1 Introduction

The nontrivial effect of fast excitation has been elaborately studied in recent years [24, 6, 7, 1214]. It is known that mechanical systems, under high frequency parametric excitation, can undergo apparent changes in system properties such as the number of equilibrium points, stability of equilibrium points, natural frequencies, stiffness, and bifurcation paths [14]. The method of direct partition of motion (DPM), formalized by Blekhman [2], serves to facilitate the study of such problems. DPM is most suitable to study motions which can be represented as the sum of a leading order slow component and an overlaid fast component. The fast component of motion is often only interesting in the extent that it affects the main slow dynamics. Unlike the averaging method or the method of multiple timescales, DPM offers no systematic way to obtain higher order terms in an asymptotic expansion of the solution, and instead is limited to the leading order dynamics of the system. In return for this limitation, one gains efficiency in terms of the required mathematical manipulations.

In most problems addressed in the literature, the fast excitation is due to an external source, that is, the system considered is non-autonomous. However, similar nontrivial effects could occur even if the fast excitation is internal to the system, instead of coming from an external source. An example of such a case would be a nonlinear oscillator coupled to a much faster oscillator [1, 9, 11, 1517]. In these latter autonomous systems with widely separated frequencies, the leading order dynamics, particularly the frequency, of the fast oscillator is unaffected by the slow oscillator. This latter condition allows the use of the standard method of averaging, that is, the fast oscillation is assumed to be, to leading order, a harmonic oscillation with a constant frequency. Then the equations of motion are averaged with respect to the fast timescale over a period of 2π [15]. This is not the case for the system we study in this paper, as the amplitude and frequency of the fast oscillation, to leading order, are found to be a function of the amplitude of the slow oscillation. This is established by observing that the equation of the fast degree of freedom can be treated as a fast oscillator with a slowly varying frequency, for which the WKB method is particularly suited, and thus using a transformation of fast time analogous to that proposed by the WKB method [19]. We present this work as an academic example that serves to shed light on the nontrivial dynamics that can arise in more general systems of this type, that is, systems with vastly different frequencies and nonlinear coupling that allows the slow variable to modulate the frequency of oscillation of the fast variable. Moreover, we use this example to illustrate how the strategy used here, which combines DPM with the WKB method, can be useful for the study of systems of this type. For simplicity, we restrict our attention here to the conservative case, ignoring dissipation and external forcing. We see this as a first basic step towards the understanding of the full dynamics that such systems are able to exhibit.

A well known variation of the mass–spring–pendulum system at hand is that in which the spring is constrained to move vertically instead of horizontally. It is known in the literature as a typical example in which autoparametric resonance can occur [18], that is, the system exhibits interesting dynamics if the ratio of the frequencies of the two degrees of freedom is 2:1 or 1:1; in such studies, the case in which the frequencies are widely separated is not given any attention. Another system similar to the one we study here is a mathematical model of two coupled Huygen’s clocks; the system represents two pendula hanging from a rigid beam support that is connected on one side to a wall through a linear spring [5]. Again in the latter system, the frequency of the linear support is considered to be of the same order of magnitude as that of the pendula. When such systems are addressed in the literature, the focus is often on the dynamics arising due to resonances, being internal, external or autoparametric resonances [8]. We emphasize that the dynamics addressed in this work is not due to any of the known types of resonance, that is, the frequencies of the two modes need not be commensurate. The prerequisite condition for the interaction addressed here is that the frequencies be of different orders of magnitude. Also, the direct modulation of the frequency and amplitude of the fast oscillation by the slow one is another feature of this interaction that is not present in the case of ordinary resonance.

In Sect. 2, we present the system of equations representing the mass–spring–pendulum that we consider, and the scaled equations that correspond to the relevant regime of small motions of the mass–spring oscillator; we also illustrate the nontrivial solutions that the system exhibits. Appendix A explains the motivation for the proposed form of solution while Sect. 3 presents the end result of the DPM procedure in the form of an autonomous equation governing the leading order slow oscillation; we also present the approximate expression of the leading order fast oscillation. The details of obtaining the latter are presented in Appendix C, while the DPM implementation is detailed in Appendix B. In Sect. 4, we discuss the bifurcation that occurs in the slow dynamics and it’s relation to the bifurcation in a Poincare map of the full system (1). The details of the analysis of the slow dynamics and the predictions based on it are presented in Appendices D and E. Section 5 briefly summarizes the main results. Then, in Sect. 6, we compare the approximate solution to that from numerical integration of the full system (1) and try to check the validity of the predictions we make based on the approximate solution.

2 The mass–spring–pendulum system

We consider a simple pendulum whose point of suspension is connected to a mass on a spring that is restricted to move horizontally, as shown in Fig. 1. Ignoring dissipation, the system is governed by the following equations of motion:

where primes denote differentiation with respect to time τ. We introduce the following change of variables:

$$x = \frac{{\tilde{x}}}{l},\qquad t = \sqrt{\frac{g}{l}} \tau . $$

The nondimensionalized equations, which we refer to as the full system, become:

$$ \everymath{\displaystyle }\begin{array}{@{}l} \ddot{\theta}+ \sin\theta= - \ddot{x}\cos\theta, \\ \noalign {\vspace {5pt}} \ddot{x} + \tilde{\varOmega}^2 x = - \mu\bigl( {\ddot{\theta}\cos\theta- \dot{\theta}^2 \sin\theta} \bigr), \end{array} $$
(1)

where \(\mu =\frac{m}{{M + m}}\), \(\tilde{\varOmega}^{2} =\frac{{kl}}{{g ( {M + m} )}}\), and dots represent differentiation with respect to time t.

Fig. 1
figure 1

Schematic for the mass–spring–pendulum system

2.1 Assumptions

We are interested in the case where the linear oscillator has a natural frequency that is an order of magnitude larger than the linearized frequency of the pendulum, and its motion has an amplitude that is an order of magnitude smaller than that of the pendulum. This is implemented through the following rescaling:

$$x=\varepsilon\chi,\qquad \tilde{\varOmega}^2 =\frac{{\varOmega^2}}{{\varepsilon^2 }}. $$

Here, Ω and χ are O(1) quantities while ε≪1. Without loss of generality, we take Ω=1. The rescaled equations become:

$$ \everymath{\displaystyle }\begin{array}{@{}l} \ddot{\theta}+ \sin\theta= - \varepsilon\ddot{\chi}\cos\theta, \\ \noalign {\vspace {5pt}} \ddot{\chi}+ \frac{1}{{\varepsilon^2 }} \chi= - \frac{\mu }{\varepsilon} \bigl( { \ddot{\theta}\cos\theta- \dot{\theta}^2 \sin\theta} \bigr). \end{array} $$
(2)

This system has a conserved quantity that can be expressed as:

2.2 Typical solutions

We numerically integrate the full system in Eq. (1) for typical parameter values and initial conditions (ICs), in order to illustrate the type of nontrivial solutions that it exhibits. As an example, we take μ=0.4 and \(\tilde{\varOmega}=50\) (ε=0.02). Since the system is conservative, we will look at how the dynamics change as the energy is increased.

For h=0.5, we choose the following set of initial conditions:

$$ \begin{array}{@{}l} \left \{\begin{array}{l} \dot{\theta}( 0 ) = 0, \\ \noalign {\vspace {3pt}} \theta ( 0 ) = \pi/9 \approx0.349, \\ \end{array} \right .\\ \noalign {\vspace {6pt}} \left \{\begin{array}{l} \dot{x} ( 0 ) = 0, \\ \noalign {\vspace {3pt}} x ( 0 ) = \varepsilon ( {{0.9756}} ) \approx0.0195. \\ \end{array} \right . \end{array} $$
(3)

For the ICs in Eq. (3), the pendulum oscillates about the downright position θ=0, it’s motion consists of a slow oscillation overlaid with a small fast oscillation, as shown in Fig. 2(a); Fig. 2(b) shows the mass–spring fast oscillation while Fig. 3 shows how its amplitude is modulated over the slow timescale.

Fig. 2
figure 2

Plot of time series for the initial conditions in Eq. (3)

Fig. 3
figure 3

Variation of the amplitude of the x oscillation over the slow timescale

Figure 6(a) shows the Poincare map \((x=0,\,\dot{x}>0 )\) for the energy level h=0.5, and the arrow points at the orbit corresponding to the ICs in Eq. (3). The shown fixed point (center) of the map corresponds to a periodic orbit in which θ≈−x. This periodic orbit is a nonlinear normal mode [10] of the coupled system that appears as a nearly straight line through the origin if viewed in the configuration plane θ vs. x. Now we look at the dynamics for h=0.7, and choose the following set of initial conditions:

$$ \begin{array}{@{}l} \left \{\begin{array}{l} \dot{\theta}( 0 ) = 0, \\ \noalign {\vspace {3pt}} \theta ( 0 ) = \pi/9 \approx0.349, \\ \end{array} \right . \\ \noalign {\vspace {5pt}} \left \{\begin{array}{l} \dot{x} ( 0 ) = 0, \\ \noalign {\vspace {3pt}} x ( 0 ) = \varepsilon ( {{1.1626}} ) \approx0.0233. \\ \end{array} \right . \end{array} $$
(4)

Figure 4(a) shows how the slow oscillation, overlaid by a small fast oscillation, is now no longer about the origin. Instead, the pendulum oscillates about an angle ≈0.33. The amplitude of the fast mass–spring oscillation is still slowly modulated, as shown in Fig. 3.

Fig. 4
figure 4

Plot of time series for the initial conditions in Eq. (4)

Keeping the energy fixed at h=0.7, we choose a different set of ICs:

$$ \begin{array}{@{}l} \left \{ \begin{array}{l} \dot{\theta}( 0 ) = 0, \\ \noalign {\vspace {3pt}} \theta ( 0 ) = \pi/6 \approx{0.5236,} \\ \end{array} \right . \\ \noalign {\vspace {5pt}} \left \{\begin{array}{l} \dot{x} ( 0 ) = 0, \\ \noalign {\vspace {3pt}} x ( 0 ) = \varepsilon ( {{1.1370}} ) \approx{0.0227.} \\ \end{array} \right . \end{array} $$
(5)

As shown in Fig. 5(a), the pendulum is back to oscillating about the downright position and the modulation of the amplitude of x (Fig. 3) is still visible.

Fig. 5
figure 5

Plot of time series for the initial conditions in Eq. (5)

Figure 6(b) shows the Poincare map for h=0.7. The arrows point to the orbits corresponding to the solutions in Fig. 4(a) (pointed arrow) and Fig. 5(a) (square head arrow). We can see from the Poincare map that the fixed point corresponding to the nonlinear normal mode with θ≈−x has lost stability and is now a saddle point of the map. Consequently, we can predict that two new fixed points (centers) were born in the process, and that the oscillations of the pendulum about a nonzero angle correspond to closed orbits of the map about the new centers.

Fig. 6
figure 6

Poincare map (x=0, \(\dot{x} >0\)) for (a) h=0.5, (b) h=0.7

The aim of this paper is to shed light on these latter nontrivial solutions, in which the pendulum oscillates about a nonzero angle, and describe their dependence on initial conditions and the parameter μ.

3 The approximate solution

We look for a solution in which the oscillation of the slow degree of freedom (the pendulum) is partitioned according to the method of direct partition of motion [2]:

$$ \left \{\begin{array}{l} \chi=\chi ( {t,T} ),\\ \noalign {\vspace {3pt}} \theta ( {t,T} ) = \theta_0 ( t ) + \varepsilon\theta_1 ( {t,T} ), \\ \end{array} \right . $$
(6)

where

$$ \frac{{dT}}{{dt}} = \frac{{\omega ( t )}}{\varepsilon}, \quad \mathrm{or}\quad T = \int_0^t {\frac{{\omega ( {t'} )}}{\varepsilon}dt'}, $$
(7)

and

$$ \omega( t ) = \frac{1 }{{\sqrt{1 - \mu\cos^2 \theta_0 } }}. $$
(8)

Here, we have introduced a new fast timescale, T, in a way similar to the WKB method [19]; the choice of ω(t) is justified in Appendix C. We note that we do not apply DPM to the fast degree of freedom since it is under the influence of a slower oscillator, while DPM is only applicable to a slow oscillator that is influenced by a much faster oscillation (as in the case of the pendulum in this system). Thus, as detailed in Appendix C, the WKB method is instead needed for the analysis of the fast degree of freedom.

After applying the standard DPM procedure [2] to the θ equation of motion, we find that, to leading order, θ 0 is governed by the following equation (see Appendix B for the details):

(9)

θ 1 is found to depend on θ 0 and χ as follows:

$$ \theta_1 = - \chi\cos\theta_0. $$
(10)

To leading order, χ is given by (see Appendix C for the details):

$$ \chi\approx C\sqrt{\omega( t )} \cos T, $$
(11)

where C is an arbitrary constant that depends on initial conditions.

Consequently, the motion of the pendulum, in the rescaled system described by Eq. (2), can be expressed as:

$$\theta\approx\theta_0 - \varepsilon\chi\cos \theta_0, $$

where θ 0 is governed by Eq. (9) and χ is given by Eq. (11).

Recall that Eq. (2) are a rescaled version of the original system of interest given by Eq. (1), where χ is related to the motion of the mass–spring oscillator as follows:

$$x = \varepsilon\chi. $$

Hence, the solution to Eq. (1), for the assumed regime of motion, can be expressed in terms of the variables of Eq. (1) as follows:

$$ \theta\approx\theta_0 - x \cos\theta_0, $$
(12)
$$ x \approx\varepsilon C\sqrt{\omega( t )} \cos T. $$
(13)

4 The slow dynamics

At the end of the procedure that is described in Appendices B and C, the solution to the two degree of freedom mass–spring–pendulum system is expressed in Eq. (12)–(13) in terms of θ 0, the leading order slow motion of the pendulum, which is governed by the following equation:

The arbitrary constant C that appears in the equation can be expressed in terms of the initial conditions. For initial zero velocities, the initial conditions take the form:

with

(14)

Now, we rewrite the equation governing θ 0 as a system of two first order equations:

$$ \everymath{\displaystyle }\begin{array}{@{}l} {\dot{\theta}_0 = \phi,}\\ \noalign {\vspace {8pt}} \dot{\phi}= - \sin\theta_0 + \frac{{1}}{2}C^2 \frac{\sin\theta_0 \cos\theta_0}{ (1-\mu\cos^2\theta_0 )\sqrt{1- \mu\cos^2\theta_0}}. \end{array} $$
(15)

For small enough values of C, the above system has a neutrally stable equilibrium point (center) at the origin (ϕ=0,θ 0=0) and two saddle points at (ϕ=0, θ 0π), so that the phase portrait resembles that of the simple pendulum. As C increases in value, a pitchfork bifurcation takes place, in which the origin becomes a saddle point and two new centers are born. The critical value of C is related to the parameter μ as follows (see Appendix D for the details):

$$ C_{cr}^2 = 2 ( {1 - \mu} )^{\frac{3}{2}} . $$
(16)

In Appendix D, it is explained how this condition on C translates into the following condition on the energy value h:

$$h_{cr}=1-\mu. $$

For the example presented in Sect. 2.2, the critical values of C and h are as follows:

So a qualitative change in the solution is expected as h increases past h=0.6, which explains the difference in solution between h=0.5 and h=0.7, cf. Figs. 6(a)–6(b).

To illustrate the relation of the solution of the full system (1) to the θ 0 dynamics, we find the value of C for the ICs presented in Sect. 2.2.

For the ICs in Eq. (3) corresponding to Fig. 2, we have:

Figure 7(a) shows the corresponding phase portrait for the system in Eq. (15) with this value of C. The slow oscillation of the pendulum in Fig. 2(a) corresponds to the closed orbit surrounding the origin in this phase portrait.

Fig. 7
figure 7

Phase portrait for the θ 0 equation for (a) ICs in Eq. (3) corresponding C=0.8749, (b) ICs in Eq. (4) corresponding C=1.0427, (c) ICs in Eq. (5), corresponding to C=1.0400

For the ICs in Eq. (4) corresponding to Fig. 4, we have:

For this value of C, Fig. 7(b) shows that the origin is a saddle point and the system in Eq. (15) has two nontrivial neutrally stable equilibrium points (centers). The slow oscillation of the pendulum in Fig. 4(a) corresponds to the small closed orbit surrounding one of the nontrivial equilibrium points in this phase portrait.

For the same energy levels, different ICs result in different values of C and different orbits in the resulting phase plane. For the ICs in Eq. (5), we get:

The resulting slow oscillation in Fig. 5(a) corresponds to the closed orbit enclosing the homoclinic orbit in the phase portrait shown in Fig. 7(c). This illustrates how, despite the presence of the two nontrivial equilibrium points, oscillations about the origin are still possible and correspond to large amplitude orbits that are outside the homoclinic orbit.

4.1 The predicted nonlinear normal modes

Note that each value of C leads to a phase portrait filled with closed orbits, however, out of those orbits, the only one which corresponds to a solution of the full system (1) is that associated with the specific ICs that led to that value of C.

An interesting case occurs when the choice of ICs results in a phase portrait that has a nontrivial equilibrium point which coincides in value with the initial θ 0 amplitude, A. That is, we start with ICs of the form:

$$\left \{\begin{array}{l} \dot{\theta}( 0 ) = 0, \\ \noalign {\vspace {3pt}} \theta ( 0 ) = \theta_0 ( 0 ) = A, \\ \end{array} \right . \qquad \left \{\begin{array}{l} \dot{x} ( 0 ) = 0, \\ \noalign {\vspace {3pt}} x ( 0 ) = \varepsilon B, \\ \end{array} \right . $$

and the corresponding value of C results in nontrivial equilibrium points (centers) for the θ 0 equation at:

$$\theta_0 = \pm E,\qquad \phi=0. $$

Then, if E=A, θ 0 will remain equal to E for all time. It would mean that we are starting at a neutrally stable equilibrium point of the θ 0 equation, so the solution will remain at that point for all time.

Appendix D shows that these special values of initial θ amplitude can be expressed in terms of h and μ as:

(17)

The corresponding value of x is expressed as:

$$ x ( 0 ) = \varepsilon B^* = \varepsilon\sqrt{2 \bigl( {h - \mu \bigl( {1 - \cos A^*} \bigr)} \bigr) }. $$
(18)

Hence, we predict that these special initial amplitudes, with zero initial velocities, will lead to a solution in which:

$$ \theta\approx A^* -x\cos A^*. $$
(19)

Such a solution would be a nonlinear normal mode of the coupled mass–spring–pendulum system.

4.2 Relation of θ 0 to the Poincare map

For a given energy level, the phase portrait of the θ 0 equation is filled with closed orbits and the picture is topologically similar to that of the Poincare map. That is, for a given initial condition, the resulting orbit in the θ 0 phase plane corresponds to a closed orbit in the Poincare map, however, the orbits are not identical. This is due to the fact that, while θθ 0, \(\dot{\theta}\) differs from \(\dot{\theta}_{0}\) by an O(1) quantity; as shown in Appendix E, for the points of the Poincare map, \(\dot{\theta}\) can be expressed in terms of \(\dot{\theta}_{0}\) as follows:

$$ \dot{\theta}_{Pm} \approx\dot{\theta}_0 - C \bigl( {1 - \mu\cos^2 \theta_0 } \bigr)^{ - \frac{3}{4}} \cos\theta_0. $$
(20)

So for a given ICs, we can obtain the corresponding orbit in the Poincare map, by first numerically integrating the θ 0 equation to obtain θ 0 and \(\dot{\theta}_{0}\) and then generating the orbit in the Poincare map by plotting the corresponding values of \(\dot{\theta}_{Pm}\) vs. θ 0. This means that we can generate an approximate picture of the Poincare map of the full system (1) by numerically integrating the slow dynamics equation governing θ 0 instead of integrating the full system (1) which contains the fast dynamics and thus requires a much smaller step size of integration.

Also, by comparing this procedure with the results of numerical integration, we can obtain a check on the accuracy of the various approximations made in this work.

5 Summary of results

We restate here the original equations governing the considered mass–spring–pendulum system:

We have shown that θθ 0+O(ε) where θ 0 is governed by the following equation:

where C is a constant that depends on the ICs. x is assumed to be O(ε) and is found to be expressed as:

$$x \approx\varepsilon C\sqrt{\omega( t )} \cos T $$

with

$$\omega( t ) = \frac{1}{{\sqrt{1 - \mu\cos^2 \theta_0 } }} \quad \mathrm{and}\quad T = \int _0^{2\pi} {\frac{{\omega ( {t'} )}}{\varepsilon}dt'}. $$

Our analysis gives that a pitchfork bifurcation occurs in the Poincare map as energy increases past the following critical value:

$$h_{cr}=1-\mu. $$

This bifurcation corresponds to a bifurcation in periodic orbits of the full system (1) in which the nonlinear normal mode corresponding to θ≈−x loses stability and two new stable nonlinear normal modes are born in which θA xcosA , where A is expressed in terms of h and μ as:

$$A^* = \pm\cos^{ - 1} \biggl( {\frac{{\mu - h \pm\sqrt{ ( {h - \mu} )^2 + 8\mu} }}{{4\mu}}} \biggr). $$

These new nonlinear normal modes correspond to the nontrivial fixed points of the Poincare map.

6 Comparison to numerics

We compare the solution resulting from the numerical integration of the original equations with that from the integration of the θ 0 equation. Figure 8 displays the Poincare map orbits for h=1. Near each of the orbits, a small arrow points to the orbit which is predicted from the θ 0 equation for corresponding initial conditions. Figures 9, 10 and 11 display comparison plots for several initial conditions. Unless otherwise mentioned, we have set μ=0.4 and \(\tilde{\varOmega}=50\) (ε=0.02). In the plots of θ vs. time, the thick line correspond to the solution of the numerical integration of the full system (1), and the apparent thickness is due to the fast component present in the oscillation of the pendulum; the thin line corresponds to the approximate solution, that is, from the numerical integration of the θ 0 equation, and captures only the leading order slow component of the pendulum oscillation. In the plots of x vs. time, the arrow points at the approximate solution.

Fig. 8
figure 8

Comparison of the predicted Poincare map orbits (arrows) with those from the integration of the full system (1)

Fig. 9
figure 9

Comparison plot of time series for ICs in Eq. (3)

Fig. 10
figure 10

Comparison plots of time series for ICs in Eq. (4)

Fig. 11
figure 11

Comparison plots of time series for ICs in Eq. (5)

We can see that the approximate solution compares well with that from numerical integration of the full system (1).

7 Conclusion

We have used the method of direct partition of motion to study the dynamics of a mass–spring–pendulum system in which the harmonic oscillator is restricted to move horizontally. We have considered the case where the stiffness of the spring is very large, so that the frequency of the oscillation of the uncoupled harmonic oscillator is an order of magnitude larger than that of the uncoupled pendulum. We have also limited our attention to the regime of motion where the amplitude of motion of the harmonic oscillator is an order of magnitude smaller than that of the pendulum. Under these assumptions, an approximate expression for the solution of the two degree of freedom system is found in terms of θ 0, the leading order slow oscillation of the pendulum. An equation governing θ 0 is presented and found to undergo a pitchfork bifurcation for a critical value of C which is a parameter related to the initial amplitudes of θ and x. It is shown that the pitchfork bifurcation in the slow dynamics equation corresponds to a pitchfork bifurcation of periodic orbits of the full system (1) that occurs as the energy is increased past a critical value which is expressed in terms of the parameter μ. This bifurcation can be seen to occur in the Poincare map of the full system (1), where the fixed point corresponding to the nonlinear normal mode θ≈−x loses stability and two new centers are born in the map. The new centers correspond to new periodic motions, which are nonlinear normal modes with θA xcosA , where the expression for A is found in terms of μ and h. For these modes, the motion of the pendulum is predicted to be a small fast oscillation about the nonzero value θ=A . Along with these special motions, quasi-periodic motions exist in which the pendulum undergoes slow oscillation about a nonzero angle, with overlaid fast oscillation. These latter orbits correspond to closed orbits about the new centers in the Poincare map. A relation between \(\dot{\theta}\) and \(\dot{\theta}_{0}\) is given for points of the Poincare map such that the orbits of the map can be generated approximately by numerically integrating the slow dynamics equation. Finally, the approximate solution, as well as the predications made based on it, is checked against numerical integration of the full system (1) and found to agree well.

Although this paper dealt with a specific system, we suggest that the strategy used to study this system, that is, DPM in combination with the WKB method, could be applicable to a general class of systems that posses two degrees of freedom with vastly different frequencies and nonlinear coupling that allows the modulation of the fast oscillation by the slow one.

As a final note, we emphasize that the nontrivial dynamics observed in this example system was primarily due to the nonlinear interaction between oscillations of vastly different frequencies. Particularly, these oscillations represented the natural response of the system to initial conditions in the absence of any dissipation or external forcing. However, preliminary numerical integration of the corresponding system with linear damping and external harmonic forcing clearly shows solutions that are qualitatively similar to those exhibited by the conservative system. These latter solutions consist of slow oscillation of the pendulum about a nonzero angle accompanied by fast oscillation of the mass on the spring with an amplitude that is modulated on the slow timescale. Such solutions are also observed in the free response of the corresponding system in which the fast degree of freedom is subject to nonlinear damping (such as a van der Pol type nonlinearity) that allows it to undergo sustained fast oscillations. Detailed analysis of such non-conservative versions of the example studied here is beyond the scope of this work and is left for future investigations.