Shaw–Pierre example [58] to higher dimensions. Subsequently, we apply Theorem 3.1 to a discretized slender beam. After concluding about the existence of a quasi-periodic steady-state response, we calculate it with the automated package proposed by Jain et al. [35].

4.1 Chain oscillator

In the following, we utilize our Theorem 3.1 to compute a steady-state response of a system of N masses depicted in Fig. 1. Two adjacent masses are connected with spring elements and linear dampers. The first mass is suspended to the wall, while the last mass is suspended to the moving base u(t). Additional to the base point excitation, the external loads \(f_j(t)\) are applied to the jth mass. For a first assessment, we assume purely external loads and select forcing frequencies in resonance with the first two eigenfrequencies. Subsequently, we show that our Theorem 3.1 also allows for complex loading scenarios, e.g., excitations arising from the natural roughness of a road surface.

Fig. 1
figure 1

Oscillator chain with base excitation

4.1.1 External loading

For the case \(u(t)=0\), the equation of motion for the jth mass is given by

$$\begin{aligned}&m_j\ddot{q}_j-c_j(\dot{q}_{j-1}-\dot{q}_{j})+c_{j+1}(\dot{q}_{j}-\dot{q}_{j+1})\nonumber \\&\quad -S_{j}(q_{j-1}-q_{j})+S_{j+1}(q_{j}-q_{j+1})=f_j(t),\nonumber \\&\qquad j=1,\ldots ,N, \end{aligned}$$
(13)

where the coordinates \(q_0\) and \(q_{N+1}\) are set to zero. The example (13) includes the classical Duffing oscillator in the case \(N=1\) and \(S_1(q_1)=\omega _0^2q+\kappa q^3\). Higher-dimensional variants of this system have been investigated by, e.g., Shaw and Pierre [58], Breunung and Haller [7] or Jain et al. [35]. In our previous work [8], we showed that if all the stiffness terms were hardening (\(\partial S_j(p)/\partial p>0 \)) then periodic forcing leads to a periodic response of system (13). For the quasi-periodic case, we have a similar statement

Fact 4.1

Assume the masses \(m_j\) and damping coefficients \(c_j\) are positive and the forcing functions \(f_j(t)\) are quasi-periodic and continuous. If additionally all springs are either hardening or softening far away from the origin, i.e., for some \(\gamma >0\) and \(r>0\)

$$\begin{aligned} \begin{aligned}&S_{j}(q_{j-1}-q_{j})> \gamma (q_{j-1}-q_{j}), \\&\qquad |q_{j-1}-q_{j}|>r,\quad j=1,\ldots , N+1, \\&\text {or } \qquad -S_{j}(q_{j-1}-q_{j})> \gamma (q_{j-1}-q_{j}), \\&\qquad |q_{j-1}-q_{j}|>r,\quad j=1,\ldots ,N+1, \end{aligned} \end{aligned}$$
(14)

holds, then system (13) has a quasi-periodic steady-state response.

Proof

We verify that system (13) satisfies the conditions of Theorem 3.1 in Appendix 5. \(\square \)

Fact 4.1 extends the result of our previous work [8] to the case of quasi-periodic forcing and also includes softening nonlinearities. As an example, we consider a chain of five masses with the parameters \(m_j=1\), \(c_j=0.02\) and softening springs of the form \(S_j(d)=d-0.5d^3\). For this configuration, Fact 4.1 guarantees the existence of a quasi-periodic steady-state response.

Selecting the quasi-periodic forcing \(f_1=0.006 (\sin (\Omega _1 t)+\sin (\Omega _2 t))\) and \(f_j=0\) for \(j=2,3,\ldots ,N\), we use the automated package of Jain et al. [35] to compute a quasi-periodic steady state. In this algorithm, the steady-state response is approximated with a finite number of Fourier coefficients. The arising nonlinear algebraic equations are solved either with the Picard or the Newton–Raphson iteration, where the response of the corresponding linear system serves as initial guess. While the existence criterion of Jain et al. [35] fails for forcing in resonance with an eigenfrequency, our Theorem 3.1 guarantees the existence of a quasi-periodic response for any forcing amplitude and frequency. We chose the forcing frequency \(\Omega _1\) to be close to the first eigenfrequency \(\omega _1\approx 0.52\) and \(\Omega _2\) in the vicinity of the second eigenfrequency \(\omega _2 \approx 1\). In Fig. 2, we depict the \(L_2\)-norm of the quasi-periodic steady-state response of the chain oscillator (13).

Fig. 2
figure 2

Quasi-periodic steady-state responses of system (13) with five masses and parameters \(m_j=1\), \(c_j=0.02\) and \(S_j(d)=d-0.5d^3\) to the external loads \(f_1=0.006(\sin (\Omega _1 t)+\sin (\Omega _2 t))\) and \(f_j=0\). We select the forcing frequencies \(\Omega _1 \) and \(\Omega _2\) in resonance with the first two natural frequencies \(\omega _1\approx 0.52 \) and \(\omega _2 \approx 1 \)

4.1.2 Base excitation through road roughness

Many engineering systems are subject to complex or even random excitations. For example, various norms (e.g., MIL-STD-810G, ISO 4866 or ISO 1032) require the assessment of vibration amplitudes to loads with a specified power spectral density. Often it is assumed that these complex loading scenarios can be accurately modeled with a finite set of Fourier modes (cf. Loprencipe and Zoccali [43]). In such a setting the arising external loads are naturally quasi-periodic. In the following, we use our Theorem 3.1 to reliably compute an arising steady-state response to a complex excitation generate by a road surface.

We assume that the base point u(t) (cf. Fig. 1) moves with a constant speed v along a rough street. We envision, e.g., parts of a vehicle. Road profiles are often classified by their spatial spectral density according to the norm ISO 8608 [34] (cf. Loprencipe and Zoccali [43]). Furthermore, the smoothened power spectral densities given in [34] are commonly used for dynamical analysis of vehicles. We follow this approach to generate a road roughness profile according to ISO 8608. The smoothened spatial power spectral density from the standard [34] is given by

$$\begin{aligned} \hat{u}(\Omega )= C \left( \frac{\Omega }{\Omega _0}\right) ^{-2} \cdot 10^{-6}, \qquad \nonumber \\ 0.022\pi \le \Omega \le 5.66\pi , \qquad \nonumber \\ \Omega _0=1~ \mathrm {rad/m},\quad C=64~ \mathrm {m}^3, \end{aligned}$$
(15)

where \(\Omega \) is the frequency in \(\mathrm {rad/m}\) and the amplitudes \(\hat{u}(\Omega )\) are given in meters. The normalization constant \(\Omega _0\) is \(1~ \mathrm {rad/m}\) and the constant C is quantifies the road quality. For our analysis, we assume that the road belongs to the class with the second highest quality and set C to be \(64~ \mathrm {m}^3\). Following Loprencipe and Zoccali [43], we create an artificial road profile by selecting a finite number of frequencies \(\Omega _k\) whereby the amplitudes are given by Eq. (15). Then, the road profile is given by

$$\begin{aligned} u(s)=\sum _{k=1}^K \sqrt{\hat{u}(\Omega _k)}\sin ( \Omega _k s+\phi _k), \end{aligned}$$
(16)

where s is the distance traveled and \(\phi _k\) denote phases. We select the frequencies \(\Omega _k\) by drawing random samples from the specified frequency interval (15). To emphasize the presence of low frequency terms, we assume an underlying beta distribution (e.g., Jonson et al. [36]) with the parameters \(\alpha =1\) and \(\beta =5\). Moreover, we sample the angles \(\phi _k\) from a uniform distribution. We depict such a road profile with ten frequencies (\(K=10\) in Eq. (16)) in Fig. 3a.

Fig. 3
figure 3

Randomly generated road profile generated according to the norm ISO 8608 [34] and frequency spectrum of the forcing arising through base excitation with then base frequencies (\(K=10\)). We consider system (17) with five masses and the parameters \(k=100 ~\mathrm {N/m}\), \(k_r=10^4~ \mathrm {N/m}\), \(c=2~ \mathrm {Ns/m}\), \(\kappa =5000~\mathrm {N/m^3}\) and \(v=10~\mathrm {m/s}\). a Road profile. b Force arising from the road profile 3a and eigenfrequencies of system (17).

We consider the chain of oscillators depicted in Fig. 1 with five masses. Assuming springs with linear and cubic coefficients the equations of motion are

$$\begin{aligned} \begin{aligned}&m\ddot{q}_j+c(2\dot{q}_{j}-\dot{q}_{ j-1}-\dot{q}_{j+1}) +k(2q_{j}-q_{j-1}-q_{j+1})\\&\quad -\kappa (q_{j-1}-q_{j})^3\\&\quad + \kappa (q_{j}-q_{j+1})^3=0, \quad j=1,\ldots ,N-1,\\&m\ddot{q}_{N}+c(2\dot{q}_{N}-\dot{q}_{N-1}) +k( q_{N}-q_{N-1})\\&\quad +k_rq_n=k_r u(vt)+c \dot{u}(t)=f_r(vt), \end{aligned} \end{aligned}$$
(17)

where we selected a linear force displacement relationship for the last spring \(S_{N+1}\) with stiffness \(k_r\) and v denotes the constant speed (cf. Fig. 1). The forcing \(f_r(t)\) arising through the base point excitation is analytic and quasi-periodic for any finite number of frequencies K. The frequency spectrum of the forcing arising from the road profile shown in Fig. 3a is depicted in Fig. 3b. Selecting the positive mass \(m=1~\mathrm {kg}\), positive stiffness coefficients \(k=100 ~\mathrm {N/m}\), \(k_r=10^4~ \mathrm {N/m}\), \(\kappa =5000~\mathrm {N/m^3}\) and positive damping coefficient \(c=2~ \mathrm {Ns/m}\), Eq. (17) satisfies Fact 4.1, hence a quasi-periodic steady-state response exists.

Figure 3b shows that in the frequency band between 10 and \(20~\mathrm {rad/s}\) three eigenfrequencies and four forcing frequencies are in an immediate vicinity. Additionally, at \(6~\mathrm {rad/s}\) one of the forcing frequencies is in resonance with the lowest eigenfrequency. Due to these multiple resonances, we expect considerable vibration amplitudes.

We compute the arising steady-state response via the harmonic balance procedure with ten base frequencies, whereby the arising nonlinear algebraic equations are solved the Newton–Raphson iteration. As initial guess, we select the response of the linearization of system (17) at the origin. In Fig. 4, we depict the arising quasi-periodic steady-state response for the first one hundred meters. Additionally, we include the linear solution and a time series obtained by numerical time integration with matlab’s ode45 algorithm serving as benchmark solution. To this end, we consider the system to be at rest one thousand seconds before reaching \(s=0\), i.e., we select the initial condition \(\mathbf {q}(t=-10^{3})=\dot{\mathbf {q}}(t=-10^{3})=\mathbf {0}\). Figure 4 reveals that the harmonic balance solution matches very well with the numeric time integration, while the linear response differs. The maximal displacement of \(q_3\) along the arising steady-state response is \(97.2~\mathrm {mm}\), while a linear analysis leads to a maximal displacement of \(83.6~\mathrm {mm}\). Hence, linear analysis underestimates the response amplitude by more than 15 percent. Thus, it is essential to perform a nonlinear analysis on our example. Therein, Theorem 3.1 justifies the expansion the coordinates of system (17) in a Fourier series.

4.2 A slender beam

After applying our existence criterion 3.1 to the spring-mass systems (13), we advance to a beam serving as more realistic engineering structure. Beam elements are frequently used to model complex engineering structures and are readily implemented in common computational packages such as Abaqus, Autodesk or ANSYS. These computational packages discretize the model’s governing partial differential equations in space. Through the discretization high-dimensional ordinary differential equations arise. In the following, we apply our existence criterion 3.1 to such a high-dimensional oscillatory system. Subsequently, it justifies numerical computations of quasi-periodic responses.

We consider the clamped–clamped beam of length L depicted in Fig. 5. The material properties and geometry such as cross-sectional area A and moment of inertia I are assumed to be constant. Nayfeh and Mook [49] derive a partial differential equation governing the vertical vibrations of slender beams. They assume that the Euler–Bernoulli hypothesis is met (i.e., plane cross sections initially perpendicular to the neutral axis remain plane and perpendicular the neutral axis after deformation), linear elastic material properties (Hooke’s law) and use a third-order approximation of the Green–Lagrange strain tensor (van Kármán strains). With the additional assumption of a small radius of gyration r, i.e., \(r=\sqrt{I/A}\ll 1\), Nayfeh and Mook state the non-dimensional partial differential equation

$$\begin{aligned}&r^2(w_{tt}+2\mu w_t+w_{xxxx})-\left( \frac{1}{2L}\int _{0}^{L} w_x^2{\mathrm{d}}x\right) w_{xx}\nonumber \\&\quad =f(x,t),~\qquad w(0)=w_x(0)\nonumber \\&\quad =w(L)=w_x(L)=0, \end{aligned}$$
(18)

where w(xt) denotes the vertical displacement as a function of the non-dimensional horizontal position x and the non-dimensional time t (cf. Fig. 5). The subscripts of w in Eq. (18) refer to partial differentiation with respect to time t or space x and the constant \(\mu \) is a damping parameter.

Fig. 4
figure 4

Steady-state response of the third mass \(q_3\) of system (17) exited by the forcing 3b. We select the parameters \(k=100 ~\mathrm {N/m}\), \(k_r=10^4~ \mathrm {N/m}\), \(c=2~ \mathrm {Ns/m}\) and \(\kappa =5000~\mathrm {N/m^3}\)

Fig. 5
figure 5

Clamped–clamped beam: undeformed configuration in black and deformed configuration in gray

We discretize the partial differential equation (18) in the spatial coordinate following a finite difference scheme with N equally space nodes \(x_j\) (\(j=1,\ldots ,N\)). The parameter h denotes the distance between neighboring nodes. We use a forward scheme for to odd spatial derivative \(w_x\) and a central scheme for the even spatial derivatives which yields

$$\begin{aligned}&w_x(x_j)=\frac{w(x_{j+1})-w(x_{j}) }{h}+\mathcal {O}(h),\nonumber \\&w_{xx}(x_j)=\frac{w(x_{j+1})-2w(x_{j})+w(x_{j-1})}{h^2}0+\mathcal {O}(h^2),\nonumber \\&w_{xxxx}(x_j)\nonumber \\&\quad =\frac{w(x_{j+2})-4w(x_{j+1})+6w(x_{j}) -4w(x_{j-1})+w(x_{j-2}) }{h^4}\nonumber \\&\qquad +\mathcal {O}(h^2).\nonumber \\ \end{aligned}$$
(19)

To approximate the spatial derivative at the node \(x_N=L\), we use a backward-difference scheme, i.e., \(w_x(L)=(w(x_{N})-w(x_{N-1}))/h+\mathcal {O}(h))\). Moreover, we approximate the integral in Eq. (18) via the trapezoidal rule yielding

$$\begin{aligned}&\int _{0}^{L} w_x^2{\mathrm{d}}x \nonumber \\&\quad = \frac{h}{2} \left( w_x^2 (x_1)+2 \sum _{j=2}^{N-1} w_x^2 (x_j)+w_x^2 (x_N)\right) +\mathcal {O}(h^2) \nonumber \\&\quad = \frac{1}{2h} \left[ (w(x_{2})-w(x_1))^2+(w(x_N)-w(x_{N-1}))^2\right] \nonumber \\&\qquad + \frac{1}{h} \sum _{j=2}^{N-1} (w(x_{j+1})-w(x_{j}))^2+ \mathcal {O}(h), \end{aligned}$$
(20)

where we have used the discretization (19). Moreover, we concentrate the distributed load f(xt) onto the closest mesh point by defining the nodal forcing functions

$$\begin{aligned} f_j(t):=\int _{x_j-h/2}^{x_j+h/2} f(x,t)~{\mathrm{d}}x. \end{aligned}$$
(21)

The boundary conditions (18) imply that the vertical displacements of the first and last two elements are zero, i.e., \(w(x_1)=w(x_2)=w(x_{N-1})=w(x_N)=0\). We collect the remaining degrees of freedom in the vector \(\mathbf {w}:=[w(x_3),w(x_4),\ldots ,w(x_{N-2})]^{\top }\) and note that for all these nodes (\(x_3,x_4\ldots ,x_{N-2}\)) the finite differences (19) are well defined. To formulate a spatially discretized version of the partial differential equation (18), we define the forcing vector \(\mathbf {f}:= [f_3(t),f_4(t),\ldots ,f_{N-2}(t)]^{\top }\) and the banded matrices

$$\begin{aligned} \displaystyle \mathbf {K}_1:= & {} \begin{bmatrix} 2 &{} -1&{}0 &{} 0 &{} \\ -1 &{} 2 &{}-1 &{} 0 &{} \\ 0 &{} \ddots &{} \ddots &{}\ddots \end{bmatrix}\nonumber \\ \mathbf {K}_2:= & {} \begin{bmatrix} 6 &{} -4&{} 1 &{} 0 &{} \\ -4 &{} 6 &{}-4 &{} 1 &{} 0 \\ 1 &{}-4 &{} 6 &{}-4 &{} 1 &{} 0 \\ 0 &{} \ddots &{} \ddots &{}\ddots &{} \ddots &{}\ddots \end{bmatrix}\nonumber \\&\quad \mathbf {K}_1,\mathbf {K}_2\in \mathbb {R}^{N-4\times N-4}. \end{aligned}$$
(22)

With the notation (22), a discretized version of the partial differential equation (18) is given by

$$\begin{aligned}&\ddot{\mathbf {w}} +2\mu \dot{\mathbf {w}} + \frac{1}{h^4} \mathbf {K}_2 \mathbf {w} +\frac{1}{2 L r^2 h^3 } (\mathbf {w}^{\top }\mathbf {K}_1\mathbf {w}) \mathbf {K}_1\mathbf {w}\nonumber \\&\quad =\frac{1}{r^2} \mathbf {f}(t). \end{aligned}$$
(23)

After deriving Eq. (23) as a discretization of the partial differential(19), we will now show that our Theorem 3.1 guarantees the existence of a quasi-periodic response of system (23). We emphasize that our result is especially relevant for Eq. (23). Competing perturbations methods commonly extract a steady-state response from linearizing Eq. (23) around the origin (\(\mathbf {w}=\dot{\mathbf {w}}=0\)). Subsequently, those approaches argue that a steady-state response of the linearization implies a close by steady-state response of the full nonlinear system if the nonlinearity is small enough. The nonlinearity in Eq. (23), however, is dominant , since the slenderness parameter r is small. Hence, the validity of regular perturbation methods is restricted to a tiny neighborhood of the origin. Our result 3.1 does not suffer from such a shortcoming.

We proceed by verifying that conditions (C1)–(C4) are satisfied for system (23). Indeed, for any nonzero value \(\mu \) condition (C1) is satisfied. Furthermore, differentiation of the function \(V(\mathbf {w}):= 1/(2 h^4) \mathbf {w}^{\top }\mathbf {K}_2 \mathbf {w}+1/(4 L r^2 h^3) (\mathbf {w}^{\top }\mathbf {K}_1\mathbf {w})^2\) yields

$$\begin{aligned} \frac{\partial V(\mathbf {w})}{\partial \mathbf {w}}=\frac{1}{h^4} \mathbf {K}_2 \mathbf {w}+\frac{1}{2 L r^2 h^3} (\mathbf {w}^{\top }\mathbf {K}_1\mathbf {w} ) \mathbf {K}_1\mathbf {w}, \end{aligned}$$
(24)

where we have used the symmetry of \(\mathbf {K}_1\) and \(\mathbf {K}_2\) (cf. Definition (22)). Equation (24) implies that \(V(\mathbf {w})\) is a potential for the stiffness terms in Eq. (23) and hence condition (C2) holds.

As noted in Remark 3.3, system (23) can be modified such that condition (C3) holds. In the current setting, a truncation of the nonlinearity for large displacements is justified, since the partial differential equation (18) was obtained by expanding the Green–Lagrange strain tensor in a Taylor series up to order three. Such an approximation is generally only valid for small enough displacements.

To verify condition (C4), we note that Eq. (8) for the discretized beam (23) is given by

$$\begin{aligned} \mathbf {w}^\top \mathbf {S}(\mathbf {w})=\frac{1}{h^4} \mathbf {w}^{\top } \mathbf {K}_2 \mathbf {w} +\frac{1}{2 r^2 h^3} \mathbf {w}^{\top }\mathbf {K}_1\mathbf {w} \mathbf {w}^{\top } \mathbf {K}_1\mathbf {w}.\nonumber \\ \end{aligned}$$
(25)

Since \(\mathbf {K}_1\) is a tridiagonal Toeplitz matrix with positive eigenvalues (e.g., Gover [27]), it is positive definite. Thus, the second summand in Eq. (25) is positive for all displacements, i.e., \( \mathbf {w}^{\top }\mathbf {K}_1\mathbf {w} \mathbf {w}^{\top } \mathbf {K}_1\mathbf {w}>0\). To analyze the first term in Eq. (25), we split the matrix \(\mathbf {K}_2\) as follows

$$\begin{aligned}&\!\!\!\mathbf {K}_2=\mathbf {K}_1^2 +\mathbf {B}(1)+\mathbf {B}(N-4),\nonumber \\&\!\!\!B_{jl}(k):= {\left\{ \begin{array}{ll} 1 &{} l=j=k \\ 0 &{} \text {otherwise} \end{array}\right. }, \quad j,l=1,\ldots ,N-4\nonumber \\ \end{aligned}$$
(26)

where all entries of the matrices \(\mathbf {B}(1)\) and \(\mathbf {B}(N-4)\) are zero expect for the first, respectively, the last entry on the main diagonal which is equal to one. Both matrices (\(\mathbf {B}(1)\) and \(\mathbf {B}(N-4)\)) are positive semi-definite. Furthermore, we note that the square of the symmetric and positive definite matrix \(\mathbf {K}_1\) is also positive definite. Since \(\mathbf {K}_2\) can be expressed as sum of the positive definite matrix \(\mathbf {K}_1^2\) with the positive semi-definite matrices \(\mathbf {B}(1)\) and \(\mathbf {B}(N-4)\), it is positive definite. The positive definiteness of \(\mathbf {K}_2\), ensures that condition (C4) is satisfied for some \(\alpha >0\) and \(r=0\), since the second summand in Eq. (25) is always positive as we have noted earlier. In summary, all conditions (C1)–(C4) are met, and our Theorem 3.1 guarantees an arising quasi-periodic response of system (23) to external quasi-periodic forcing.

We calculate a forced response of the discretized beam (23) with the numerical package proposed by Jain et al. [35] and choose \(N=10\) node points for the spatial discretization of the partial differential equation (18). The arising system (23) has six degrees of freedom. Moreover, we select the parameter values \(r=0.01\), \(\mu =0.05\) and \(L=1\) and assume quasi-periodic forcing of the form \(f(x,t)=0.06 r^2 (\sin (2\pi x /L)+1) (\sin (\Omega _1 t)+\sin (\Omega _2 t))\). We sweep the forcing frequencies \(\Omega _1\) and \(\Omega _2\) close to the first two eigenfrequencies \(\omega _1\approx 7.13 ~\mathrm {rad/s} \) and \(\omega _2 \approx 19 ~\mathrm {rad/s} \) and depict the \(L_2\)-norm of the arising quasi-periodic steady-state response in Fig. 6.

Fig. 6
figure 6

Quasi-periodic steady-state responses of system (23) with six degrees-of-freedom and non-dimensional parameters \(\mu =0.05\), \(r=0.01\) and \(L=1\) and forcing \(f(x,t)=0.06 r^2 (\sin (2\pi x/L )+1) (\sin (\Omega _1 t)+\sin (\Omega _2 t))\). We select the forcing frequencies \(\Omega _1 \) and \(\Omega _2\) in resonance with the first two natural frequencies \(\omega _1\approx 7.13 \) and \(\omega _2 \approx 19 \)

The results depicted in Fig. 6 arise from two incommensurate frequencies. Our Theorem 3.1, however, guarantees the existence of a quasi-periodic steady-state response for any finite number of incommensurate frequencies. To illustrate the applicability of Theorem 3.1 in the case of more than two incommensurate forcing frequencies, we consider the excitation

$$\begin{aligned} f(x,t,\alpha )=0.05r^2(\sin (2\pi x /L)+1) \sum _{k=1}^3 \sin ( \gamma \omega _k t) \nonumber \\ \end{aligned}$$
(27)

where \(\omega _k\) denote the eigenfrequencies of the discretized beam (23). For \(\gamma \) close to one, the forcing (27) is in resonance with the first three eigenfrequencies. As previously, Theorem 3.1 guarantees the existence of a quasi-periodic steady-state response also for the forcing (27). Once again, we use the numerical package proposed by Jain et al. [35] and compute a quasi-periodic steady-state response. While the package handles two-dimensional tori fast and efficiently, its memory requirements become challenging and computation times increase significantly for three incommensurate frequencies.

We depict the maximal displacement of the arising quasi-periodic steady-state response of the discretized beam (23) for the forcing (27) in resonance with the first three eigenfrequencies in Fig. 7. Moreover, we include a solution by direct time integration with Matlab’s ode45 algorithm serving as a benchmark solution, which we initialize on the computed quasi-periodic orbit at \(t=0\). While the maximal vibration amplitude of the nonlinear response agrees well with the benchmark solution, the linear response differs significantly. Thus, a nonlinear analysis of system (23) is essential to accurately predict the response amplitudes.

Fig. 7
figure 7

Quasi-periodic steady-state responses of system (23) with three incommensurate base frequencies (\(K=3\)). We consider system (23) with six degrees-of-freedom, non-dimensional parameters \(\mu =0.05\), \(r=0.01\) and \(L=1\) and select the forcing (27) to be in resonance with the first three natural frequencies

5 Conclusions

We have presented a criterion guaranteeing the existence of a quasi-periodic response for wide class of externally driven multi-degree-of-freedom vibratory systems. Our Theorem 3.1 guarantees the existence of a quasi-periodic orbit for arbitrary large forcing and response amplitudes, whereby it overcomes the major shortcoming of competing perturbation results. Thus, it can provide a priori justification for formal perturbation methods or numerical methods such as harmonic balance. Thereby, the results of such heuristic approaches become valid and computational resources can be efficiently employed.

Our criterion requires the damping to act on all coordinates, the stiffness terms to be derived from a potential and satisfy a growths conditions far away from the origin. These conditions are similar to our results in the periodic case [8] except for the growth condition (C4) which is stricter compared to the sign condition in the periodic case. Moreover, the nonlinear terms are required to be globally Lipschitz, which is generally satisfied if Eq. (1) accurately model a physical system (cf. Remark 3.3). We also cover the case of discontinuous external forcing, such as square waves or saw-tooth functions (cf. Theorem 3.3).

We utilize our results to revisit a chain of oscillators investigated by Breunung and Haller [7] and a slender beam model proposed by Nayfeh and Mook [49]. Both systems satisfy the conditions of our Theorem 3.1 and hence an approximation of the solution in a Fourier series is justified. We compute the arising steady-state vibrations with the numerical package proposed by Jain et al. [35]. Moreover, we consider a complex loading scenario, where the vibrations are excited by the natural roughness of a road. In this scenario, quasi-periodic functions arise naturally and our Theorem 3.1 justifies subsequent computations with the harmonic balance procedure.

We have limited our discussion to position depending nonlinearities, since these are customary in the vibrations literature. For more complex damping forces or oscillatory systems derived from Lagrangian with position dependent mass matrix arising in, e.g., robotics, an extension to account for velocity dependent nonlinearities is desirable.

After computing a steady-state response, it is important to assess its stability, since only stable solutions are observable. Numerical calculation of the monodromy matrix and calculating its eigenvalues (often referred as Floquet theory) is a well-established and universal procedure to assess the stability of periodic orbits. To the best of our knowledge such a procedure is absent for quasi-periodic orbits.

Furthermore, we have focused on quasi-periodic solutions, since these type of vibrations have been reported for a variety of mechanical and even more general systems (cf. Introduction 1). From a theoretical perspective, an extension of our results to the case of almost-periodic or bounded solutions seems tempting.