This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.
Paper The following article is Open access

Coherent dynamics in frustrated coupled parametric oscillators

, , , , and

Published 19 August 2020 © 2020 The Author(s). Published by IOP Publishing Ltd on behalf of the Institute of Physics and Deutsche Physikalische Gesellschaft
, , Focus on Time Crystals Citation Marcello Calvanese Strinati et al 2020 New J. Phys. 22 085005 DOI 10.1088/1367-2630/aba573

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

1367-2630/22/8/085005

Abstract

We explore the coherent dynamics in a small network of three coupled parametric oscillators and demonstrate the effect of frustration on the persistent beating between them. Since a single-mode parametric oscillator represents an analogue of a classical Ising spin, networks of coupled parametric oscillators are considered as simulators of Ising spin models, aiming to efficiently calculate the ground state of an Ising network—a computationally hard problem. However, the coherent dynamics of coupled parametric oscillators can be considerably richer than that of Ising spins, depending on the nature of the coupling between them (energy preserving or dissipative), as was recently shown for two coupled parametric oscillators. In particular, when the energy-preserving coupling is dominant, the system displays everlasting coherent beats, transcending the Ising description. Here, we extend these findings to three coupled parametric oscillators, focussing in particular on the effect of frustration of the dissipative coupling. We theoretically analyse the dynamics using coupled nonlinear Mathieu's equations, and corroborate our theoretical findings by a numerical simulation that closely mimics the dynamics of the system in an actual experiment. Our main finding is that frustration drastically modifies the dynamics. While in the absence of frustration the system is analogous to the two-oscillator case, frustration reverses the role of the coupling completely, and beats are found for small energy-preserving couplings.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Parametric oscillators are a viable experimental platform to study the physics of time crystals, i.e., systems that can spontaneously break time translational symmetry [1, 2]. The possibility of the existence of such a phase of matter at equilibrium was first proposed in 2012 by Frank Wilczek and collaborators [3, 4], both for quantum and classical systems. The original proposal evokes the possibility for a system to break continuous time translational symmetry, in analogy with the formation of space crystals in condensed matter where space translational symmetry is broken. Shortly after its proposal, it became clear that this kind of time-crystalline phase cannot exist at equilibrium [57]. However, following Wilczek's original idea, it was understood that time crystals can be realized out of equilibrium, in periodically-driven system, also referred to as Floquet systems. This new type of time crystals, dubbed Floquet time crystals, accounts for the fact that, under certain conditions, a periodically-driven system can break the discrete time translational symmetry enforced by the external drive [818]: instead of merely following the external drive, the system undergoes a periodic motion at a frequency that is different from that of the drive (see reference [1] for a review).

The periodically driven single-mode classical parametric oscillator is the canonical example of period-doubling instability (see reference [19, 20] for an introduction), and represents the simplest case of a classical Floquet time crystal. Indeed, when excited above the amplification threshold, the parametric oscillator oscillates at half the frequency of the drive and admits only two distinct phase solutions, dubbed '0' and 'π', with a relative shift in time by one period of the drive. One of the two solutions is chosen by the system depending on the initial conditions, a phenomenology analogous to a spontaneous breaking of a ${\mathbb{Z}}_{2}$ (Ising) symmetry. Because of this, a single degenerate parametric oscillator may be regarded as a classical bit, or an Ising spin, where the two states 'up' or 'down' of the spin are given by the two distinct '0' and 'π' solutions. Exploiting this property, networks of many coupled parametric oscillators have been proposed as a platform, called coherent Ising machine (CIM) [21], to simulate the behaviour of a network of many coupled Ising spins. Such a machine, whose experimental realization has been reported in reference [2224], is envisioned to solve the NP-hard problem of finding the ground state of the classical Ising model [25]. In the last years, the analysis of various issues related to the computational performance of CIMs has been the focus of a remarkable amount of work, not only implementing CIMs using parametric-oscillator networks [2630], which are the focus of this paper, but also using digital computers [31, 32], polariton networks [33], electrical oscillators [34], optoelectronical oscillators [35], and laser networks [36].

While the underlying assumption in parametric-oscillator-based CIMs (henceforth PO-CIMs) is that a system of coupled parametric oscillators behaves as a set of coupled Ising spins, we pointed out recently that already a pair of coupled parametric oscillators may display a much richer dynamics, beyond the Ising description, depending on the nature of the coupling (energy-preserving or dissipative) between the oscillators [37, 38]. Specifically, we studied in detail both theoretically and experimentally in a radio-frequency experiment a pair of coupled parametric oscillators, which is the minimal system to explore nontrivial coupling effects. Our main finding was that, when driven above the amplification threshold at the parametric resonance condition, the two oscillators can either display persistent coherent beats when the coupling is mostly energy preserving, or behave as a PO-CIM [21] when the coupling is mostly dissipative.

The existence of such a nontrivial dynamics in just a pair of coupled parametric oscillators opens the question on how the nature of the coupling affects the dynamics of a larger network composed by more than two parametric oscillators, specifically with potential implications in the context of PO-CIMs and in the view of exploiting large-scale networks of coupled parametric oscillators to realize classical many-body time crystals [39]. Motivated by these perspectives, we present in this paper a detailed theoretical and numerical analysis of three coupled parametric oscillators, which is the minimal system where nontrivial connectivity effects can be studied. The coupling between any two oscillators is parametrized by two coupling components—energy-preserving and dissipative. The main focus of this paper is to analyse for specific choices of the coupling matrix the effect of frustration (defined as the situation where the dissipative couplings prevent the oscillators from adjusting their phases to energetically minimize every link [40]), which turns out to be dramatic.

To reach this goal, we model each parametric oscillator as a classical variable and describe the system by three coupled nonlinear Mathieu's equations [38], in the presence of an external pump, intrinsic dissipation, and pump depletion nonlinearity, to analyse the phase diagram of the system for different values of the system parameters, as detailed hereon. Our theoretical predictions are confirmed by a low-level numerical simulation of the field propagation within the parametric oscillators both in time and space, as close as possible to an actual experimental setup. Our numerical scheme emulates directly the dynamics of the field inside a cavity with parametric gain, with no explicit mention of the equations of motion studied in our analytical model.

Our main finding is that frustration totally inverts the dynamical picture of the coupled system. While in the absence of frustration the system behaves similar to the two-oscillator case, where beats are observed only when the energy-preserving coupling is larger than the dissipative one, in the presence of frustration we find two main differences: first, the system shows coherent everlasting beats for small energy-preserving couplings. This finding can be reasoned by the fact that a frustrated system cannot distinguish between two (or more) degenerate Ising states that are found when the coupling is purely dissipative. Thus, any non-vanishing value of the energy-preserving coupling induces beating between those degenerate states. Second, for large energy-preserving couplings and large frustration, the network converges to a phase-locked oscillation, which however is not the Ising ground state.

This paper is organized as follows. In section 2, we briefly review our previous results of reference [37, 38] for the simpler case of two coupled parametric oscillators, introducing our model and notations. We then present our theoretical analysis for the case of three coupled parametric oscillators in section 3. We discuss in section 4 a possible experimental implementation of our system, and present the results of the low-level numerical simulation of such an experiment. We then draw our conclusions in section 5, and report some relevant details on the calculations in the appendixes.

2. Two parametric oscillators

Before moving to the analysis of three-coupled oscillators, let us shortly review the relevant notation and analytical tools of our previous work in reference [37, 38]. The familiar reader can directly move on to section 3 for the discussion on three-coupled parametric oscillators.

2.1. Model and notation

We consider a system of two identical single-mode parametric oscillators, with equal proper frequency ω0, driven by an external pump field at frequency 2ω0 and with amplitude h, injected into a parametric amplifier (PA) [41] as depicted in figure 1. The field inside each oscillator 1 (or 2) is identified by a classical variable x1 (x2). The two oscillators are coupled by a power-splitter coupling [37], which accounts for: (i) transmission coefficients c11 and c22 for oscillator 1 and 2, respectively, which renormalize the intrinsic loss of each oscillator, providing an overall loss rate that we denote by g, and (ii) coupling coefficients c12 and c21, which give the rate of energy exchange between the two oscillators. In this framework, the fields x1 and x2 are coupled according to the equation (see appendix A)

Equation (1)

where the dot denotes the time derivative. The dynamics of the two-oscillator system is described by a pair of coupled Mathieu's equations [38]

Equation (2)
Figure 1.

Figure 1. Schematic representation of the two-parametric-oscillator system. Each parametric oscillator is described by a classical field xk, with k = 1, 2, driven by an external pump h(t) = h sin(2ω0t) injected into a PA. The coupling between the oscillators is described in the general case by a coupling matrix that accounts for (i) transmittance coefficients c11 and c22, which renormalize the intrinsic losses of the oscillators, and (ii) coupling coefficients c12 and c21, which determine the rate of energy flow from oscillator 1 to 2, and vice versa.

Standard image High-resolution image

Equation (2) also includes a second-order nonlinearity in the amplitude of the pump field (hereafter referred to as 'pump-depletion nonlinearity'), whose strength is quantified by β. Such a nonlinearity describes the fact that the intensity of the pump field inside each oscillator is depleted by x1 and x2, and in many experimental contexts captures the most relevant nonlinear process [38]. In general, the rate of energy flow between the two oscillators can be unbalanced, i.e., c12c21, indicating dissipation in the coupling itself. Without loss of generality, one can parametrize the coupling coefficients in terms of an antisymmetric and symmetric part with respect to the exchange x1x2 in equation (2): c12 = rα and c21 = r + α, where the antisymmetric part r ⩾ 0 represents the energy-preserving component of the coupling, whereas the symmetric part α ⩾ 0 is the dissipative one.

The energy-preserving coupling r induces a coherent exchange of energy between the two oscillators. Its energy-preserving nature follows from the fact that the equations of motion (2), with β = 0, g = 0, and c12 = c21 = r, can be derived from the Hamilton's equations [19] starting from the Hamiltonian

Equation (3)

where p1 and p2 are the canonical momentum variables for x1 and x2, respectively. Such an Hamiltonian is analogous to that of a charged particle (charge q = 0r/2) moving on a two-dimensional plane identified by the spatial coordinates (x1, x2, z = 0), subject to a vector potential $\mathbf{A}={\left(-{x}_{2},{x}_{1},0\right)}^{\text{T}}$, where T denotes the transposition (for the details of the derivation, see appendix B).

Figure 2.

Figure 2. Dynamical phase diagram of two coupled parametric oscillators, described by equation (4). (Left) phase diagram in the h/(2g) vs. r/g plane, for $\tilde {\alpha }=0.15$, and (right) configuration of the fixed points in the BR vs AR plane (while AI = BI = 0), where black and green dots represent unstable and stable points, respectively. The flow of the slow varying amplitudes from the solution of equation (4) is marked by red curves. The phase diagram consists of four main phases, characterized by different configurations of the fixed points, as shown in right panels: (a) below the oscillation threshold, where only the origin A = B = 0 is a stable attractor; (b) the Ising or PO-CIM region, where the system has two stable fixed points, corresponding to the two ground-state Ising solutions (00) and (ππ) (for 'ferromagnetic' $\tilde {\alpha }{ >}0$); (c) a phase in which a stable limit cycle stabilizes the long-time dynamics, and the system displays coherent beats; (d) a phase with four stable fixed points, corresponding to both ground-state (00) and (ππ), and excited-state (0π) and (π0) configurations. The red dashed line in the phase diagram is the boundary for the oscillation threshold ${\tilde {h}}_{\text{th}}$. Other phases with more than four fixed points [38] are not labelled and not relevant for the present discussion, and additional unstable fixed points different from the origin are not shown for the sake of clarity.

Standard image High-resolution image

The dissipative coupling α, in contrast, introduces additional loss or gain terms [38], which give rise to the Ising-type coupling between the oscillators that is usually considered in the standard analysis of PO-CIMs [2124, 2630]. This coupling guides the convergence of the two-oscillator system to the desired Ising ground state [38]. In the long-time limit, the two oscillators will prefer to lock according to the sign of α: in-phase for 'ferromagnetic' coupling (α > 0), yielding the two 'ferromagnetic' configurations (00) or (ππ), or in anti-phase for 'anti-ferromagnetic' coupling (α < 0), yielding the two 'anti-ferromagnetic' configurations (0π) or (π0), where the first (second) label denotes the corresponding phase solution the first (second) oscillator.

2.2. Phase-locking and beats

We now review the effect of the interplay between r and α on the long-time dynamics of the system. We focus on the dynamics of the slow-varying amplitudes that modulate the fast-varying oscillations at half the pump frequency, which are instead integrated out, by employing the multiple-scale perturbative expansion in [38]. We take the intra-cavity loss g as a small expansion parameter, and identify the fast-varying and slow-varying time scales as t = 2π/ω0 and τ = gt, respectively. By writing ${x}_{1}\left(t,\tau \right)=A\left(\tau \right)\enspace {\mathrm{e}}^{i{\omega }_{0}t}+{A}^{{\ast}}\left(\tau \right)\enspace {\mathrm{e}}^{-i{\omega }_{0}t}$ and ${x}_{2}\left(t,\tau \right)=B\left(\tau \right)\enspace {\mathrm{e}}^{i{\omega }_{0}t}+{B}^{{\ast}}\left(\tau \right)\enspace {\mathrm{e}}^{-i{\omega }_{0}t}$, where A(τ) and B(τ) are the complex amplitudes that encode the slow-varying dynamics, and by rescaling $\tilde {h}=h/g$, $\tilde {r}=r/g$, $\tilde {\alpha }=\alpha /g$, and $\tilde {\tau }={\omega }_{0}\tau $, one finds that A and B obey the following set of coupled first-order differential equations [38]:

Equation (4)

Equation (4) can be further recast in terms of the real and imaginary parts of the complex amplitudes, A = AR + iAI and B = BR + iBI, where AR (BR) and AI (BI) are, respectively, the real and imaginary parts of A (B). The long-time dynamics is determined by the configuration and stability of the fixed points $\left({\overline{A}}_{\text{R}},{\overline{A}}_{\text{I}},{\overline{B}}_{\text{R}},{\overline{B}}_{\text{I}}\right)$ of equation (4), where the overline is used to denote the steady-state values of the quadratures. Note however that, sufficiently close to the oscillation threshold, AI and BI decay very quickly $\left({\overline{A}}_{\text{I}}={\overline{B}}_{\text{I}}=0\right)$ [38] due to the phase dependent amplification and squeezing in parametric oscillators, allowing to focus the discussion only on the dynamics of the real parts AR and BR.

While in general the configuration of the fixed points depends on the form of the nonlinearity, especially far from the amplification threshold, most of the interesting physics throughout this paper occurs close to the threshold, where nonlinear effects are negligible and the system is almost linear. The properties of the system at threshold can be therefore found by focussing on the spectrum of the Jacobian matrix around the origin A = B = 0, analysing specifically the eigenvalue with largest real part (λmax, which we dub 'most efficient eigenvalue' from now on). Importantly, when the system exceeds the amplification threshold ${\tilde {h}}_{\text{th}}$, defined by the condition Re[λmax] = 0+, the imaginary part of λmax determines the frequency of the beats at threshold ωB = ω0g|Im[λmax]| between the two oscillators. The beat frequency ωB is the key observable to describe the behaviour of the system as the oscillators are driven above the oscillation threshold. Specifically, when ωB = 0, A and B reach eventually constant values ${\overline{A}}_{\text{R}}$ and ${\overline{B}}_{\text{R}}$, and the system behaves as a time crystal, and can simulate Ising spins. Indeed, for positive (negative) ${\overline{A}}_{\text{R}}$, x1 converges to the '0' ('π') solution, and analogously for x2. Instead, for ωB > 0, A and B display persistent coherent beats. The presence of the beats implies that each oscillator x1,2 periodically flips between the '0' and 'π' solutions, and therefore the system neither obeys the Ising description, nor it behaves as a time crystal.

A concrete calculation of the phase diagram of the system in equation (4) is shown in figure 2, where we identify the following main phases:

  • (a)  
    the sub-threshold phase, for a pump amplitude $\tilde {h}{< }{\tilde {h}}_{\text{th}}$, where the origin A = B = 0 is the only stable fixed point of equation (4);
  • (b)  
    the Ising or PO-CIM phase, for $\tilde {h}{ >}{\tilde {h}}_{\text{th}}$, where two stable fixed points are found, the origin being unstable. Phase locking occurs at (00) or (ππ), for 'ferromagnetic' $\tilde {\alpha }{ >}0$, or at (0π) or (π0) for 'antiferromagnetic' $\tilde {\alpha }{< }0$ (not shown);
  • (c)  
    the beating phase, for $\tilde {h}{ >}{\tilde {h}}_{\text{th}}$, where the long-time dynamics of the oscillators' amplitudes is attracted into a stable limit cycle;
  • (d)  
    a phase with four stable fixed points, corresponding to all the four Ising configurations (00), (ππ), (0π), and (π0), in which the system behaves as two decoupled spins. This behaviour matches the one previously discussed in reference [21].

Slightly above the amplification threshold ${\tilde {h}}_{\text{th}}$, the PO-CIM phase is found for $\tilde {r}{< }\tilde {\alpha }$ (ωB = 0), whereas the beating phase is found for $\tilde {r}{ >}\tilde {\alpha }$ (ωB > 0). At threshold, the system therefore undergoes a transition between the PO-CIM and the beating behaviour when $\tilde {r}=\tilde {\alpha }$.

3. Three coupled parametric oscillators

The results of reference [37, 38], reviewed in the previous section, pointed out the existence of a persistent coherent beating dynamics in coupled parametric oscillators, not considered in the standard analysis of PO-CIMs. A natural question that arises is how the presence of such a dynamics affects a more structured network with more than two coupled parametric oscillators. Here, we begin to address this question by extending the previous discussion to the case of three degenerate coupled parametric oscillators, which is the simplest configuration where one can systematically study the role of connectivity. Specifically, our main focus is to study how frustration of the dissipative coupling (which reflects the underlying Ising model) affects the coherent dynamics of the system.

3.1. Model

A schematic representation of a three-oscillator system is shown in figure 3. Now, the coupling matrix c between the oscillators, according to equation (1), is

Equation (5)

The system is now described by a set of three coupled nonlinear Mathieu's equations, which we write in a compact form for the sake of simplicity (j, k = 1, 2, 3):

Equation (6)

where sgn(⋅) denotes the sign function. From equation (6), one obtains the corresponding multiple-scale equations for the slow-varying amplitudes of the fields {xj} as in equation (4). Here, we renormalize each element of the coupling matrix as ${\tilde {c}}_{jk}={c}_{jk}/g$, and to ease the notation, we use the symbols {Xj(τ)} of the amplitudes, such that ${x}_{j}\left(t,\tau \right)={X}_{j}\left(\tau \right){\mathrm{e}}^{i{\omega }_{0}t}+{X}_{j}^{{\ast}}\left(\tau \right){\mathrm{e}}^{-i{\omega }_{0}t}$. The equations for the complex amplitudes are then determined (j, k = 1, 2, 3):

Equation (7)

As in section 2, we decompose the coupling cjk in equations (5)–(7) between any two oscillators in terms of an energy-preserving (antisymmetric) and dissipative (symmetric) part, respectively rjk and αjk. Due to this increase of parameter space with respect to the case in section 2 (the coupling matrix now has in general six independent components), we focus on a specific choice of the coupling matrix, with the ambition to highlight the role of frustration in the dissipative components of the coupling matrix. We choose rjk = r for all j and k, and introduce two different dissipative couplings: α12 = η and α13 = α23 = α, so that the coupling matrix in equation (5) reads

Equation (8)

In the rest of the paper, our goal is to study the physics of the system near threshold as a function of the coupling parameters $\tilde {r}$, $\tilde {\alpha }$, and $\tilde {\eta }$, where as before all quadratures are real $\left({\overline{X}}_{1,\text{I}}={\overline{X}}_{2,\text{I}}={\overline{X}}_{3,\text{I}}=0\right)$. Before discussing this general case, we first focus on two main configurations of interest for the coupling matrix in equation (8). Namely, assuming $\tilde {r},\tilde {\alpha }{\geqslant}0$: (i) the non-frustrated case, for $\tilde {\eta }=\tilde {\alpha }$, and (ii) the fully-frustrated case, for $\tilde {\eta }=-\tilde {\alpha }$. The reason why we focus on these two fine-tuned cases is because they are, on one hand, easily analytically tractable, and on the other hand, they capture the dramatic effect of frustration in the dissipative coupling. We discuss the general case later in section 3.4.

Figure 3.

Figure 3. Schematic representation of the three-parametric-oscillator system. Similar to figure 1, the parametric oscillators are described by a classical field xk (k = 1, 2, 3), driven by the same external pump h(t) = h sin(2ω0t). The coupling between the oscillators, describing the connectivity of the system, includes all the mutual couplings cjk for j, k = 1, 2, 3. The transmittance coefficients cjj for j = 1, 2, 3 have been omitted from the figure for the sake of simplicity.

Standard image High-resolution image

3.2. Non-frustrated network

First, we analytically discuss the threshold properties of the non-frustrated network, for $\tilde {\eta }=\tilde {\alpha }$ in equation (8). We analyse separately the cases of $\tilde {r}{ >}\tilde {\alpha }$ and $\tilde {r}{< }\tilde {\alpha }$.

For $\tilde {r}{ >}\tilde {\alpha }$, we have ${\lambda }_{\mathrm{max}}=-1/2+\tilde {h}/4-{\mathrm{e}}^{i\pi /3}\enspace F\left(\tilde {r},\tilde {\alpha }\right)+{\mathrm{e}}^{-i\enspace \pi /3}\enspace G\left(\tilde {r},\tilde {\alpha }\right)$, where

Equation (9)

Since λmax is complex, beats are found. From the imaginary parts of λmax, one can find the expression of the beat frequency:

Equation (10)

In particular, for $\tilde {r}$ approaching $\tilde {\alpha }$, the frequency of the beats reduces towards zero with the critical behaviour of ${\omega }_{\text{B}}\sim {\left(\tilde {r}-\tilde {\alpha }\right)}^{1/3}$, differently from the critical exponent of 1/2 in the two-oscillator case [38].

When $\tilde {r}{< }\tilde {\alpha }$, we find that ${\lambda }_{\mathrm{max}}=-1/2+\tilde {h}/4+F\left(\tilde {\alpha },-\tilde {r}\right)+G\left(\tilde {\alpha },-\tilde {r}\right)$. Now, λmax is real, and above the oscillation threshold, parametric amplification occurs without beats, ωB = 0. In this case, the phase-locked steady-state oscillations correspond to the two 'ferromagnetic' configurations (000) or (πππ), as shown in the left panel figure 4 (in the figure specifically for $\tilde {r}=0$). As one may expect, the behaviour of the non-frustrated network is qualitatively the same as the behaviour of two coupled oscillators (section 2).

Figure 4.

Figure 4. Stability diagram of three parametric oscillators in the (X1,R, X2,R, X3,R) space. Green dots represent stable configurations of oscillations for $\tilde {r}=0$ and $\tilde {\alpha }{ >}0$. The unstable origin is represented by a black dot. The orange plane marks the X3,R = 0 plane. (Left panel) in the non-frustrated network ($\tilde {\eta }=\tilde {\alpha }$), the two possible phase-locked configurations are (000) and (πππ), which correspond to the ferromagnetic Ising solutions. The system converges to one of them as long as $\tilde {r}{< }\tilde {\alpha }$ (ωB = 0 for $\tilde {r}{< }\tilde {\alpha }$). (Right panel) in the frustrated case ($\tilde {\eta }=-\tilde {\alpha }$) there are six possible phase-locked configurations: (0ππ), (0π0), (000), (π00), (π0π), and (πππ), which correspond to the six degenerate Ising ground states. In contrast to the non-frustrated case, any infinitesimal $\tilde {r}{ >}0$ induces beats within these configurations (ωB > 0 for $0{< }\tilde {r}{< }\tilde {\alpha }$).

Standard image High-resolution image

3.3. Fully-frustrated network

We now move to the case of the fully-frustrated network, i.e., $\tilde {\eta }=-\tilde {\alpha }$ in equation (8). By proceeding as before, we find that, for $\tilde {r}{ >}\tilde {\alpha }$, the most efficient eigenvalue is ${\lambda }_{\mathrm{max}}=-1/2+\tilde {h}/4+F\left(\tilde {r},-\tilde {\alpha }\right)-G\left(\tilde {r},-\tilde {\alpha }\right)$. Now, in stark contrast to the non-frustrated case, λmax is real, and above the oscillation threshold parametric amplification occurs without beats, ωB = 0.

For $\tilde {r}{< }\tilde {\alpha }$, instead, λmax reads as ${\lambda }_{\mathrm{max}}=-1/2+\tilde {h}/4+{\mathrm{e}}^{i\enspace \pi /3}\enspace F\left(\tilde {\alpha },\tilde {r}\right)+{\mathrm{e}}^{-i\enspace \pi /3}\enspace G\left(\tilde {\alpha },\tilde {r}\right)$. Therefore, the parametric oscillation occurs with beats, where the beat frequency is

Equation (11)

One can see by inspection that, when $\tilde {r}{< }\tilde {\alpha }$, the presence of the limit cycle at threshold makes the system periodically flip between six possible phase-locked configurations, namely, (0ππ), (0π0), (000), (π00), (π0π), and (πππ), corresponding to the six degenerate ground-state configurations of the frustrated Ising model. One of these configurations stabilizes the long-time dynamics only when $\tilde {r}=0$ (see figure 4, right panel). The frustration of one of the dissipative components of the coupling has therefore a dramatic effect on the coherent dynamics of the network. In the non-frustrated case, the system at threshold converges to a phase-locked configuration for $\tilde {r}{< }\tilde {\alpha }$, and displays persistent beats otherwise. Instead, the behaviour of the fully-frustrated network is reversed with respect to the non-frustrated case: it presents beats for $\tilde {r}{< }\tilde {\alpha }$, and converges to phase-locked oscillations otherwise. For fixed $\tilde {\alpha }$, the frequency of the beats increases linearly from zero for small $\tilde {r}$, and goes to zero as $\tilde {r}$ approaches $\tilde {\alpha }$ with the same critical exponent 1/3 as before.

3.4. Interpolating case

We now expand the discussion to consider the general case of equation (8), which interpolates between the non-frustrated and the frustrated cases, and generalizes the analysis in sections 3.2 and 3.3. Here, one can find the frequency of the beats at threshold ωB from the imaginary part of the most efficient eigenvalue, for a fixed $\tilde {\alpha }$, as a function of $\tilde {r}$ and $\tilde {\eta }$, and discern regions in parameter space where phase-locked oscillations or beats are observed. We present our findings in figure 5 for both theory [panel (a)] and low-level simulation of the experiment [panel (b)], which will be discussed in detail in the next section. To ease the comparison between the analytical prediction and the low-level simulation results, we express the frequency of the beats in units of α. This makes the frequency of the beats ωB/α be a function of r/α multiplied by pure numbers and independent of g (see sections 3.2 and 3.3, and appendix C).

Figure 5.

Figure 5. Phase diagram of the three-oscillator case, showing the colormap plots of the frequency of the beats ωB/α at threshold in the r/α vs η/α plane. The vertical red lines mark the values η = ±α of non-frustrated and fully-frustrated network (see sections 3.2 and 3.3), and green lines are calculated boundaries between the phase-locking regions where ωB = 0 (dark blue) and the beating regions where ωB > 0 (other colours, see legends). (a) Analytical value from the theory presented in section 3.4 and appendix C. The two phase-locking regions are labelled by 'Ising' and 'Non Ising' depending on whether the systems behaves correctly as a PO-CIM or not (see section 3.4). (b) Low-level simulation of the experiment (see section 4). The experimental simulation agrees exceptionally well with the predicted theoretical behaviour, apart from small deviations (some noisy regions and small discrepancies near the phase boundaries) that we ascribe mostly to the difficulty of estimating the oscillation threshold in the experimental simulation for some values of the system parameters.

Standard image High-resolution image

Panel (a) of figure 5 shows ωB/α in the r/α vs η/α plane, where the red lines mark the special cases of fully frustrated and totally non-frustrated Ising coupling (η/α = ±1) that were discussed in sections 3.2 and 3.3. Green boundaries separate the regions where phase-locking is found (ωB = 0, dark blue) from those where persistent beating is manifested (ωB > 0, other colours). We see that, for η/α > 0, a region of beats is found when the energy-preserving coupling dominates over the dissipative coupling (r > α), and phase-locking is found otherwise, similar to the two-oscillator analysis and to the non-frustrated case. However, when η/α < 0, a 'tooth-shaped' region of beats appears when the dissipative coupling dominates (r < α), and phase locking is found otherwise. As r is lowered towards r = 0, the width of the tooth region of beats decreases until it collapses to a point when r = 0, at η = −α, i.e., the fully-frustrated point of section 3.3.

This peculiar behaviour of beating for small r/α may have important implications in the context of PO-CIMs. For example, fixing r and scanning η from positive to negative allows to interpolate between the ferromagnetic non-frustrated (η = α), and fully frustrated (η = −α) Ising models, where the transition to the fully-frustrated case occurs at η/α = −1. At the transition point, the Ising gap (i.e., the energy difference between the ground- and first-excited configurations) closes, causing the multiplicity of the ground state to increase (in our case, from two to six, see blue and red curves in figure 6). Our findings show that, in the system of three parametric oscillators coupled by the matrix in equation (8), any vanishingly small energy-preserving coupling induces coherent beating between the oscillators as the Ising gap becomes vanishingly small, preventing the system from converging to the Ising ground-state configuration.

Figure 6.

Figure 6. Ising energy levels relative to the ground-state energy as a function of η/α, from the classical Ising energy that the dissipative components of the coupling reflect: Eising = −ησ1σ2α(σ1σ3 + σ2σ3), where σj = ±1 is the Ising spin variable corresponding to the phase solution (respectively '0' or 'π') of xj. The three energy levels are shown in different colours for different states: blue for (000) and (πππ), red for (π00), (π0π), (0ππ), and (0π0), and green for (00π) and (ππ0). The Ising gap goes to zero at the fully-frustrated point η/α = −1, where the six states (π00), (π0π), (0ππ), (0π0), (000) and (πππ) become degenerate.

Standard image High-resolution image

In addition, we find that, while our three-oscillator system correctly behaves as a PO-CIM at the fully-frustrated point and in the phase-locking 'Ising' region in figure 5, in the other phase-locking ('Non Ising') region, the system does not yield the expected Ising behaviour. Indeed, the Ising model predicts a four-fold degenerate ground-state when η < −α (see red curve in figure 6). However, in the 'Non Ising' region in figure 5, the oscillator system slightly above the threshold converges only to two fixed points. In particular, we find the following behavour:

  • for r = 0 and r = α, the two fixed points are found on the X3,R = 0 plane, implying that limtx3(t) = 0, and the other two oscillators converge to (0π) or (π0). Clearly, the suppression of one oscillator in the long-time limit is not a valid Ising configuration;
  • for 0 < r < α, the two fixed points correspond to the states (0ππ) and (π00), which are only two of the four ground states of the Ising model;
  • for r > α, the two fixed points correspond to the states (0π0) and (π0π). These two configurations, for η < −α, are two of the four ground states, as before, but for −α < η < 0, they correspond to two of the four excited states of the Ising model.

This finding hints that frustration may cause phase-locked oscillation at threshold that however transcend the Ising description. Before concluding, we stress that the presence and details of the beating region for vanishingly small energy-preserving coupling strongly depend on the form of the coupling matrix, as well as on the number of oscillators. Indeed, while frustrated spin models with a small number of spins, simulated with PO-CIMs, have been experimentally studied in previous work [42, 43], coherent beats were not reported. This fact can be due to both the different coupling matrix considered in [42, 43], as well as to the fact that the energy-preserving coupling was possibly suppressed. A deeper analysis on how our results translates to general number of coupled oscillators and general coupling topology requires further analysis of larger spin models, which is beyond the scope of the present manuscript, and it is left for future work.

4. Numerical simulation of the experimental implementation

To corroborate the previous results and confirm our analytical predictions, we conducted a direct numerical simulation of the dynamics of the field inside a parametric-oscillator cavity with three (or more) modes, aiming to emulate as closely as possible the dynamics of a future experimental setup. In such an experiment, we intend to couple between parametrically driven modes of a multi-mode radio-frequency cavity in a fully-controlled and tunable manner. The dynamical coupling will be controlled by a field-programmable gate array (FPGA). Full details of the planned actual implementation will be reported in future work. Our numerical approach is completely distinct from the analysis presented above, and makes no explicit mention (or use) of the coupled Matheiu's equations of motion. In what follows, we first discuss our numerical procedure, and then compare the results of the simulated experiment with the analytical results presented in the previous sections.

4.1. Numerical procedure

We consider a multimode cavity where each temporal slot acts as an independent parametric oscillator, dynamically coupled to the other modes. In our simulation, the field inside the cavity propagates as illustrated in the block diagram in figure 7. At each round trip inside the cavity, the time signal is partitioned into N = 3 time slots, and parametrized as a three-dimensional vector $\mathbf{S}={\left({S}_{1},{S}_{2},{S}_{3}\right)}^{\text{T}}$. In each such interval, the field is assumed to vary slowly and is amplified independently of the other time slots. This is a reasonable assumption since the parametric gain is an instantaneous process and the pump for each time slot is uncoupled from that of the other slots. Thus, each time slot defines a distinct parametrically driven mode that without coupling evolves independently from the others. Furthermore, additive noise is fed inside the cavity at each round trip through an output coupler device to simulate thermal noise or vacuum fluctuations. During the first round trip, before being injected into the PA, the signal inside the cavity consists of noise alone.

Figure 7.

Figure 7. Schematic of the simulated experiment. At each round trip, the field inside the cavity is fed with noise and injected together with the pump field into a PA. After the parametric amplification, part of the signal is sent into the coupling mechanism that couples between the time slots, and injected back into the cavity. At the end of the round trip, part of the signal is extracted from the cavity and measured. The green backslashed boxes denote a coupler device, identified by reflection and transmission coefficients (R, T), whose values are chosen differently depending on the simulation step [(Rc, Tc) for mode coupling, or (Rout, Tout) for output coupling and noise feeding].

Standard image High-resolution image

A round trip inside the cavity is identified by the following steps: first, the pump field and the signal are injected into the PA. Importantly, since our goal is to probe the linear, near threshold, properties of the system, the pump intensity is set slightly above the oscillation threshold (section 3). At the output of the PA, the residual pump field is blocked (dark grey parallel lines in figure 7) and the signal is injected into a coupler, which splits the field according to transmission and reflection coefficients Tc = 1/4 and ${R}_{\mathrm{c}}=\sqrt{1-{T}_{\mathrm{c}}^{2}}$. The transmitted signal TcS is sent into a coupling mechanism, which implements the coupling matrix c of equation (8). In a future experiment, such a coupling mechanism will be implemented by an FPGA. After the coupling, the coupled signal F = TccS and the reflected signal RcS are combined on another coupling device, again with transmission and reflection coefficients Tc and Rc. At this step, the reflected signal TcRcS + RcF is blocked (contributing to the overall cavity losses), and the transmitted signal ${R}_{\mathrm{c}}^{2}\mathbf{S}+{T}_{\mathrm{c}}\mathbf{F}$ is fed back into the cavity. Last, the output coupler with transmission and reflection coefficients Tout = 1/2 and ${R}_{\text{out}}=\sqrt{1-{T}_{\text{out}}^{2}}$ allows to couple out the field ${T}_{\text{out}}\left({R}_{\mathrm{c}}^{2}\mathbf{S}+{T}_{\mathrm{c}}\mathbf{F}\right)$ from the cavity and analyse it both it in time and frequency, and the remaining field ${R}_{\text{out}}\left({R}_{\mathrm{c}}^{2}\mathbf{S}+{T}_{\mathrm{c}}\mathbf{F}\right)$, fed with noise, is input together with the pump field into the PA to start the next round trip.

For a given set of parameters, we run the simulation over a sufficient number of round trips nrt in order for the oscillation to reach a steady state. We then examine the steady-state dynamics to obtain the slow-varying amplitude of all oscillators and numerically extract the frequency of the beats.

4.2. Numerical results

The numerical frequency of the beats is measured by computing the fast Fourier transform (FFT) of the signal at the output to identify the frequency component with largest amplitude, for different values of r/α and/or η/α. In the numerics, the frequencies obtained from the FFT are given in units of 1/τsim, where the total simulation time is τsim = nrtτrt, and round-trip time τrt in our numerical context is an arbitrary time scale. In order to quantitatively compare the numerical results with the analytical prediction, we express the frequency of the beats in units of α (section 3.4) and use τrt = 0.0195 (see appendix D for more details).

Figure 8 shows the frequency of the beats as a function of r/α for both the numerical simulation of the experiment and the theoretical analysis. We compare the results in the special cases of the non-frustrated network (η = α) and the fully-frustrated one (η = −α). As evident, the numerical results in both cases agree well with the theoretical curves, with some slight deviations especially in the fully-frustrated cases, which we ascribe to the difficulty of correctly estimating the oscillation threshold and therefore choosing the proper value of h, due to both noise and nonlinearities. Indeed, it has been shown that pump-depletion nonlinearity tends to lower the beating frequency (divergence of the period of the beats), eventually inducing phase-locking as h is increased above the oscillation threshold [37].

Figure 8.

Figure 8. Frequency of the beats at threshold ωB/α as a function of r/α, (left) in the non-frustrated network (η = α), and (right) in the fully-frustrated network (η = −α). The numerical data (red points) obtained from the simulation of the experiment with τrt = 0.0195 are superimposed to the analytical curve [blue line, equations (10) and (11)].

Standard image High-resolution image

Last, in figure 5, panel (b), we evaluate the beating frequency ωB/α as a function of r/α and η/α, in order to verify the theoretical phase diagram in panel (a). The analytical phase boundary, marked by the green line, which is the same for both panels of figure 5, is superimposed to the numerical phase diagram to ease comparison between theory and simulated experiment. As evident, our numerical data agree exceptionally well with the analytical prediction also in the interpolating case.

5. Conclusions

We analysed the behaviour of three coupled degenerate parametric oscillators—the minimal case to study nontrivial coupling and connectivity effects. By extending our previous work on two coupled parametric oscillators, we modelled the system as three coupled Mathieu's equations, where the coupling between any two oscillators is comprised of both energy-preserving and dissipative components. We analysed the role of frustration of the dissipative component of the coupling for specific choices of the coupling matrix. We focussed in particular on two main cases of connectivity, namely, for frustrated and non-frustrated dissipative coupling. Our theoretical predictions, obtained by linearizing the effective equations of motion, were confirmed by a direct numerical simulation in time of the dynamics inside a parametric oscillator cavity, as it would be implemented in an actual experiment. The good agreement between the results obtained by these two different approaches strengthens the fact that the coupled nonlinear Mathieu's equations capture the relevant dynamics of coupled parametric oscillators.

Our main finding was that frustration of the dissipative component of the coupling has a dramatic effect on the coherent dynamics of the system. While in the non-frustrated case the system phase locks once the dissipative coupling exceeds the energy-preserving one, and behaves as a PO-CIM, in qualitative agreement with the behaviour of two coupled parametric oscillators, the frustrated case shows a totally reversed behaviour. In particular, when the dissipative coupling dominates close to full frustration, the Ising gap is vanishingly small and any vanishingly small energy-preserving coupling induces coherent beats between the (quasi) degenerate Ising configurations. For large values of the frustration parameter and for large energy-preserving coupling, the system phase locks. Interestingly, in this phase-locking region the system of three coupled oscillators does not obey the Ising description.

Our results provide an additional piece of evidence of the highly nontrivial dynamics in networks of coupled parametric oscillators, which is considerably richer than an Ising network of spins. In the view of using coupled parametric oscillators to simulate Ising models, our results hint that, in situations where the energy gap of the corresponding Ising model is very small or vanishes, the presence of even a small energy-preserving coupling between the oscillators may induce coherent beats, which should be considered in the context of PO-CIMs. Because of these intriguing implications, the theoretical and experimental investigation of large-scale networks is now highly desirable in order to see how our results translate to larger sets of parametric oscillators, as well as for more general forms of the coupling matrix [27]. We are currently planning the experimental implementation of such a large-scale network, whose analysis will be reported in future work. Ultimately, in light of our findings and the correspondence between the phase-locked behaviour and classical time-crystals, it will be important to understand how nontrivial connectivities affect the stability of classical many-body time crystals.

Acknowledgments

We thank Itzhack Dana for fruitful discussions. AP acknowledges support from the Israel Science Foundation (ISF) Grants No. 44/14 and US-Israel Binational Science Foundation (BSF) Grant No. 2017743. MCS acknowledges support from the ISF Grants No. 231/14, 1452/14, and 993/19, and BSF Grants No. 2016130 and 2018726.

Appendix A.: Derivation of the power-splitter coupling

In this appendix, we explicit the origin of the power-splitter coupling as in equation (1). In an actual physical implementation, the parametric oscillators are realized by a nonlinear cavity. The fields x1 and x2 inside each cavity propagate with a characteristic round-trip time τrt = D/v, which depends on the linear dimension D of the cavity, and on the field propagation velocity v inside the cavity.

The fields after n + 1 round-trip times, tn+1 relate to the fields after n round-trip times, tn = rt, for n = 0, 1, 2, ..., via the splitter matrix as

Equation (A.1)

We consider c11 = c22c, where 0 ⩽ c ⩽ 1 represents the transmittance coefficient of the coupling. With this choice, equation (A.1) is equivalently recast as

Equation (A.2)

and by rewriting cx1,2 = x1,2 − (1 − c)x1,2, equation (A.2) becomes

Equation (A.3)

Since x1,2(tn+1) − x1,2(tn) ∝ dot x1,2(t)/ω0, equation (A.3) can be rewritten as

Equation (A.4)

Without loss of generality, we consider c12 > 0. The terms proportional to 1 − c in equation (A.4) can be seen as loss terms that can be absorbed into the definition of g, the intrinsic loss of the cavities. Therefore, by taking the time derivative on both sides of equation (A.4), and by including this coupling in the equations of motion, equation (2) is obtained.

Appendix B.: Hamiltonian for the power-splitter coupling

In this appendix, we report the derivation of the equations of motion (2) (with β = 0 and g = 0) from the Hamilton's equations, in order to show that the coupling with c12 = c21 = r is indeed energy preserving. First, one takes the two oscillators fields, x1 and x2, and their conjugate momentum variables, p1 and p2, and defines the vectors of canonical coordinates $\mathbf{p}={\left({p}_{1},{p}_{2}\right)}^{\text{T}}$ and $\mathbf{x}={\left({x}_{1},{x}_{2}\right)}^{\text{T}}$, where T denotes the transposition. The Hamiltonian of the system is analogous to that of a particle of charge q in a two-dimensional plane, in a vector potential along the z-axis (i.e., perpendicular to the plane), given by $\mathbf{A}={\left(-{x}_{2},{x}_{1},0\right)}^{\text{T}}$, so that the corresponding effective magnetic field is $\mathbf{B}=\nabla {\times}\mathbf{A}=\hat{z}\left({\partial }_{{x}_{1}}{A}_{{x}_{2}}-{\partial }_{{x}_{2}}{A}_{{x}_{1}}\right)=\hat{z}\enspace 2$:

Equation (B.1)

or explicitly in terms of the canonical variables x1, x2, and p1, p2

Equation (B.2)

The Hamilton's equations [19] for the {dot x} variables are

Equation (B.3)

and the Hamilton's equations for the {dot p} variables are

Equation (B.4)

From equation (B.3), by deriving both sides with respect to time, one has

Equation (B.5)

By substituting dot p1 and dot p2 in the left-hand sides of equation (B.5) with the expressions in equation (B.4), one has

Equation (B.6)

but, from equation (B.3), one has p1/m = dot x1 − (q/m)x2 and p2/m = dot x2 + (q/m)x1 that, when substituted in the right-hand side of equation (B.6), yields

Equation (B.7)

from which one obtains the equations of motion

Equation (B.8)

which are indeed the equations of motion in equation (2) with c12 = c21 = r, β = 0, and g = 0, where ω0r = 2q/m.

Appendix C.: Jacobian matrix spectrum in the interpolating case

In this appendix, we report the expressions of the eigenvalues of the Jacobian matrix around the origin in the interpolating case discussed in section 3.4 (see also figure 5). By generalizing the functions in equation (9), one has

Equation (C.1)

Equation (C.2)

and then the eigenvalues of the Jacobian that can have a positive real part can be written as

Equation (C.3)

Equation (C.4)

Equation (C.5)

By finding the most efficient eigenvalue ${\lambda }_{\mathrm{max}}\left(\tilde {r},\tilde {\alpha },\tilde {\eta }\right)$ according to the usual condition (the eigenvalue with largest real part), the frequency of the beats at threshold reads ${\omega }_{\mathrm{B}}\left(\tilde {r},\tilde {\alpha },\tilde {\eta }\right)={\omega }_{0}g\enspace \vert \mathrm{I}\mathrm{m}\left[{\lambda }_{\mathrm{max}}\left(\tilde {r},\tilde {\alpha },\tilde {\eta }\right)\right]\vert $.

Appendix D.: Additional details on the choice of the round-trip time

In this appendix, we provide some details on the choice of τrt discussed in figure 8. In figure 9, we show the comparison between the theoretical expressions of the frequency of the beats at threshold [equations (10) and (11)] ωB/α, as a function of r/α, and the data from the simulated experiment for different values of the round-trip time τrt, as in the legends. The effect of changing τrt is to renormalize the frequency units for the simulated experiment. Because of the excellent agreement between theory and data from the simulated experiment in the non-frustrated case for τrt = 0.0195, this value of τrt was chosen to quantitatively match the theoretical phase diagram in figure 5. Indeed, as evident, for this τrt, theory and data are essentially overlapped.

Figure 9.

Figure 9. Beat frequency at threshold ωB as a function of r/α, (left panel) in the non-frustrated case and (right panel) fully-frustrated case, as in figure 8. The data from the simulated experiment for three different values of τrt as in the legends (purple circles for τrt = 0.0175, red crosses for τrt = 0.195, and green squares for τrt = 0.0215) are compared to the theoretical behaviour [blue line, equations (10) and (11)].

Standard image High-resolution image
Please wait… references are loading.
10.1088/1367-2630/aba573