Abstract

For two combinations of a dimensionless rotational damping parameter and a dimensionless inertial coupling parameter, we consider free response of a rectilinearly vibrating linearly sprung primary mass inertially coupled to damped rotation of a second mass, for which Gendelman et al. (2012, “Dynamics of an Eccentric Rotational Nonlinear Energy Sink,” ASME J. Appl. Mech. 79(1), 011012) developed equations of motion in the context of a rotational nonlinear energy sink (NES) with no direct damping of the rectilinear motion. For dimensionless initial rectilinear displacements comparable with those considered by Gendelman et al., we identify a region in the motionless projection of the initial condition space (i.e., for zero values of the initial rectilinear and rotational velocities) in which every initial condition leads to a previously unrecognized zero-energy solution, with all initial energy dissipated by rotation. We also show that the long-time nonrotating, rectilinear solutions of the type found by Gendelman et al. are (orbitally) stable only in limited ranges of amplitude. Finally, we show how direct viscous damping of rectilinear motion of the primary mass affects dissipation, and that results with no direct rectilinear dissipation provide excellent guidance for performance when direct rectilinear dissipation occurs. Some applications are discussed.

1 Introduction

Inertial coupling of rectilinear motion of a linearly sprung primary mass to viscously damped rotation of a second mass about a vertical axis can lead to complicated dynamics, and “targeted energy transfer” in which kinetic energy of the rectilinear motion is transferred to rotational motion, which is then dissipated [13]. In such a rotational “nonlinear energy sink” (NES), the second mass can rotate at any angular speed, with no “preferred” frequency, and so a properly designed system can be expected to extract and dissipate energy from the rectilinear motion of the primary structure over a range of natural frequency of the latter. The approach has been demonstrated in experiments [1], and in simulations of flow-induced vibration (FIV) of a circular cylinder [46] and impulsively driven structural vibration [7].

For a point mass attached to and allowed to rotate about a vertical axis perpendicular to and translating with the rectilinear motion of a linearly sprung primary mass, Gendelman et al. [1] developed equations of motion
(M+m)d2xdt2+Cx=mr0ddt[dθdtsinθ]
(1a)
mr02d2θdt2+γdθdt=mr0d2xdt2sinθ
(1b)
where x is the rectilinear displacement of the primary mass M in the lab frame relative to its equilibrium position, θ is the angular position of the point mass m rotating about the vertical axis at a radius r0, C is a linear spring constant, and γ is a viscous damping coefficient. On the right-hand side (RHS) of Eq. (1a), Sigalov et al. [2,3] introduced a term ζdx/dt to account for (direct) viscous damping of the rectilinear motion of the primary mass, and nondimensionalized the resulting equations as
d2ηdτ2+λdηdτ+η=σddτ[dθdτsinθ]
(2a)
d2θdτ2+αdθdτ=d2ηdτ2sinθ
(2b)
where in our notation η = x/r0 and τ=tC/(M+m), and the dimensionless parameters are a mass ratio σ = m/(M + m), a rotational damping coefficient α=γ/[mr02C/(m+M)], and a rectilinear damping coefficient λ=ζ/C(M+m). We specify the initial conditions as
η(0)=ηi,dη(0)/dτ=vi,θ(0)=θi,anddθ(0)/dτ=Ωi(3ad)
(3a–d)

With λ = 0, Eqs. (2a,b) have also been shown to apply if the rotating mass is distributed [4], with the distance from its center of mass to the axis (Rcm) allowed to differ from the radius of gyration (Rg). In this case, σ = m(Rcm/Rg)2/(m + M), and α=γ/[mRg2C/(m+M)]. As shown in Sec. A available in the Supplemental Materials, Eqs. (2a,b) also describe the situation in which a single linearly sprung, rigid, distributed mass undergoes viscously damped rotation about a vertical axis which does not pass through its center of mass, and which moves horizontally with the rectilinear vibration of the body. Because Eqs. (2a,b) apply equally well to these other cases (for which σ is not a mass ratio), we hereafter refer to σ as a dimensionless coupling parameter.

For (α, σ, λ) = (0.1, 0.1, 0), Gendelman et al. [1] presented numerical solutions of Eqs. (2a,b) for six different motionless initial conditions (MICs, with vi = Ωi = 0), and in each case found that the long-time solution was one for which the primary mass oscillates harmonically with a nonzero amplitude, with the angular position an integer multiple of π. Sigalov et al. [2,3] subsequently considered σ = 0.1 with α = 0.05 and 0.1, again with MICs, and for λ = 0 reported a range of dynamical behavior, including a sequence of transitions between regular motion and chaos. They also stated [3] that computations with σ = 0.03 and α = 0.2 produced similar results. They presented no results for λ > 0, but stated that computations for λ > 0 were “similar” to those for λ = 0 and were “omitted for the sake of brevity”. Here, we reconsider the same dynamical equations studied previously [13], with three main objectives.

First, we show that for λ = 0, the long-time solutions include not only the “semi-trivial” [8] solutions
(η(τ),θ(τ))=(Bsin(τ+φ),nπ),naninteger
(4)
with B > 0 found previously [13], but also a previously unrecognized class of zero-energy solutions in which the initial energy has been completely dissipated. For two combinations of α and σ, including α = σ = 0.1 [13], we then establish a range of initial conditions in the MIC space for which complete dissipation of initial energy actually occurs. Notwithstanding the statement [2] that the initial angular position “θ0 is expected to have little effect on the dynamics as long as θ0 for any integer n,” we show that the range of initial energy for which complete dissipation occurs depends strongly on the initial angular position.

Second, for λ = 0, we show that the semi-trivial solutions, with harmonic rectilinear motion and no rotation, are orbitally stable only in limited ranges of amplitude and are unstable for all other amplitudes.

Third, we add damping of the primary mass, corresponding to λ > 0, in which case the only long-time solutions are zero-energy, regardless of the initial condition. We show that the time required to dissipate 99% of the initial energy correlates well with the predictions of the λ = 0 analysis, provided that the damping of the primary mass is weak.

The remainder of the paper is organized as follows. For λ = 0, we identify two classes of long-time solutions in Sec. 2, where we briefly describe the numerical methods used, and present results of a Floquet-based stability analysis, showing that for each α, the semi-trivial solutions are orbitally stable only in finite ranges of amplitude and are unstable for all other amplitudes. In Sec. 3, we present representative trajectories for several MICs and delineate the boundaries between three distinct regions in the MIC space, including one in which every initial condition leads to a zero-energy solution, with the initial energy completely dissipated by the NES. This is followed by a discussion of trajectories that are (a) asymptotically motionless (Sec. 4), or (b) asymptotically semi-trivial (Sec. 5). In Sec. 6, we consider the effects of viscous damping of the rectilinear motion of the primary mass, followed by a discussion in Sec. 7, and conclusions in Sec. 8.

2 Preliminary Considerations

2.1 Nature of the Long-Time Solutions.

All solutions of Eqs. (2a,b) reported previously [13] were computed for λ = 0 and are of the form shown in Eq. (4). In addition to those solutions, it is evident that Eqs. (2a,b) have zero-energy solutions with η(τ) = 0 and θ(τ) = θ, where θ is arbitrary.

That these are the only long-time solutions is easily shown by considering the dimensionless total energy, which can be written as
E(τ)=12[η2+(dηdτ)22σdηdτdθdτsinθ+σ(dθdτ)2]
(5)
Using Eqs. (2a,b), we can show that
dEdτ=λ(dηdτ)2ασ(dθdτ)2
(6)

It is clear that the long-time solutions must correspond to constant values of θ (denoted by θ). (In the alternative, the RHS of Eq. (6) shows that E must decrease, showing that no long-time solution is possible unless θ is constant.) For λ > 0, every long-time solution must have zero energy, with (η(τ), θ(τ)) = (0, θ), again with θ arbitrary.

2.2 Numerical Methods.

For numerical integration, we used the matlab code ode45, which implements a (4,5) Runge–Kutta method, with a step size adjustable to satisfy specified error tolerances. The absolute and relative error tolerances, taken to be equal here, and referred to as δ, are set to 10−8. For the values of ηi considered here, the results are insensitive to small changes in initial conditions, and this tolerance yields results sufficiently converged for the intended purpose.

To efficiently extract the final amplitudes, we fit each trajectory to its asymptote ηasymp(τ) = B sin(τ + φ) (sometimes with B = 0) and estimate the constants φ and
B=limτB(τ)Bi=ηi2+vi2
(7)
where B(τ)=η2+(dη/dτ)2, by applying the Levenberg–Marquardt algorithm [9] to perform a nonlinear least-squares fit using at least the last 200 dimensionless time units of the time series of η(τ) (with asymptotic period 2π). Except where otherwise specified in Sec. 3, we used 50 Levenberg–Marquardt iterations.

2.3 Floquet Analysis of the Stability of Semi-trivial Solutions for λ = 0.

For λ = 0, the stability of a semi-trivial solution with respect to infinitesimal disturbances can be assessed by substituting θ(τ)=nπ+θ*(τ) and η(τ)=Bsin(τ+φ)+η*(τ) into Eqs. (2a,b) and linearizing to get
d2η*dτ2+η*=0
(8a)
d2θ*dτ2+αdθ*dτ+B(1)nsin(τ+φ)θ*=0
(8b)
with the objective being to determine the ranges of B for which infinitesimal disturbances to the semi-trivial solution grow or decay. We define s = τ + φ + [1 − (−1)n] π/2, and rewrite Eqs. (8a,b) as
d2η*ds2+η*=0
(9a)
d2θ*ds2+αdθ*ds+Aθ*sins=0
(9b)
where A = B, and note that Eqs. (9a,b) are uncoupled, with no dependence on σ.
Although Eq. (9b) can be transformed into the Mathieu equation using θ*=eαs/2H(s) and s=2s^+3π/2, it is not possible to determine the stability of solutions of Eqs. (2a,b) by reference to the stability boundary for the Mathieu equation. To use results from the Mathieu equation, one would need to know the locus in the αA plane on which the growth rate is α/2. While this can certainly be done, we instead conduct a Floquet analysis, writing Eqs. (9a,b) as a system of four first-order ordinary differential equations (ODEs)
dη*ds=v*,dv*ds=η*,dθ*ds=Ω*,anddΩ*ds=αΩ*Aθ*sins(10ad)
(10a–d)
and then integrating for four sets of initial conditions
ηj*(0)=δ1j,vj*(0)=δ2j,θj*(0)=δ3j,Ωj*(0)=δ4j,1j4(11ad)
(11a–d)
over 0 ≤ s ≤ 2π, where δlm is the Kronecker delta, to form the matrix
G=[η1*(2π)η2*(2π)η3*(2π)η4*(2π)v1*(2π)v2*(2π)v3*(2π)v4*(2π)θ1*(2π)θ2*(2π)θ3*(2π)θ4*(2π)Ω1*(2π)Ω2*(2π)Ω3*(2π)Ω4*(2π)]
(12)
It is easy to see that
G=[1000010000θ3*(2π)θ4*(2π)00Ω3*(2π)Ω4*(2π)]
(13)
from which it is evident that the four eigenvalues of G are 1, 1, and the eigenvalues of
H=[θ3*(2π)θ4*(2π)Ω3*(2π)Ω4*(2π)]
(14)

The nature of the repeated root at 1 means that no semi-trivial solution is asymptotically stable, even for infinitesimal disturbances. From a physical standpoint, we see that any initial condition with [θimodπ]2+Ωi20 leads to dissipation, so that the final amplitude of the rectilinear motion will always be less than the initial value. If one or both eigenvalues of H lie outside the unit circle, then the semi-trivial solution is unstable, whereas if both eigenvalues lie inside the unit circle, the semi-trivial solution is orbitally stable, in the sense that every infinitesimal initial disturbance leads to only an infinitesimal departure from the nominal solution. The case of instability is separated from the case of orbital stability by what we will call the stability boundary, corresponding to the situation where the eigenvalues of H lie on the unit circle.

Figure 1 shows a multi-valued stability boundary, where critical values lie on branches (denoted by Ci) terminating on the A-axis and connected at turning points (denoted by Fi), with the first two turning points (F1 and F2) and first five branches (C0, C1, C2, C3, and C4) shown in Fig. 1. Consecutive branches (C2k−1 and C2k, k ≥ 1) bound “tongues” (denoted by Tk), each extending as a cusp to the A-axis, reminiscent of the tongues of instability for the Mathieu equation [10]. That is not surprising, in light of the fact that our boundaries correspond to the boundaries for which the Mathieu equation is unstable with a (positive) growth rate of α/2.

Fig. 1
Stability boundary for solutions of Eqs. (2a,b) with λ = 0. “U” and “S” denote unstable and orbitally stable portions of the α − A plane, respectively. The first two turning points are denoted by F1 and F2. Each point on a branch C2k (x) or C2k+1 (○) is the midpoint between a point for which the time-periodic solution is unstable and a nearby point for which it is orbitally stable.
Fig. 1
Stability boundary for solutions of Eqs. (2a,b) with λ = 0. “U” and “S” denote unstable and orbitally stable portions of the α − A plane, respectively. The first two turning points are denoted by F1 and F2. Each point on a branch C2k (x) or C2k+1 (○) is the midpoint between a point for which the time-periodic solution is unstable and a nearby point for which it is orbitally stable.
Close modal

For λ = 0, the stability boundary separates combinations of α and A (and hence B) for which semi-trivial solutions of Eqs. (2a,b) are orbitally stable (denoted by “S”), from those for which disturbances grow (denoted by “U”). For α > 0, the number of critical values lying below any A is a nonincreasing function of α, decreasing by two at each passage of a constant-α line beyond a turning point. Table 1 shows the first five critical values of Ak(α) for the values of α considered here. We note that A2kA2k−1 (i.e., the tongue “width”) decreases rapidly with increasing k and with decreasing α.

Table 1

Critical values of Ak as a function of α

α
0.11
A00.46381.18341
A13.76174.2030
A23.79614.406423
A310.654811.1843
A410.657111.2021
α
0.11
A00.46381.18341
A13.76174.2030
A23.79614.406423
A310.654811.1843
A410.657111.2021

These results are consistent with the numerical simulations of Gendelman et al. [1], in which all reported long-time solutions corresponded to harmonic rectilinear motion of the primary mass, with the angular position of the rotating mass being an integer multiple of π. (The zero-energy alternative B = 0, discussed in Secs. 3 and 4, was not recognized in Refs. [13].)

3 Distinct Regions in the Motionless Initial Condition Space Without Rectilinear Dissipation

For λ = 0, we show here that different MICs can lead to qualitatively different long-time solutions, including solutions from which all initial energy is dissipated, a result not reported in the previous work [13].

For (α, σ, λ) = (0.1, 0.1, 0) [13] and (1, 0.01, 0), we restrict consideration to 0 < ηi < A0(α). (For larger initial displacements, the dynamics and the dependence of solutions on initial conditions are considerably more complicated and will be considered separately.) For the range of initial rectilinear displacements considered, we show that the MIC space is divided into three simply connected regions, in which (a) B = 0 and all initial energy is dissipated by the NES (Region I), (b) B > 0 and θ = 0 (Region IIA), and (c) B > 0 and θ = π (Region IIB).

For (α, σ, λ) = (0.1, 0.1, 0) and (1, 0.01, 0), the values of the coefficient ασ in the energy dissipation rate (Eq. (6)) are equal, but the coupling (represented by σ) and damping (represented by α) differ by factors of ten. For both combinations of α and σ, we illustrate the solutions in Regions I, IIA, and IIB by discussing specific trajectories, followed by a detailed delineation of the boundaries between regions. Dependence of the long-time solutions (characterized by B and θ) on ηi and θi within the three regions is discussed in Secs. 4 and 5.

3.1 (α, σ, λ) = (0.1, 0.1, 0)

3.1.1 Region I Trajectories (B = 0).

For (ηi, θi) = (0.03, 0.3π), Figs. 2(a)2(f) show η, /, θ, /, B, and D = σ[d2θ/2 sin θ + (/)2 cos θ] (an inertial coupling term). All initial energy is dissipated, and B decays to zero nearly monotonically with very small superimposed decaying oscillations. The asymptotic angular position is 0.1357π. To characterize the decay, we found the local maxima of the very small superimposed oscillations in B(τ). [The small time-step size (imposed by the small integration tolerances) compared with the characteristic oscillation time (2π) means that this can be done very accurately without interpolation.] We then used 500 Levenberg–Marquardt iterations to directly perform a nonlinear least-squares fit (i.e., without a logarithmic transformation of the B values) of 642 local maxima to the exponential model Bexp(τ) = g1 + g2eγτ over the interval 2000 ≤ τ ≤ 4000. This gave g1 = 1.484 × 10−5, g2 = 2.384 × 10−3, and γ = 8.816 × 10−4, with mean absolute values of the absolute and relative errors of 6.235 × 10−7 and 6.644 × 10−4, respectively. The decay rate is in excellent agreement with linear theory for decay of an infinitesimal disturbance to a zero-energy solution (see Sec. B available in the Supplemental Materials). We also fitted the local maxima, over the same range of τ, to a Landau model
BL(τ)=[ace2aτ+b]1/2
(15)
corresponding to one of the two nonzero solutions of dy/ = ayby3 and found a = −8.623 × 10−4, b = 4.078, and c = −153.6, with mean absolute values of the absolute and relative errors of 4.594 × 10−8 and 5.426 × 10−5, respectively. Since both fits have three parameters, it is clear that the Landau model is superior. The orientation θ and inertial coupling term D also undergo oscillations having the same dominant frequency, and decaying similarly in time. For the entire response, power spectra were calculated, and wavelet transformations were performed according to the procedure described in Sec. C available in the Supplemental Materials, showing (Figs. S1(a,b)) that the dominant frequencies for the rectilinear and rotational responses are 0.1622 and 0.1617, respectively, both very close to the natural frequency of the primary mass, namely, 1/(2π).
Fig. 2
Free response for (α, σ, λ) = (0.1, 0.1, 0) with (ηi, θi) = (0.03, 0.3π). Time series of (a) η, (b) dη/dτ, (c) θ/π, (d) dθ/dτ, (e) B, and (f) D.
Fig. 2
Free response for (α, σ, λ) = (0.1, 0.1, 0) with (ηi, θi) = (0.03, 0.3π). Time series of (a) η, (b) dη/dτ, (c) θ/π, (d) dθ/dτ, (e) B, and (f) D.
Close modal

For (ηi, θi) = (0.03, 0.7π), all initial energy is again dissipated, with the time series (Figs. S2(a–f), available in the Supplemental Materials) being similar to those for (ηi, θi) = (0.03, 0.3π) (Figs. 2(a)2(f)), except that the final orientation is θ = 0.8618π. For both MICs, θ is approximately 0.16π closer to the nearest integer multiple of π than is the initial orientation. Approximate symmetry of the behavior for these two cases is a consequence of the equal proximity of the values of θi to an integer multiple of π (leading to small values of the RHS of Eq. (2b)), as well as to the small value of ηi, for which linear behavior is a good approximation.

For (ηi, θi) = (0.06, 0.5π), Figs. 3(a)3(f) show that the response is similar to the cases discussed above, except that θ undergoes noticeably asymmetric oscillation about θ = 0.4866π, as opposed to the essentially one-sided approach of θ to its final value for the previous cases.

Fig. 3
Free response for (α, σ, λ) = (0.1, 0.1, 0) with (ηi, θi) = (0.06, 0.5π). Time series of (a) η, (b) dη/dτ, (c) θ/π, (d) dθ/dτ, (e) B, and (f) D.
Fig. 3
Free response for (α, σ, λ) = (0.1, 0.1, 0) with (ηi, θi) = (0.06, 0.5π). Time series of (a) η, (b) dη/dτ, (c) θ/π, (d) dθ/dτ, (e) B, and (f) D.
Close modal
3.1.2 Region IIA Trajectories (B > 0, θ = 0).

Figures 4(a)4(f) show time series for (ηi, θi) = (0.06, 0.4π). We see that B initially decreases rapidly, and then asymptotically approaches B = 0.03127 (Fig. 4(e)). The time series for θ undergoes oscillatory decay to zero (Fig. 4(c)). After an initial decrease from zero to −0.05397 (at τ = 1.42), and an increase to 0.05886 (at τ = 4.427), / undergoes oscillatory decay to zero (Fig. 4(d)). Wavelet transformations and power spectra (Figs. S3(a–f), available in the Supplemental Materials) again affirm that the dominant frequencies for the rectilinear and rotational responses are very close to 1/(2π).

Fig. 4
Free response for (α, σ, λ) = (0.1, 0.1, 0) with (ηi, θi) = (0.06, 0.4π). Time series of (a) η, (b) dη/dτ, (c) θ/π, (d) dθ/dτ, (e) B, and (f) D.
Fig. 4
Free response for (α, σ, λ) = (0.1, 0.1, 0) with (ηi, θi) = (0.06, 0.4π). Time series of (a) η, (b) dη/dτ, (c) θ/π, (d) dθ/dτ, (e) B, and (f) D.
Close modal
3.1.3 Region IIB Trajectories (B > 0, θ = π).

Figures 5(a)5(f) show the response for (ηi, θi) = (0.06, 0.8π), with long-time solution (B, θ) = (0.05509, π). The time series are qualitatively similar to those shown in Figs. 4(a)4(f), except that the approach to the semi-trivial solution is faster than for θi = 0.4π. This is because for θi = 0.4π, (a) the relative closeness of the initial angular displacement to 0.5π gives rise to considerably more “leverage” so that more initial energy is dissipated, and (b) θ remains closer to 0.5π, so that rotation persists longer than when the initial angular displacement is 0.8π.

Fig. 5
Free response for (α, σ, λ) = (0.1, 0.1, 0) with (ηi, θi) = (0.06, 0.8π). Time series of (a) η, (b) dη/dτ, (c) θ/π, (d) dθ/dτ, (e) B, and (f) D.
Fig. 5
Free response for (α, σ, λ) = (0.1, 0.1, 0) with (ηi, θi) = (0.06, 0.8π). Time series of (a) η, (b) dη/dτ, (c) θ/π, (d) dθ/dτ, (e) B, and (f) D.
Close modal
3.1.4 Boundaries Between Regions.

Here, we delineate the boundaries in the MIC space between the three regions discussed above. Figures 6(a) and 6(b) show B and θ for 0 < ηi ≤ 0.46 and 0 < θi < π, obtained from 14,268 trajectories (using nonuniform increments of ηi and θi). (Values of B and θ for many of these initial conditions are provided in Tables S1(a,b), available in the Supplemental Materials) Region I extends over the entire range of θi for ηi = 0, narrowing rapidly as ηi increases, and becoming exceedingly narrow for ηi greater than about 0.14. For ηi ≤ 0.10, the boundaries between Regions I and IIA and between Regions I and IIB (in both cases denoted by ηi,crit) rapidly approach and were determined as follows. For the boundary between Regions I and IIA, we rewrote the local approximation B = h(ηiηi,crit)1/2 as B2=h2ηih2ηi,crit, and then at each θi performed a least-squares fit (using the fitting parameters h2 and h2ηi,crit) to the computed values of B2 at the six or so smallest values of ηi for which B is between 10−3 and 10−2. The lower branch of the boundary between Regions I and IIB was determined by the same procedure. The two branches of the boundary separating Region I from IIB meet at a turning point near (ηi, θi) = (0.10, 0.5128π). For 0.10 < ηi ≤ 0.14, we computed the boundaries θi,crit,min (between Regions I and IIA) and θi,crit,max (between Regions I and IIB) by distinguishing trajectories for which B = 0 (Region I) from those for which either θ = 0 (Region IIA) or θ = π (Region IIB). For larger ηi, Δθ = θi,crit,maxθi,crit,min < 2 × 10−4π, with Δθ narrowing rapidly as ηi increases. In this part of Region I, which we refer to as the “depression,” the existence and location of the range for which B = 0 is inferred from the transition from θ = 0 to π as θi increases. As ηi increases, the dependence of B on θi shows a pronounced depression in Region I, which shifts from near θi = 0.5π for ηi < 0.10 to approximately 0.60π at ηi = 0.46, as indicated in Fig. 6(a). Behavior of the long-time solutions near this depression is discussed in more detail in Sec. 5.1.

Fig. 6
For (α, σ, λ) = (0.1, 0.1, 0), variation of (a) B∞ and (b) θ∞ with ηi and θi (Color version online.)
Fig. 6
For (α, σ, λ) = (0.1, 0.1, 0), variation of (a) B∞ and (b) θ∞ with ηi and θi (Color version online.)
Close modal

3.2 (α, σ, λ) = (1, 0.01, 0)

3.2.1 Trajectories in Region I, IIA, and IIB.

For (α, σ, λ) = (1, 0.01, 0), trajectories emanating from (ηi, θi) = (0.2, 0.5π) (in Region I), (0.2, 0.3π) (in Region IIA), and (0.2, 0.7π) (in Region IIB) are shown in Figs. S4(a–f), S5(a–f), and S6(a–f), respectively, available in the Supplemental Materials. By comparing Figs. S4, S5, and S6, available in the Supplemental Materials, to Figs. 3, 4, and 5, we note that the time series for (α, σ, λ) = (1, 0.01, 0) are qualitatively similar to those for (α, σ, λ) = (0.1, 0.1, 0) in each region. For these values of ηi and θi, the approximate symmetry discussed in Sec. 3.1 obtains. This is illustrated by time series for (ηi, θi) = (0.05, 0.2π) and (0.05, 0.8π) (shown in Figs. S7(a–f) and S8(a–f), respectively, available in the Supplemental Materials), in which the dynamics are approximately linear, owing to the small value of σ and the proximity of θi to an integer multiple of π.

3.2.2 Boundaries Between Regions.

For (α, σ, λ) = (1, 0.01, 0), in the MIC space covering 0 < ηi ≤ 1.18 and 0 < θi < π, we established the boundaries separating Region I from Regions IIA and IIB using the same procedure as for (α, σ, λ) = (0.1, 0.1, 0), except that the “cutoff” value of ηi is now 0.36, below which the fit to B2=h2ηih2ηi,crit was used. Figures 7(a) and 7(b) show the distributions of B and θ (interpolated based on trajectories computed using 4730 MICs), with the turning point of the boundary between Regions I and IIB being located near (ηi, θi) = (0.36, 0.5024π). Values of B and θ are shown in Tables S2(a) and S2(b), available in the Supplemental Materials, respectively, for all 4730 trajectories. For ηi > 0.36, the gap between the left and right branches is Δθ < 9 × 10−4π. Similar to results for (α, σ, λ) = (0.1, 0.1, 0), as ηi increases, the depression of Region I shifts from near θi = 0.5π to approximately 0.55π as ηi approaches 1.18. Behavior of the long-time solutions near this depression is discussed in more detail in Sec. 5.1.

Fig. 7
For (α, σ, λ) = (1, 0.01, 0), variation of (a) B∞ and (b) θ∞ with ηi and θi (Color version online.)
Fig. 7
For (α, σ, λ) = (1, 0.01, 0), variation of (a) B∞ and (b) θ∞ with ηi and θi (Color version online.)
Close modal

4 Complete Dissipation of Initial Energy Without Rectilinear Dissipation

For (α, σ, λ) = (0.1, 0.1, 0) and (1, 0.01, 0), the asymptotic approach of B(τ) to zero illustrated in Sec. 3 shows that the initial energy can be completely dissipated in Region I, a possibility not recognized in previous work [13]. This is of particular interest when the objective of design is to completely suppress vibration.

The dependence of B and θ on ηi for several values of θi is shown in Figs. 8(a)8(c) and Figs. S9(a–c), available in the Supplemental Materials, for (α, σ, λ) = (0.1, 0.1, 0) and (1, 0.01, 0), respectively. Results for both sets of dimensionless parameters are qualitatively similar, with increases in ηi leading to re-entry into Region I at θi = 0.52π and 0.55π for (α, σ, λ) = (0.1, 0.1, 0), and at θi = 0.5025π, 0.503π, and 0.504π for (α, σ, λ) = (1, 0.01, 0).

Fig. 8
Asymptotic states for (α, σ, λ) = (0.1, 0.1, 0) with 0 < ηi ≤ 0.46. Variation of B∞ with ηi for selected θi for (a) θi ≤ 0.5π and (b) θi > 0.5π. (c) Variation of θ∞ with ηi for selected θi. θi = 0.1π, ; 0.3π, ; 0.5π, ; 0.51π, ; 0.52π, ; 0.55π, ; 0.65π, ; 0.7π, ; 0.9π, . For θi = 0.52π, points near the depression are identified by solid dots. For θi = 0.55π, the depression is so narrow that no values of θ∞/π in the range 0 < θ∞/π < 1 were computed.
Fig. 8
Asymptotic states for (α, σ, λ) = (0.1, 0.1, 0) with 0 < ηi ≤ 0.46. Variation of B∞ with ηi for selected θi for (a) θi ≤ 0.5π and (b) θi > 0.5π. (c) Variation of θ∞ with ηi for selected θi. θi = 0.1π, ; 0.3π, ; 0.5π, ; 0.51π, ; 0.52π, ; 0.55π, ; 0.65π, ; 0.7π, ; 0.9π, . For θi = 0.52π, points near the depression are identified by solid dots. For θi = 0.55π, the depression is so narrow that no values of θ∞/π in the range 0 < θ∞/π < 1 were computed.
Close modal

For trajectories leading to B = 0, a key question is how long it takes to dissipate, say, 99% of the initial energy. In Region I, we define τ0.01 as the time at which E(τ0.01) = 0.01 E(0). Note that τ0.01 is finite in Region I and is not defined in most of Regions IIA and IIB, because less than 99% of the initial energy is dissipated in most of those regions. In each of those regions, there is, however, a very small zone adjacent to Region I, in which τ0.01 is defined and grows very rapidly as one moves away from the boundary with Region I, as shown in Fig. S10, available in the Supplemental Materials, for (α, σ, λ) = (0.1, 0.1, 0). On the scale of Figs. 9(a) and 9(b), that zone is too small to be apparent.

Fig. 9
Variation of τ0.01 in Region I with ηi and θi for (a) (α, σ, λ) = (0.1, 0.1, 0) and (b) (α, σ, λ) = (1, 0.01, 0) (Color version online.)
Fig. 9
Variation of τ0.01 in Region I with ηi and θi for (a) (α, σ, λ) = (0.1, 0.1, 0) and (b) (α, σ, λ) = (1, 0.01, 0) (Color version online.)
Close modal

Results in Figs. 9(a) and 9(b) for (α, σ, λ) = (0.1, 0.1, 0) and (1, 0.01, 0) show that values of τ0.01 for (α, σ, λ) = (1, 0.01, 0) are generally greater than those for (α, σ, λ) = (0.1, 0.1, 0). For these two sets of parameters with identical MICs [(ηi, θi) = (0.04, 0.4π)] lying within Region I, Figs. 10(a)10(c) show time series of dE/, D, and B. It is clear that coupling in the (α, σ, λ) = (0.1, 0.1, 0) case is stronger than for (α, σ, λ) = (1, 0.01, 0), giving rise to a greater initial dissipation rate, which ultimately leads to shorter τ0.01.

Fig. 10
For (ηi, θi) = (0.04, 0.4π), time series over 0 ≤ τ ≤ 400 of (a) dE/dτ, (b) D, and (c) B. (α, σ, λ) = (0.1, 0.1, 0), black; (α, σ, λ) = (1, 0.01, 0), red. (Color version online.)
Fig. 10
For (ηi, θi) = (0.04, 0.4π), time series over 0 ≤ τ ≤ 400 of (a) dE/dτ, (b) D, and (c) B. (α, σ, λ) = (0.1, 0.1, 0), black; (α, σ, λ) = (1, 0.01, 0), red. (Color version online.)
Close modal

Figures 9(a) and 9(b) show that for each ηi, the smallest value of τ0.01 occurs near the center of Region I (near θi = 0.5π for sufficiently small ηi), and that τ0.01 increases monotonically as θi departs from the center, consistent with the fact that Eq. (2b) shows that the strongest inertial coupling (and hence most efficient energy transfer) occurs when sin θ achieves its maximum value. For larger ηi (i.e., beyond values shown in Figs. 9(a) and 9(b)), the only values of θi in Region I for which all of the initial energy is dissipated lie in a very narrow range of θi (in the depression) greater than 0.5π. For these ηi, departure of the depression from θi = 0.5π occurs because initial rectilinear motion in the −η direction has the effect of immediately reducing θ, and it is only for θi in a narrow range greater than 0.5π that θ(τ) remains close enough to 0.5π for long enough that the initial energy can be completely dissipated. For (α, σ, λ) = (1, 0.01, 0) at θi = 0.503π, time series shown in Figs. 11(a)11(f), 12(a)12(f), and S11(a,b), available in the Supplemental Materials, illustrate this for ηi = 0.421 (in the depression) and 0.425 (above the depression, in Region IIA). For ηi = 0.421 (Figs. 11(a)11(f) and S11(a), available in the Supplemental Materials), θ remains near 0.5π much longer than for the ηi = 0.425 trajectory (Figs. 12(a)12(f) and S11(b), available in the Supplemental Materials). Similar behavior is observed for (α, σ, λ) = (0.1, 0.1, 0), where for ηi = 0.12, Figs. S12(a,b), available in the Supplemental Materials, show that the initial decay of the mean value of θ(τ) is considerably more rapid for θi = 0.511π (outside of the depression) than for θi = 0.512π (inside the depression). On the other hand, the “settling time,” expressed in terms of the time required for the energy to fall, say, 99% of the difference between its initial and final values, is considerably faster for θi = 0.512π. The dependence of τ0.01 on ηi and θi follows accordingly.

Fig. 11
Free response for (α, σ, λ) = (1, 0.01, 0) with (ηi, θi) = (0.421, 0.503π). Time series of (a) η, (b) dη/dτ, (c) θ/π, (d) dθ/dτ, (e) B, and (f) D.
Fig. 11
Free response for (α, σ, λ) = (1, 0.01, 0) with (ηi, θi) = (0.421, 0.503π). Time series of (a) η, (b) dη/dτ, (c) θ/π, (d) dθ/dτ, (e) B, and (f) D.
Close modal
Fig. 12
Free response for (α, σ, λ) = (1, 0.01, 0) with (ηi, θi) = (0.425, 0.503π). Time series of (a) η, (b) dη/dτ, (c) θ/π, (d) dθ/dτ, (e) B, and (f) D.
Fig. 12
Free response for (α, σ, λ) = (1, 0.01, 0) with (ηi, θi) = (0.425, 0.503π). Time series of (a) η, (b) dη/dτ, (c) θ/π, (d) dθ/dτ, (e) B, and (f) D.
Close modal

5 Asymptotically Rectilinear Harmonic Motion

For initial conditions outside of Region I, it is not possible to dissipate all initial energy, and the final state will be one for which B > 0, with θ an integer multiple of π. Here, for Regions IIA and IIB, we characterize the dependence of B, θ, and χ=1B2/ηi2 (the fraction of the initial energy ultimately dissipated) on ηi and θi.

5.1 Dependence of Asymptotic Rectilinear Amplitude on Initial Conditions.

For (α, σ, λ) = (1, 0.01, 0), Fig. 7(a) and Table S2(a), available in the Supplemental Materials, show that for ηi < A0(1), B increases monotonically with increasing ηi, except for values of θi for which there is more than one crossing from Region I into either Region IIA or IIB. Moreover, B generally varies unimodally with θi, having maximum values at θi = 0 and π, and falling to zero as Region I is approached from either side. The final angular position, θ, is zero everywhere in Region IIA and π everywhere in Region IIB, with a continuous variation through the intervening Region I. For (α, σ, λ) = (0.1, 0.1, 0), Fig. 6(a) and Table S1(a), available in the Supplemental Materials, show that the variation of B with ηi is monotonic for ηi less than approximately 0.4, again with the exception of those values of θi for which there is more than one crossing from Region I into either Region IIA or IIB. The more complicated dependence of B on ηi and θi just below ηi = A0(0.1) is discussed below.

For (α, σ, λ) = (0.1, 0.1, 0) and several θi, Figs. 8(a) and 8(b) show the local square-root dependence B = h(ηiηi,crit)1/2 of B on ηi used in Sec. 3 to determine the boundary of Region I. Figure 8(c) shows that there is a range of θi for which θ depends nonmonotonically on ηi. For each θi considered, the dependence of B on ηi is nearly linear over 0.3 ≤ ηi ≤ 0.4, beyond which the dependence is complicated near θi = 0.5π. There is a narrow range of θi, between the turning point (near θi = 0.5128π) and the value of θi (near 0.60π) in the depression for ηi = 0.46, for which there are two disjoint ranges of ηi wherein B = 0. We illustrate this for θi = 0.52π. Increasing ηi brings the initial state from Region I (with B = 0) into Region IIB (with B > 0) near ηi = 0.0848. Further increases in ηi result in B decreasing to zero as the initial condition again approaches Region I (near ηi = 0.1541). After passage through Region I (with B = 0), the initial condition emerges into Region IIA (near ηi = 0.1545), with further increases in ηi leading again to square-root dependence of B on ηi. At still higher ηi, B increases nearly linearly with ηi for ηi < 0.4. For still larger ηi, B increases monotonically but sublinearly as ηi increases. For (α, σ, λ) = (1, 0.01, 0), a similar dependence of B and θ on ηi and θi is shown in Figs. S9(a–c), available in the Supplemental Materials. Note that the nonmonotonic dependence of B on ηi and θi above ηi = 0.4 in the (α, σ, λ) = (0.1, 0.1, 0) case has no counterpart for (α, σ, λ) = (1, 0.01, 0).

For (α, σ, λ) = (0.1, 0.1, 0) and ηi = 0.1, Fig. 13 shows that B decreases monotonically with θi from B = ηi at θi = 0 to B = 0 at the boundary between Regions I and IIA and remains zero through the depression. Emerging from the depression, B grows rapidly as θi increases, ultimately reaching ηi at θi = π. Similar unimodal dependence of B on θi is observed for ηi = 0.2 and 0.3, with the width of the depression in Region I being less than Δθ = 10−5π, compared with 5 × 10−3π for ηi = 0.1. For ηi = 0.4, dependence of B on θi is more complex than the unimodal behavior found for ηi = 0.1, 0.2, and 0.3. For ηi = 0.4, B = 0 in only a narrow range of θi (0.584857π < θi < 0.584858π), where the lower and upper bounds are ones for which we compute θ = 0 and π, respectively. Thus, MICs leading to values of B smaller than the minima shown near θi = 0.58π (Fig. 13) are difficult to locate. The corresponding variation of θ with θi is shown in Figs. 14(a) and 14(b) for ηi = 0.1 and 0.4, respectively, and is used to identify the transition at larger ηi. For ηi = 0.1, the transition from θ = 0 to θ = π (across the width of Region I) occurs over a narrow range of θi, for which intermediate values of θ/π are shown at two values of θi. For ηi = 0.4, the range is exceedingly narrow, and no intermediate values of θ/π were computed.

Fig. 13
For (α, σ, λ) = (0.1, 0.1, 0), B∞ as a function of θi for ηi = 0.1 in red (labeled as A), ηi = 0.2 in green (labeled as B), ηi = 0.3 in blue (labeled as C), and ηi = 0.4 in black (labeled as D). (Color version online.)
Fig. 13
For (α, σ, λ) = (0.1, 0.1, 0), B∞ as a function of θi for ηi = 0.1 in red (labeled as A), ηi = 0.2 in green (labeled as B), ηi = 0.3 in blue (labeled as C), and ηi = 0.4 in black (labeled as D). (Color version online.)
Close modal
Fig. 14
For (α, σ, λ) = (0.1, 0.1, 0), variation of θ∞/π with θi for (a) ηi = 0.1 and (b) ηi = 0.4. Open circles (○) and squares (□) denote values of θi at which transitions occur from B∞ > 0 to B∞ = 0, and from B∞ = 0 to B∞ > 0, respectively. For ηi = 0.4, the depression is so narrow that no values of θ∞/π in the range 0 < θ∞/π < 1 were computed.
Fig. 14
For (α, σ, λ) = (0.1, 0.1, 0), variation of θ∞/π with θi for (a) ηi = 0.1 and (b) ηi = 0.4. Open circles (○) and squares (□) denote values of θi at which transitions occur from B∞ > 0 to B∞ = 0, and from B∞ = 0 to B∞ > 0, respectively. For ηi = 0.4, the depression is so narrow that no values of θ∞/π in the range 0 < θ∞/π < 1 were computed.
Close modal

To better understand the more complicated dependence of B on θi described above for ηi = 0.4 (see Fig. 13), we show time series of θ and / in Figs. 15(a)15(d) for θi = 0.563π and 0.5849π at ηi = 0.4. For θi = 0.563π (in Region IIA), Figs. 15(a) and 15(b) show that the orientation and angular velocity undergo three excursions, after which / settles into a multi-peak oscillation whose amplitude decays slowly to zero, while the angular displacement approaches θ = 0. On the other hand, for θi = 0.5849π (in Region IIB, very close to the boundary with Region I), Fig. 15(c) shows that θ undergoes seven oscillations about θi before (nonmonotonically) approaching π. Again following seven early oscillations, / also decays nonmonotonically to zero (Fig. 15(d)). That the amount of energy dissipated increases with the number of excursions explains why B is smaller for θi = 0.5849π than for θi = 0.563π. For (α, σ, λ) = (1, 0.01, 0), complicated dependence of B on θi near the depression of Region I also occurs (Fig. S13, available in the Supplemental Materials).

Fig. 15
Free response for (α, σ, λ) = (0.1, 0.1, 0). For (ηi, θi) = (0.4, 0.563π), time series of (a) θ/π and (b) dθ/dτ. For (ηi, θi) = (0.4, 0.5849π), time series of (c) θ/π and (d) dθ/dτ.
Fig. 15
Free response for (α, σ, λ) = (0.1, 0.1, 0). For (ηi, θi) = (0.4, 0.563π), time series of (a) θ/π and (b) dθ/dτ. For (ηi, θi) = (0.4, 0.5849π), time series of (c) θ/π and (d) dθ/dτ.
Close modal

5.2 Dependence of Energy Dissipation on Initial Conditions.

From Fig. 6(a), it is clear that for sufficiently large ηi, complete dissipation of initial energy is possible only in a very narrow range of θi. The ηi- and θi-dependence of the fraction of initial energy ultimately dissipated (χ=1B2/ηi2) is shown in Figs. 16(a) and 16(b) for (α, σ, λ) = (0.1, 0.1, 0) and (1, 0.01, 0), respectively. In each white area and in the associated depression extending upward from it (Region I in Figs. 8(a) and 8(b)), all initial energy is dissipated, and χ = 1. Over the entire range of θi in Regions IIA and IIB, χ generally decreases to a value less than 0.36 as ηi increases to 0.46. Complex dependence of χ on θi sets in near the depression (as reflected in the complex dependence of B on θi shown in Fig. 13 for ηi = 0.4). For (α, σ, λ) = (1, 0.01, 0), similar dependence of χ on MICs is shown in Fig. 16(b), with χ in Regions IIA and IIB being generally less than for (α, σ, λ) = (0.1, 0.1, 0).

Fig. 16
Variation of χ, the fraction of initial energy dissipated, with ηi and θi for (a) (α, σ, λ) = (0.1, 0.1, 0) and (b) (α, σ, λ) = (1, 0.01, 0). In Region I (white) all of the initial energy is dissipated (Color version online.)
Fig. 16
Variation of χ, the fraction of initial energy dissipated, with ηi and θi for (a) (α, σ, λ) = (0.1, 0.1, 0) and (b) (α, σ, λ) = (1, 0.01, 0). In Region I (white) all of the initial energy is dissipated (Color version online.)
Close modal

From the standpoint of energy dissipation, Figs. 16(a) and 16(b) show that the contours of the “worst” parts of the MIC space (i.e., with smallest χ) are very different for the two combinations of α and σ. For both cases, there is essentially no dissipation for initial conditions in significant portions of Regions IIA and IIB, near whose θi = 0 and θi = π boundaries there is very little inertial coupling between the translational and rotational motions, and hence little energy transfer from the former to the latter. The (α, σ, λ) = (0.1, 0.1, 0) and (1, 0.01, 0) cases differ markedly, however, in the shape and extent of the portions of Regions IIA and IIB in which there is essentially no dissipation, and indeed in the distribution of χ more generally. It is particularly noteworthy that for (α, σ, λ) = (0.1, 0.1, 0), the portion of Region IIB in which there is essentially no dissipation narrows as ηi increases beyond about 0.2, whereas that portion of Region IIA widens continuously as ηi increases.

6 Damping of the Rectilinear Motion of the Primary Mass

For λ > 0, it is clear that the long-time solution has zero energy, regardless of initial conditions. Thus, unlike the λ = 0 case discussed above, τ0.01 (the time required to dissipate 99% of the initial energy) is defined for all initial conditions. Here, we focus on how τ0.01 depends on ηi and θi for both combinations of α and σ.

Figures 17(a)17(d) show τ0.01 for (α, σ, λ) = (0.1, 0.1, λ) with λ = 10−4, 10−3, 3 × 10−3, and 10−2. At any combination of ηi and θi, τ0.01 decreases monotonically with increasing λ, as expected. A typical case is (ηi, θi) = (0.02, 0.3π), for which τ0.01 = 837.3, 817.8, 677.2, 499.7, and 269.3 for λ = 0, 10−4, 10−3, 3 × 10−3, and 10−2, respectively. We see that for combinations of ηi and θi lying within λ = 0's Region I, the values of τ0.01 for λ = 10−4 and 10−3 are quite similar to the λ = 0 case. The very large values of τ0.01 for λ = 0 in Region I near the boundaries with Regions IIA and IIB (red in Fig. 9(a)) have given way at λ = 10−4 to considerably smaller values of τ0.01 for the same values of ηi and θi, a clear influence of the rectilinear damping. (Note that the false-color scale is different for Fig. 9(a) than for Figs. 17(a)17(d). A more direct comparison of τ0.01 is provided in Fig. S14, available in the Supplemental Materials) Even for λ = 3 × 10−3 and 10−2, the largest values of τ0.01 continue to be centered about θi/π = 0.5, although at the same values of ηi and θi, τ0.01 is considerably smaller than it is for smaller λ. For each λ ≥ 0, the minimum value of τ0.01 occurs at θi/π = 0.5 in the limit as ηi → 0. (This is because θi/π = 0.5 provides the most “leverage,” and ηi → 0 is the limit in which temporal deviations from θi/π = 0.5 are smallest.)

Fig. 17
Variation of τ0.01 with ηi and θi for (a) (α, σ, λ) = (0.1, 0.1, 10−4), (b) (α, σ, λ) = (0.1, 0.1, 10−3), (c) (α, σ, λ) = (0.1, 0.1, 3 × 10−3), and (d) (α, σ, λ) = (0.1, 0.1, 10−2) (Color version online.)
Fig. 17
Variation of τ0.01 with ηi and θi for (a) (α, σ, λ) = (0.1, 0.1, 10−4), (b) (α, σ, λ) = (0.1, 0.1, 10−3), (c) (α, σ, λ) = (0.1, 0.1, 3 × 10−3), and (d) (α, σ, λ) = (0.1, 0.1, 10−2) (Color version online.)
Close modal
We can assess the relative importance of rectilinear and rotational dissipation from the ratio β = Grot/(Grect + Grot), where
Grect=λ0(dη/dτ)2dτandGrot=ασ0(dθ/dτ)2dτ
(16)
are the integrals of the instantaneous rates at which energy is dissipated by damping of the rectilinear and rotational motions, respectively, corresponding to the terms on the RHS of Eq. (6). The fact that dissipation by rotational damping (and the requisite transfer of energy from rectilinear motion to rotational motion) plays a dominant role for initial conditions lying in λ = 0's Region I for λ = 10−4 and 10−3 is made clear in Figs. 18(a) and 18(b). Even at larger values of λ, Figs. 18(c) and 18(d) show that the effect of rotational dissipation remains important for these initial conditions.
Fig. 18
Variation of β with ηi and θi for (a) (α, σ, λ) = (0.1, 0.1, 10−4), (b) (α, σ, λ) = (0.1, 0.1, 10−3), (c) (α, σ, λ) = (0.1, 0.1, 3 × 10−3), and (d) (α, σ, λ) = (0.1, 0.1, 10−2) (Color version online.)
Fig. 18
Variation of β with ηi and θi for (a) (α, σ, λ) = (0.1, 0.1, 10−4), (b) (α, σ, λ) = (0.1, 0.1, 10−3), (c) (α, σ, λ) = (0.1, 0.1, 3 × 10−3), and (d) (α, σ, λ) = (0.1, 0.1, 10−2) (Color version online.)
Close modal

For (α, σ, λ) = (1, 0.01, λ), Figs. 19(a)19(d) show τ0.01 with the same values of λ. The results, and their relationship to the λ = 0 case (Fig. 9(b)), are qualitatively quite similar to those for (α, σ, λ) = (0.1, 0.1, λ), with the main difference being that the values of τ0.01 are larger than for (α, σ, λ) = (0.1, 0.1, λ). This is a direct “carry-over” from the λ = 0 case and reflects the fact that the stronger inertial coupling (larger σ) plays a dominant role in transferring energy from the rectilinear mode to the rotational mode, where it is dissipated. Figures S15(a–d), available in the Supplemental Materials, show that, as for (α, σ, λ) = (1, 0.01, λ), rotational dissipation is dominant for small λ and remains important even at larger λ.

Fig. 19
Variation of τ0.01 with ηi and θi for (a) (α, σ, λ) = (1, 0.01, 10−4), (b) (α, σ, λ) = (1, 0.01, 10−3), (c) (α, σ, λ) = (1, 0.01, 3 × 10−3), and (d) (α, σ, λ) = (1, 0.01, 10−2) (Color version online.)
Fig. 19
Variation of τ0.01 with ηi and θi for (a) (α, σ, λ) = (1, 0.01, 10−4), (b) (α, σ, λ) = (1, 0.01, 10−3), (c) (α, σ, λ) = (1, 0.01, 3 × 10−3), and (d) (α, σ, λ) = (1, 0.01, 10−2) (Color version online.)
Close modal

7 Discussion

For the simple two-body point-mass system considered by Gendelman et al. [1] (i.e., with no direct damping of the primary mass), and for the equivalent distributed-mass and one-body systems governed by the same equations (see Ref. [4] and Sec. A available in the Supplemental Materials), we have shown that there is a part of the MIC space (Region I) in which every trajectory leads to a zero-energy final state. That result, not recognized in earlier work [13], along with the energy dissipation results in Secs. 4 and 5.2, are of particular interest from the standpoint of applications. Specifically, they show that advantageous selection of the dimensionless coupling parameter σ and dimensionless damping parameter α can significantly affect the portion of the initial condition space in which all initial energy is dissipated when there is no direct damping of the motion of the primary structure (Region I), and how fast that occurs. Such selection can also affect the fraction of the initial energy dissipated in other portions of the initial condition space (Regions IIA and IIB), as recognized by Saeed et al. [7] in a higher-dimensional structural application. Not surprisingly, the results depend significantly on the individual values of α and σ, and not simply on their product (see Eq. (6)). When direct damping of the motion of the primary structure is accounted for (Sec. 6), the λ = 0 results provide very good guidance as to how fast the zero-energy state is approached, especially when damping of the motion of the primary structure is relatively weak.

In some applications, the initial conditions are not motionless [7], and so it is important to know whether our analysis carries over to the more general case. While a complete exploration of the four-dimensional initial condition space for Eqs. (2a,b) is beyond the scope of this work, for (α, σ, λ) = (1, 0.01, 0) we have computed the range of initial rectilinear velocities vi for which initial conditions (ηi, vi, θi, Ωi) = (0.2, vi, 0.5π, 0) lead to trajectories with B = 0. We find that all initial potential and kinetic energy is dissipated for −0.0914 ≤ vi ≤ 0.0903. For (ηi, vi, θi, Ωi) = (0.2, 0.05, 0.5π, Ωi), similar computations over the range 0.06Ωi0.17 show that all of the initial potential and kinetic energy can be dissipated when all of the initial displacements and velocities are nonzero.

To put into context the initial dimensionless rectilinear displacements ηi considered here, we compare the ranges of initial energy corresponding to our initial rectilinear displacements of 0 < ηi ≤ 0.46 for (α, σ) = (0.1, 0.1), and 0 < ηi ≤ 1.18 for (α, σ) = (1, 0.01), to the ranges of initial energy considered by Saeed et al. [7]. In that computational study of damping by a rotational NES of impulsively driven structural vibration, the mass and stiffness matrices were taken from experiments with 51 kg two-story [11] and 10,000 kg nine-story [12] structures. In Sec. D available in the Supplemental Materials, we show that ηi = 0.46 corresponds to an initial energy nearly equal to the largest value considered by Saeed et al. for the nine-story structure, and to an energy approximately five times that at which their simulation predicted maximum dissipation (over a specified time interval) for that structure. For the two-story structure, maximum dissipation occurred at an initial energy only slightly greater than that corresponding to ηi = 0.46. On the other hand, ηi = 1.18 corresponds to an initial energy more than five times the largest considered by Saeed et al. for the nine-story structure, and slightly larger than the largest energy considered for the two-story structure. This shows that the initial energies considered here are comparable with, and in some cases considerably larger than, those relevant to structural vibration mitigation.

When design criteria allow for rotation of the primary structure about an axis not containing its center of mass, then as discussed in Sec. 1 and in Sec. A available in the Supplemental Materials, the dimensionless damping coefficient and the distance of the center of mass of the primary structure from the axis of rotation can be selected to achieve the same dynamical response that could have been effected by a second rotating mass, thus eliminating the second mass and its connection to the shaft about which it would rotate.

Finally, in addition to structural applications, we note that a rotational NES can both suppress and excite FIV in flow past a cylinder [46]. This suggests potential applications to enhancement of FIV of submerged cylinders in energy harvesting by vibrating cylinders submerged in riverine, estuarine, or marine currents [13], as well as in low Reynolds number mixing applications [14,15].

8 Conclusions

For a rotational nonlinear energy sink with no direct damping of the motion of the primary mass, we have examined part of the motionless projection of the initial condition space (i.e., in which the initial rectilinear and angular velocities are zero) for two combinations of the dimensionless coupling parameter and dimensionless rotational damping coefficient. We have identified a range of motionless initial conditions for which every trajectory leads to previously unrecognized zero-energy final states and characterized the time required for the initial energy to dissipate. We have also shown that the previously identified solutions, with harmonic rectilinear motion of the primary mass, and no motion of the rotatable mass, are (orbitally) stable only for limited ranges of the amplitude, and unstable outside of those ranges. Finally, when direct damping of the rectilinear motion of the primary mass is weak, we show that the characteristic time to dissipate the initial energy correlates well with predictions for the undamped case.

Funding Data

  • The authors gratefully acknowledge support from National Science Foundation (NSF) Grant CMMI-1363231.

Conflict of Interest

There are no conflicts of interest.

Data Availability Statement

The datasets generated and supporting the findings of this article are obtainable from the corresponding author upon reasonable request. The authors attest that all data for this study are included in the paper.

References

1.
Gendelman
,
O. V.
,
Sigalov
,
G.
,
Manevitch
,
L. I.
,
Mane
,
M.
,
Vakakis
,
A. F.
, and
Bergman
,
L. A.
,
2012
, “
Dynamics of an Eccentric Rotational Nonlinear Energy Sink
,”
ASME J. Appl. Mech.
,
79
(
1
), p.
011012
. 10.1115/1.4005402
2.
Sigalov
,
G.
,
Gendelman
,
O. V.
,
AL-Shudeifat
,
M. A.
,
Manevitch
,
L. I.
,
Vakakis
,
A. F.
, and
Bergman
,
L. A.
,
2012
, “
Alternation of Regular and Chaotic Dynamics in a Simple Two-Degree-of-Freedom System With Nonlinear Inertial Coupling
,”
Chaos
,
22
(
1
), p.
013118
. 10.1063/1.3683480
3.
Sigalov
,
G.
,
Gendelman
,
O. V.
,
AL-Shudeifat
,
M. A.
,
Manevitch
,
L. I.
,
Vakakis
,
A. F.
, and
Bergman
,
L. A.
,
2012
, “
Resonance Captures and Targeted Energy Transfers in an Inertially Coupled Rotational Nonlinear Energy Sink
,”
Nonlin. Dyn.
,
69
(
4
), pp.
1693
1704
. 10.1007/s11071-012-0379-1
4.
Tumkur
,
R. K. R.
,
Pearlstein
,
A. J.
,
Masud
,
A.
,
Gendelman
,
O. V.
,
Blanchard
,
A.
,
Bergman
,
L. A.
, and
Vakakis
,
A. F.
,
2017
, “
Intermediate Reynolds Number Flow Past a Sprung Circular Cylinder With an Internal Nonlinear Rotational Dissipative Element
,”
J. Fluid Mech.
,
828
, pp.
196
235
. 10.1017/jfm.2017.504
5.
Blanchard
,
A. B.
,
Bergman
,
L. A.
,
Vakakis
,
A. F.
, and
Pearlstein
,
A. J.
,
2019
, “
Coexistence of Multiple Long-Time Solutions for Two-Dimensional Laminar Flow Past a Linearly Sprung Circular Cylinder With a Rotational Nonlinear Energy Sink
,”
Phys. Rev. Fluids
,
4
(
5
), p.
054401
. 10.1103/PhysRevFluids.4.054401
6.
Blanchard
,
A. B.
, and
Pearlstein
,
A. J.
,
2020
, “
On-Off Switching of Vortex Shedding and Vortex-Induced Vibration in Cross-Flow Past a Circular Cylinder by Locking or Releasing a Rotational Nonlinear Energy Sink
,”
Phys. Rev. Fluids
,
5
(
2
), p.
023902
. 10.1103/PhysRevFluids.5.023902
7.
Saeed
,
A. S.
,
AL-Shudeifat
,
M. A.
,
Vakakis
,
A. F.
, and
Cantwell
,
W. J.
,
2019
, “
Rotary-Impact Nonlinear Energy Sink for Shock Mitigation: Analytical and Numerical Investigations
,”
Arch. Appl. Mech.
,
90
(
3
), pp.
495
521
. 10.1007/s00419-019-01622-0
8.
Tondl
,
A.
,
Ruijgrok
,
T.
,
Verhulst
,
F.
, and
Nabergoj
,
R.
,
2000
,
Autoparametric Resonance in Mechanical Systems
,
Cambridge University Press
,
New York
.
9.
Marquardt
,
D.
,
1963
, “
An Algorithm for Least-Squares Estimation of Nonlinear Parameters
,”
SIAM J. Appl. Math.
,
11
(
2
), pp.
431
441
. 10.1137/0111030
10.
Kovacic
,
I.
,
Rand
,
D. H.
, and
Sah
,
S. M.
,
2018
, “
Mathieu’s Equation and Its Generalizations: Overview of Stability Charts and Their Features
,”
ASME Appl. Mech. Rev.
,
70
(
2
), p.
020802
. 10.1115/1.4039144
11.
Wierschem
,
N. E.
,
Quinn
,
D. D.
,
Hubbard
,
S. A.
,
AL-Shudeifat
,
M. A.
,
McFarland
,
D. M.
,
Luo
,
J.
,
Fahnestock
,
L. A.
,
Spencer
,
B. F.
,
Vakakis
,
A. F.
, and
Bergman
,
L. A.
,
2012
, “
Passive Damping Enhancement of a Two-Degree-of-Freedom System Through a Strongly Nonlinear Two-Degree-of-Freedom Attachment
,”
J. Sound Vib.
,
331
(
25
), pp.
5393
5407
. 10.1016/j.jsv.2012.06.023
12.
Luo
,
J.
,
Wierschem
,
N. E.
,
Fahnestock
,
L. A.
,
Spencer
,
B. F.
,
Quinn
,
D. D.
,
McFarland
,
D. M.
,
Vakakis
,
A. F.
, and
Bergman
,
L. A.
,
2014
, “
Design, Simulation, and Large-Scale Testing of an Innovative Vibration Mitigation Device Employing Essentially Nonlinear Elastomeric Springs
,”
Earthquake Eng. Struct. Dyn.
,
43
(
12
), pp.
1829
1851
. 10.1002/eqe.2424
13.
Lee
,
J. H.
, and
Bernitsas
,
M. M.
,
2011
, “
High-Damping, High-Re VIV Tests for Energy Harvesting Using the VIVACE Converter
,”
Ocean Eng.
,
38
(
16
), pp.
1697
1712
. 10.1016/j.oceaneng.2011.06.007
14.
Deshmukh
,
S. R.
, and
Vlachos
,
D. G.
,
2005
, “
Novel Micromixers Driven by Flow Instabilities: Application to Post-Reactors
,”
AIChE J.
,
51
(
12
), pp.
3193
3204
. 10.1002/aic.10591
15.
Ortega-Casanova
,
J.
,
2017
, “
On the Onset of Vortex Shedding From 2D Confined Rectangular Cylinders Having Different Aspect Ratios: Applications to Promote Mixing Fluids
,”
Chem. Eng. Proc.: Proc. Intensif.
,
120
, pp.
81
92
. 10.1016/j.cep.2017.06.014