License: arXiv.org perpetual non-exclusive license
arXiv:2312.04232v1 [physics.class-ph] 07 Dec 2023

Emergent Error Correcting States in Networks of Nonlinear Oscillators

X. **    C. G. Baker    E. Romero    N. P. Mauranyapin    T. M. F. Hirsch    W. P. Bowen [email protected]    G. I. Harris School of Mathematics and Physics, The University of Queensland, QLD 4072, Australia
Abstract

Networks of nonlinear oscillators can exhibit complex collective behaviour ranging from synchronised states to chaos. Here, we simulate the dynamics of three coupled Duffing oscillators whose multiple equilibrium states can be used for information processing and storage. Our analysis reveals that even for this small network, there is the emergence of an error correcting phase where the system autonomously corrects errors from random impulses. The system has several surprising and attractive features, including dynamic isolation of resonators exposed to extreme impulses and the ability to correct simultaneous errors. The existence of an error correcting phase opens the prospect of fault-tolerant information storage, with particular applications in nanomechanical computing.

Networked nonlinear oscillators often display intricate emergent phenomena that defy prediction from extrapolation of the behaviours of individual oscillators. A prime example is when a network of unsynchronized oscillators spontaneously achieves synchronization once their coupling strength surpasses a critical threshold [1]. Adjusting system parameters can further lead to the formation of chimera states [2], where various subsets of oscillators exhibit distinct dynamic patterns, some of which may even evolve into chaotic behavior. Understanding of these intricate behaviors has yielded valuable insights into a diverse range of physical occurrences, such as the synchronized flashing of fireflies [3], the neurological underpinnings of Parkinson’s disease [4], and the emergence of novel phases of matter in quantum systems [5]. They can also be utilised in technological applications, such as clocks with improved timing uncertainty [6], neuromorphic computing [7], and sensing networks [8].

Many physical systems exhibit nonlinear oscillations, such as superconducting circuits [9], optical parametric oscillators/amplifiers [10], and spin-waves in magnetic materials [11]. One particularly important example is the nanomechanical oscillator, which has played a critical role in the development of modern technologies, with applications ranging from ultra-precise chronometry [12] and sensing [13] to filtering [14] and imaging [15]. Nanomechanical oscillators have also been identified as a promising platform for alternative forms of computation, enabling efficient and robust operation in harsh environments, such as those in medical facilities, nuclear power plants, and outer space [16, 17, 18, 19, 20, 21].

Here, we explore the emergent behaviour of simple networks of three all-to-all coupled oscillators. We include a third-order nonlinearity; the dominant nonlinearity for many oscillators, and of particular relevance in nanomechanical systems where it is referred to as a Duffing nonlinearity. The characteristic bistability of each Duffing oscillator is used to define logical states for information processing [22]. We hypothesise that, in appropriate parameter regimes, the coupling may act as a stabiliser correcting for logical errors in individual oscillators. A computational search over the parameter space of the system confirms this hypothesis, showing that over a wide region, a collective error correcting phase emerges. That is, the proposed system returns to its initial state after being disturbed by an instantaneous perturbation. The system has several striking features that distinguish it from a traditional 3-bit majority-vote error correction scheme, including autonomous operation, enhanced performance for extreme perturbations, and the ability to probabilistically correct simultaneous errors.

Refer to caption
Figure 1: a) Schematic of a single Duffing oscillator and its applications in various computing platforms. Images of the nanomechanical circuit, superconducting circuit, and optical circuit reproduced from Ref [18, 9, 10], respectively. b) Frequency/drive response of Duffing oscillator. Solid (dashed) lines represent stable (unstable) solutions of the Duffing equation. Jumps between the stable solutions are labelled with up and down arrows. c) Time dynamics of a single Duffing oscillator subject to an SEU. The parameters of the oscillator are given by: m=1012𝑚superscript1012m=10^{-12}italic_m = 10 start_POSTSUPERSCRIPT - 12 end_POSTSUPERSCRIPT kg, Γ=105Γsuperscript105\Gamma=10^{5}roman_Γ = 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT, ω0=106subscript𝜔0superscript106\omega_{0}=10^{6}italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT, α=2×1022𝛼2superscript1022\alpha=2\times 10^{22}italic_α = 2 × 10 start_POSTSUPERSCRIPT 22 end_POSTSUPERSCRIPT ms22superscriptsuperscript𝑠22{}^{-2}s^{-2}start_FLOATSUPERSCRIPT - 2 end_FLOATSUPERSCRIPT italic_s start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT, ω=1.152×106𝜔1.152superscript106\omega=1.152\times 10^{6}italic_ω = 1.152 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT, F=5×107𝐹5superscript107F=5\times 10^{-7}italic_F = 5 × 10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT N, Δp=6×1012Δ𝑝6superscript1012\Delta p=6\times 10^{-12}roman_Δ italic_p = 6 × 10 start_POSTSUPERSCRIPT - 12 end_POSTSUPERSCRIPT kg.m.s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT. The oscillator is first initialised in the ‘0’ state, but then subject to an SEU which leads to a bit flip. The displacement is expressed in arbitrary units, such that the magnitude of a ‘1’ signal is 1111.

Despite their robustness to environmental influence, nanomechanical logic elements are susceptible to transient errors such as those arising from thermal effects, electromagnetic pulses, ionising radiation, or mechanical shock [23]. These types of transient bit-errors, which also occur frequently in conventional computers [24], are often referred to as single-event upsets (SEUs) [25, 26]. For nanomechanical logic, SEUs originate from an instantaneous impulse on the mechanical oscillator that is sufficiently large to shift the oscillator from one logical state to the other. Our discovery of an error-correcting phase, which applies broadly to third-order nonlinear oscillators, provides a means to address these errors, paving the way towards fault-tolerant nanomechanical computing.

The Duffing oscillator is one of the most widely studied nonlinear systems [27, 28], with some common applications illustrated in Fig 1a). It is described by the following equation [29]:

x¨+Γx˙+ω02x+αx3¨𝑥Γ˙𝑥superscriptsubscript𝜔02𝑥𝛼superscript𝑥3\displaystyle\ddot{x}+\Gamma\dot{x}+\omega_{0}^{2}x+\alpha x^{3}over¨ start_ARG italic_x end_ARG + roman_Γ over˙ start_ARG italic_x end_ARG + italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_x + italic_α italic_x start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT =Fmcos(ωt),absent𝐹𝑚𝜔𝑡\displaystyle=\frac{F}{m}\cos(\omega t),= divide start_ARG italic_F end_ARG start_ARG italic_m end_ARG roman_cos ( start_ARG italic_ω italic_t end_ARG ) , (1)

where, for a mechanical oscillator, x𝑥xitalic_x is the displacement, m𝑚mitalic_m the mass, ΓΓ\Gammaroman_Γ the dissipation, ω0subscript𝜔0\omega_{0}italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT the resonant frequency, α𝛼\alphaitalic_α the strength of nonlinearity and F𝐹Fitalic_F the amplitude of the sinusoidal drive provided to the system at frequency ω𝜔\omegaitalic_ω. At certain drive amplitudes and frequencies, the steady-state dynamics exhibit bistability [29], with one stable solution corresponding to high displacement and the other corresponding to low displacement [22]. This bistability is shown in Fig. 1b) for the case of spring hardening (α𝛼\alphaitalic_α >0), wherein the oscillation amplitude depends on the history of the drive frequency (left) and amplitude (right).

We use numerical methods to simulate the responses of Duffing oscillators to incident impulses, which are modelled as instantaneous changes to the momentum of the oscillators. By convention, we use high (low) displacements to represent binary ‘1’ (‘0’) signals (red/blue shaded regions in Fig. 1b)) [28, 22]. Figure 1c) illustrates the time dynamics of a single Duffing oscillator based on Eq. 1, with an impulse introduced at the time indicated by the dashed line. The oscillator is initially prepared in a ‘0’ state but the impulse causes it to latch onto the ‘1’ state, which corresponds to an erroneous bit-flip.

In conventional computing systems, a common technique to mitigate the effects of SEUs is to use majority voting logic  [30]. Within a 3-bit majority voting algorithm, each bit is replicated three times; if one bit is affected by a SEU, then the algorithm corrects this error by assuming that the majority state of the 3 bits is correct. This is traditionally achieved through bit post-processing with a logic circuit. Here, we design a similar scheme in nanomechanical logic, in which the error correction occurs autonomously, without the need for additional comparator circuitry. We consider three identical, equally forced, Duffing oscillators that are all linearly coupled (see Fig. 2a)). Each individual oscillator is driven into the bistable region and represents a single logical bit. The coupling provides an additional avenue for driving that can be either large or small depending on the amplitude of the neighbouring oscillators. The equations of motion of these coupled oscillators are:

x¨1+Γx˙1+ω02x1+αx13subscript¨𝑥1Γsubscript˙𝑥1superscriptsubscript𝜔02subscript𝑥1𝛼superscriptsubscript𝑥13\displaystyle\ddot{x}_{1}+\Gamma\dot{x}_{1}+\omega_{0}^{2}x_{1}+\alpha x_{1}^{3}over¨ start_ARG italic_x end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + roman_Γ over˙ start_ARG italic_x end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_α italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT =Fmcos(ωt)+βx2+βx3absent𝐹𝑚𝜔𝑡𝛽subscript𝑥2𝛽subscript𝑥3\displaystyle=\frac{F}{m}\cos(\omega t)+\beta x_{2}+\beta x_{3}= divide start_ARG italic_F end_ARG start_ARG italic_m end_ARG roman_cos ( start_ARG italic_ω italic_t end_ARG ) + italic_β italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_β italic_x start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT (2)
x¨2+Γx˙2+ω02x2+αx23subscript¨𝑥2Γsubscript˙𝑥2superscriptsubscript𝜔02subscript𝑥2𝛼superscriptsubscript𝑥23\displaystyle\ddot{x}_{2}+\Gamma\dot{x}_{2}+\omega_{0}^{2}x_{2}+\alpha x_{2}^{3}over¨ start_ARG italic_x end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + roman_Γ over˙ start_ARG italic_x end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_α italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT =Fmcos(ωt)+βx1+βx3absent𝐹𝑚𝜔𝑡𝛽subscript𝑥1𝛽subscript𝑥3\displaystyle=\frac{F}{m}\cos(\omega t)+\beta x_{1}+\beta x_{3}= divide start_ARG italic_F end_ARG start_ARG italic_m end_ARG roman_cos ( start_ARG italic_ω italic_t end_ARG ) + italic_β italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_β italic_x start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT (3)
x¨3+Γx˙3+ω02x3+αx33subscript¨𝑥3Γsubscript˙𝑥3superscriptsubscript𝜔02subscript𝑥3𝛼superscriptsubscript𝑥33\displaystyle\ddot{x}_{3}+\Gamma\dot{x}_{3}+\omega_{0}^{2}x_{3}+\alpha x_{3}^{3}over¨ start_ARG italic_x end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT + roman_Γ over˙ start_ARG italic_x end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT + italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT + italic_α italic_x start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT =Fmcos(ωt)+βx1+βx2absent𝐹𝑚𝜔𝑡𝛽subscript𝑥1𝛽subscript𝑥2\displaystyle=\frac{F}{m}\cos(\omega t)+\beta x_{1}+\beta x_{2}= divide start_ARG italic_F end_ARG start_ARG italic_m end_ARG roman_cos ( start_ARG italic_ω italic_t end_ARG ) + italic_β italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_β italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT (4)

where β𝛽\betaitalic_β is the coupling rate between oscillators, and x1subscript𝑥1x_{1}italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, x2subscript𝑥2x_{2}italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, and x3subscript𝑥3x_{3}italic_x start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT represent the displacement of oscillators 1111, 2222 and 3333, respectively.

Refer to caption
Figure 2: a) Schematic of mechanical error correction system, composed of three coupled Duffing oscillators. This functions as a majority voting system and can correct for single SEUs. b) Time dynamics of error correction device. The coupled oscillators are initialised in their ‘0’ states and after 4 periods of oscillation, an impulse is applied to the third oscillator. The third oscillator temporarily transitions into its ‘1’ state, but quickly equilibrates back. The parameters of the oscillator are given by: m=1012𝑚superscript1012m=10^{-12}italic_m = 10 start_POSTSUPERSCRIPT - 12 end_POSTSUPERSCRIPT kg, Γ=105Γsuperscript105\Gamma=10^{5}roman_Γ = 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT, ω0=106subscript𝜔0superscript106\omega_{0}=10^{6}italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT , α=2×1022𝛼2superscript1022\alpha=2\times 10^{22}italic_α = 2 × 10 start_POSTSUPERSCRIPT 22 end_POSTSUPERSCRIPT m.2s2{}^{-2}.s^{-2}start_FLOATSUPERSCRIPT - 2 end_FLOATSUPERSCRIPT . italic_s start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT, ω=1.152×106𝜔1.152superscript106\omega=1.152\times 10^{6}italic_ω = 1.152 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT, F=1.23×106𝐹1.23superscript106F=1.23\times 10^{-6}italic_F = 1.23 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT N, Δp=6×1012Δ𝑝6superscript1012\Delta p=6\times 10^{-12}roman_Δ italic_p = 6 × 10 start_POSTSUPERSCRIPT - 12 end_POSTSUPERSCRIPT kg.m.s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT, β=2×1011𝛽2superscript1011\beta=2\times 10^{11}italic_β = 2 × 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT s22{}^{-2}start_FLOATSUPERSCRIPT - 2 end_FLOATSUPERSCRIPT. c) Phase map of error correction device. The map is divided into four main regions, ‘1’ bias, ‘0’ bias, initialise and error correction. This map is produced using the same parameters as b).

To determine the collective behaviour of the proposed error correcting network, we simulate its time dynamics for a range of different parameters and initial conditions. We search for emergent phases where the collective dynamics exhibit only two steady states (’111’ and ’000’), irrespective of when impulses occur and insensitive to small deviations in system parameters. We find that for specific sets of parameters, error correcting behaviour is possible. One example is shown in Fig. 2b), where each oscillator is initially prepared in the ‘0’ state, before oscillator 3 is subjected to an impulse of the same magnitude as that in Fig. 1c) (dashed line). We see that the impulse causes the amplitude of oscillator 3 to temporarily reach the ‘1’ state, however, it does not latch to this state. After an equilibration period, all oscillators settle back into the ‘0’ state. Similar error correcting dynamics are seen when the oscillators are initialised in the ’1’ states (see Supplementary Information).

Next, we repeat these simulations varying the drive frequency, drive amplitude and time of impulse. We find the behaviour of the coupled system can be categorized into four distinct phases (see Fig. 2c)). When the drive is strong and near resonance, all oscillators evolve into their ‘1’ states. Conversely, when the oscillators are provided with weak drives, or are driven far away from resonance, they always evolve into their ‘0’ states. We term these two regions ‘1’ bias and ‘0’ bias, respectively. In between the two bias regions, the oscillators can be initialized in either their collective ‘1’ or ‘0’ states, depending on the history of the applied forcing. We call this the initialise region. It should be noted that operating in this region does not guarantee error correction. We find that, for some parameters and impulse timings, it is possible for a single impulse to cause the system to transition between collective states. However, there remains a subspace that is guaranteed to correct all single impulses. This region is labelled error correction in Fig. 2c).

For sufficiently high coupling rates, the error correction region can be made to be large, enabling stable operation. For instance, Fig. 2c) was generated with a coupling rate equal to twice the intrinsic dissipation (i.e. β/ω0=2Γ𝛽subscript𝜔02Γ\beta/\omega_{0}=2\Gammaitalic_β / italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 2 roman_Γ). Here, near the centre of the error correction region, error correction is maintained even with large deviations in drive force. For instance, at ω/ω0=1.152𝜔subscript𝜔01.152\omega/\omega_{0}=1.152italic_ω / italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 1.152, error correction is predicted to occur for drive forces in the range 4.3FcritF6.8Fcrit4.3subscript𝐹𝑐𝑟𝑖𝑡𝐹6.8subscript𝐹𝑐𝑟𝑖𝑡4.3F_{crit}\leq F\leq 6.8F_{crit}4.3 italic_F start_POSTSUBSCRIPT italic_c italic_r italic_i italic_t end_POSTSUBSCRIPT ≤ italic_F ≤ 6.8 italic_F start_POSTSUBSCRIPT italic_c italic_r italic_i italic_t end_POSTSUBSCRIPT. The size of the error correction region decreases with decreasing coupling rate, eventually disappearing altogether in the weak-coupling regime where β/ω0<Γ𝛽subscript𝜔0Γ\beta/\omega_{0}<\Gammaitalic_β / italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT < roman_Γ.

.

The error correction of all single impulses can be further understood by considering the instantaneous energy of the system as a function of time. Figure 3a) shows that the energy of the perturbed oscillator (purple line) increases rapidly due to the impulse, as expected.

The two unperturbed oscillators then extract the excess energy and dissipate it to the environment, bringing the perturbed oscillator back into its original state. As the impulse amplitude is increased, the ability of the unperturbed oscillators to dissipate the excess energy is reduced, causing failures in error correction. This is shown in Fig. 3b) where impulses that cause the system momentum to exceed p>1.42pmax,‘1’𝑝1.42subscript𝑝max,‘1’p>1.42p_{\text{max,`1'}}italic_p > 1.42 italic_p start_POSTSUBSCRIPT max,‘1’ end_POSTSUBSCRIPT result in failures of error correction (where pmax,‘1’subscript𝑝max,‘1’p_{\text{max,`1'}}italic_p start_POSTSUBSCRIPT max,‘1’ end_POSTSUBSCRIPT is the maximum momentum of the equilibrated ‘1’ state).

Unexpectedly, we find that the system also exhibits error correction of extremely large impulses (i.e. impulse energy large enough to transition all oscillators into the error state). This occurs due to an effect we call dynamic decoupling. An extremely large impulse will temporarily increase the amplitude of the perturbed oscillator, nonlinearly shifting its oscillation frequency away from the remaining oscillators. This decouples the perturbed oscillator from the others and precludes efficient energy exchange between them. The perturbed oscillator then dissipates the excess energy to the environment before reducing its amplitude and re-establishing coupling. The dashed line in Fig 3a) indicates the approximate energy threshold above which dynamic decoupling starts to occur, defined as the point when the Duffing induced frequency shift exceeds the coupling rate between oscillators β/ω0𝛽subscript𝜔0\leavevmode\nobreak\ \beta/\omega_{0}italic_β / italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (See Supplemental Information). Indeed, dynamic decoupling becomes more pronounced for larger impulse amplitudes. As a result, for our set of parameters, there are two regimes that allow perfect error correction of SEUs; impulse amplitudes in the range 0<Δp/pmax,‘1’<1.420Δ𝑝subscript𝑝max,‘1’1.420<\Delta p/p_{\text{max,`1'}}<1.420 < roman_Δ italic_p / italic_p start_POSTSUBSCRIPT max,‘1’ end_POSTSUBSCRIPT < 1.42 and those with Δp/pmax,‘1’>8.5Δ𝑝subscript𝑝max,‘1’8.5\Delta p/p_{\text{max,`1'}}>8.5roman_Δ italic_p / italic_p start_POSTSUBSCRIPT max,‘1’ end_POSTSUBSCRIPT > 8.5. This is shown in Fig. 3b) where the failure probability is zero for both small (main figure) and large (inset) impulse amplitudes.

Refer to caption
Figure 3: Single event upset on a 3-bit majority system. a) Instantaneous energy of coupled system. Oscillators 1 and 2 are represented by the yellow trajectory and oscillator 3 is represented by the purple trajectory. The phase of the impulse relative to motion of the oscillator is indicated by the diagram on the top right corner of the figure. Since the amplitude of oscillator 3 is momentarily increased, its resonance frequency is up-shifted and it decouples from the other oscillators. The energy required for this to occur is indicated by a horizontal dashed line, and the region of decoupling is represented by grey shading. b) Probability of error-correction failure (i.e. error occurs from impulse) with increasing amplitude of impulse. The horizontal axis is normalised to the maximum momentum of the oscillator when in the ‘1’ state. Inset: probability of error-correction failure for a larger range of impulse amplitudes. Grey region is the range of the main figure. Note that at high impulse amplitudes (i.e. Δp>8.5pmax,‘1’Δ𝑝8.5subscript𝑝max,‘1’\Delta p>8.5p_{\text{max,`1'}}roman_Δ italic_p > 8.5 italic_p start_POSTSUBSCRIPT max,‘1’ end_POSTSUBSCRIPT) the error-correction is always successful.

The error correction protocol can perfectly correct single impulses, but we expect it to be susceptible to multiple simultaneous impulses. To explore this, we run simulations in which each oscillator has a fixed probability, Pkicksubscript𝑃𝑘𝑖𝑐𝑘P_{kick}italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT, of experiencing an impulse at a given time. Consequently, there is a finite probability of two or all three oscillators experiencing impulses simultaneously. We define an event as the occurrence of one or more impulses. If an event occurs and causes an error, then it is recorded and the simulation is reinitialised (see Supplementary Information for details). This allows one to determine the probability of failure for a given event probability. When this process is repeated for a range of event probabilities, the probability of failure (normalised by the event probability itself) is obtained.

We compare to equivalent simulations on a single oscillator (red in Fig. 4). The probability of failure per event for the single oscillator is constant at approximately 91%. The 9%percent99\%9 % success rate can be attributed to a portion of impulses opposing the momentum of the oscillator and therefore having a reduced effect.

The performance of the system was investigated with respect to an ideal majority voting scheme. Three way majority voting logic should correct all single impulses, but fail if two or more oscillators experience them. One can predict the theoretical failure rate, so defined, from a Binomial distribution for a given impulse probability (see Supplementary Information). When comparing the three oscillator system (blue dots) to the theoretical performance of a conventional majority voting scheme (blue dashed line), Fig. 4 shows that the coupled system surprisingly exhibits superior error-correcting performance.

Interestingly, the coupled system corrects all single impulses and 65% of the double impulses. We believe that the ability of the coupled oscillators to correct a double impulse is related to non-trivial transients of the system. Quite generally, as the oscillator transitions between its two stable states, it must re-phase its oscillation with respect to the external drive. In this situation, the displacement from an impulse may not have the correct phase to enable latching onto the stable state, causing the impulse energy to be extracted from the system through dissipation and the external drive. In some sense, the state of each oscillator is encoded twice in the dynamics, once in its amplitude and again in its phase. This creates additional redundancy of the logical bit, enabling double errors to be corrected.

Refer to caption
Figure 4: Simulated probability of system failure per event as a function of event probability per time interval. Red and blue dots represent the simulation results of single and coupled systems, respectively. Red and blue dashed lines represent the predicted model for single oscillator and majority vote of three independent oscillators, respectively. Red and blue solid lines represent the corrected model for single oscillator and coupled oscillators, respectively. Note: simulation parameters are the same as in Figure 2.

In this letter, we have explored the emergence of an error correcting phase in systems of coupled nonlinear oscillators. Using numerical simulations, we identify four distinct regions of behaviour in parameter space. We show that one of these regions enables autonomous error-correction from randomly occurring impulses, removing the need for additional logic operations to perform a majority-vote. Interestingly, error-correction is greatly enhanced for large impulses due to a dynamic decoupling of the perturbed oscillator, enabling error correction for impulses that are over ten times larger than the momentum of the ‘1’ state. Furthermore, the system is capable of correcting two simultaneous errors 65% of the time due to transient effects that de-phase the oscillators with respect to the drive.

Our work shows that the complexity of networks of nonlinear oscillators can be harnessed to perform useful computational tasks, and potentially enable fault-tolerant nanomechanical computing.

This research was primarily funded by the Australian Research Council and Lockheed Martin Corporation through the Australian Research Council Linkage Grant No. LP190101159. Support was also provided by the Australian Research Council Centre of Excellence for Engineered Quantum Systems (No. CE170100009). G.I.H. and C.G.B. acknowledge their Australian Research Council grants (No. DE210100848 and No. DE190100318), respectively.

References

  • Restrepo et al. [2005] J. G. Restrepo, E. Ott, and B. R. Hunt, Phys. Rev. E 71, 036151 (2005).
  • Abrams and Strogatz [2004] D. M. Abrams and S. H. Strogatz, Phys. Rev. Lett. 93, 174102 (2004).
  • Sarfati and Peleg [2022] R. Sarfati and O. Peleg, Science Advances 8, eadd6690 (2022).
  • Uhlhaas and Singer [2006] P. J. Uhlhaas and W. Singer, Neuron 52, 155 (2006).
  • Sakurai et al. [2021] A. Sakurai, V. M. Bastidas, W. J. Munro, and K. Nemoto, Phys. Rev. Lett. 126, 120606 (2021).
  • Kómár et al. [2014] P. Kómár, E. M. Kessler, M. Bishof, L. Jiang, A. S. Sørensen, J. Ye, and M. D. Lukin, Nature Physics 10, 582 (2014).
  • Torrejon et al. [2017] J. Torrejon, M. Riou, F. A. Araujo, S. Tsunegi, G. Khalsa, D. Querlioz, P. Bortolotti, V. Cros, K. Yakushiji, A. Fukushima, H. Kubota, S. Yuasa, M. D. Stiles, and J. Grollier, Nature 547, 428 (2017).
  • Xu et al. [2020] L. Xu, S. Wang, Z. Jiang, and X. Wei, Microsystems & Nanoengineering 6, 63 (2020).
  • Boaknin et al. [2007] E. Boaknin, V. E. Manucharyan, S. Fissette, M. Metcalfe, L. Frunzio, R. Vijay, I. Siddiqi, A. Wallraff, R. J. Schoelkopf, and M. Devoret, Dispersive microwave bifurcation of a superconducting resonator cavity incorporating a josephson junction (2007), arXiv:cond-mat/0702445 [cond-mat.supr-con] .
  • MOSS and EGGLETON [2008] D. J. MOSS and B. J. EGGLETON, Advances in Information Optics and Photonics, Vol. 6 (SPIE, 2008) pp. 657–686.
  • Dreyer et al. [2022] R. Dreyer, A. F. Schäffer, H. G. Bauer, N. Liebing, J. Berakdar, and G. Woltersdorf, Nature Communications 13, 4939 (2022).
  • Gavartin et al. [2013] E. Gavartin, P. Verlot, and T. J. Kippenberg, Nature communications 4, 1 (2013).
  • Zhang and Turner [2005] W. Zhang and K. L. Turner, Sensors and Actuators A: Physical 122, 23 (2005).
  • Delsing et al. [2019] P. Delsing, A. N. Cleland, M. J. Schuetz, J. Knörzer, G. Giedke, J. I. Cirac, K. Srinivasan, M. Wu, K. C. Balram, C. Bäuerle, et al., Journal of Physics D: Applied Physics 52, 353001 (2019).
  • Rützel et al. [2003] S. Rützel, S. I. Lee, and A. Raman, Proceedings of the Royal Society of London. Series A: Mathematical, Physical and Engineering Sciences 459, 1925 (2003).
  • Coulombe et al. [2017] J. C. Coulombe, M. C. A. York, and J. Sylvestre, PLOS ONE 12, 1 (2017).
  • Song et al. [2019] Y. Song, R. M. Panas, S. Chizari, L. A. Shaw, J. A. Jackson, J. B. Hopkins, and A. J. Pascall, Nature communications 10, 882 (2019).
  • Wenzler et al. [2014] J.-S. Wenzler, T. Dunn, T. Toffoli, and P. Mohanty, Nano Lett. 14, 89 (2014).
  • Yamaguchi et al. [2011] H. Yamaguchi, K. Nishiguchi, E. Flurin, A. Fujiwara, and I. Mahboob, Nature communications 2, 198 (2011).
  • Yao and Hikihara [2014a] A. Yao and T. Hikihara, Applied physics letters 105, 123104 (2014a).
  • Roukes [2004] M. Roukes, in IEDM Technical Digest. IEEE International Electron Devices Meeting, 2004 (IEEE, 2004) pp. 539–542.
  • Romero et al. [2022] E. Romero, N. P. Mauranyapin, T. M. F. Hirsch, R. Kalra, C. G. Baker, G. I. Harris, and W. P. Bowen, Scalable nanomechanical logic gate (2022), arXiv:cond-mat/2206.11661v2 [cond-mat.mes-hall] .
  • Lee et al. [2023] Y.-B. Lee, M.-H. Kang, P.-K. Choi, S.-H. Kim, T.-S. Kim, S.-Y. Lee, and J.-B. Yoon, Nature Communications 14, 460 (2023).
  • Schroeder et al. [2009] B. Schroeder, E. Pinheiro, and W.-D. Weber, ACM SIGMETRICS Performance Evaluation Review 37, 193 (2009).
  • Karnik and Hazucha [2004] T. Karnik and P. Hazucha, IEEE Transactions on Dependable and Secure Computing 1, 128 (2004).
  • Ziegler and Lanford [1979] J. F. Ziegler and W. A. Lanford, Science 206, 776 (1979).
  • Yao and Hikihara [2014b] A. Yao and T. Hikihara, Applied physics letters 105, 123104 (2014b).
  • Tadokoro and Tanaka [2021] Y. Tadokoro and H. Tanaka, Phys. Rev. Appl. 15, 024058 (2021).
  • Schmid et al. [2016] S. Schmid, L. G. Villanueva, and M. L. Roukes, Fundamentals of Nanomechanical Resonators (Springer International Publishing : Imprint: Springer, 2016).
  • She and McElvain [2009] X. She and K. McElvain, IEEE transactions on nuclear science 56, 2443 (2009).

Supplemental Material

A Strong Coupling Threshold of Three Coupled Harmonic Oscillators

The equations of motion for three coupled harmonic oscillators are given by:

x¨1+ω02x1subscript¨𝑥1superscriptsubscript𝜔02subscript𝑥1\displaystyle\ddot{x}_{1}+\omega_{0}^{2}x_{1}over¨ start_ARG italic_x end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT =βx2+βx3absent𝛽subscript𝑥2𝛽subscript𝑥3\displaystyle=\beta x_{2}+\beta x_{3}= italic_β italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_β italic_x start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT (5)
x¨2+ω02x2subscript¨𝑥2superscriptsubscript𝜔02subscript𝑥2\displaystyle\ddot{x}_{2}+\omega_{0}^{2}x_{2}over¨ start_ARG italic_x end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT =βx1+βx3absent𝛽subscript𝑥1𝛽subscript𝑥3\displaystyle=\beta x_{1}+\beta x_{3}= italic_β italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_β italic_x start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT (6)
x¨3+ω02x3subscript¨𝑥3superscriptsubscript𝜔02subscript𝑥3\displaystyle\ddot{x}_{3}+\omega_{0}^{2}x_{3}over¨ start_ARG italic_x end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT + italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT =βx1+βx2absent𝛽subscript𝑥1𝛽subscript𝑥2\displaystyle=\beta x_{1}+\beta x_{2}= italic_β italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_β italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT (7)

where ω0subscript𝜔0\omega_{0}italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the resonant frequency of an uncoupled oscillator and β𝛽\betaitalic_β is the coupling strength. Assuming that the solution takes the form of xj(t)=xj0exp(iωt)subscript𝑥𝑗𝑡superscriptsubscript𝑥𝑗0𝑖𝜔𝑡x_{j}(t)=x_{j}^{0}\exp(-i\omega t)italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) = italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT roman_exp ( start_ARG - italic_i italic_ω italic_t end_ARG ), the equations of motion can be rewritten in matrix form:

𝑨x𝑨𝑥\displaystyle\bm{A}\ xbold_italic_A italic_x =𝟎absent0\displaystyle=\bm{0}= bold_0
[ω2+ω02βββω2+ω02βββω2+ω02][x10x20x30]matrixsuperscript𝜔2superscriptsubscript𝜔02𝛽𝛽𝛽superscript𝜔2superscriptsubscript𝜔02𝛽𝛽𝛽superscript𝜔2superscriptsubscript𝜔02matrixsuperscriptsubscript𝑥10superscriptsubscript𝑥20superscriptsubscript𝑥30\displaystyle\begin{bmatrix}-\omega^{2}+\omega_{0}^{2}&-\beta&-\beta\\ -\beta&-\omega^{2}+\omega_{0}^{2}&-\beta\\ -\beta&-\beta&-\omega^{2}+\omega_{0}^{2}\end{bmatrix}\begin{bmatrix}x_{1}^{0}% \\ x_{2}^{0}\\ x_{3}^{0}\end{bmatrix}[ start_ARG start_ROW start_CELL - italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL start_CELL - italic_β end_CELL start_CELL - italic_β end_CELL end_ROW start_ROW start_CELL - italic_β end_CELL start_CELL - italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL start_CELL - italic_β end_CELL end_ROW start_ROW start_CELL - italic_β end_CELL start_CELL - italic_β end_CELL start_CELL - italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL end_ROW end_ARG ] [ start_ARG start_ROW start_CELL italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL italic_x start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT end_CELL end_ROW end_ARG ] =[000]absentmatrix000\displaystyle=\begin{bmatrix}0\\ 0\\ 0\end{bmatrix}= [ start_ARG start_ROW start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL end_ROW end_ARG ]

The characteristic equation of the matrix 𝑨𝑨\bm{A}bold_italic_A is given by:

2β3+3β2ω2ω63β2ω02+3ω4ω023ω02ω4+ω062superscript𝛽33superscript𝛽2superscript𝜔2superscript𝜔63superscript𝛽2superscriptsubscript𝜔023superscript𝜔4superscriptsubscript𝜔023superscriptsubscript𝜔02superscript𝜔4superscriptsubscript𝜔06\displaystyle-2\beta^{3}+3\beta^{2}\omega^{2}-\omega^{6}-3\beta^{2}\omega_{0}^% {2}+3\omega^{4}\omega_{0}^{2}-3\omega_{0}^{2}\omega^{4}+\omega_{0}^{6}- 2 italic_β start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT + 3 italic_β start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_ω start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT - 3 italic_β start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 3 italic_ω start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 3 italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ω start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT + italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT =0absent0\displaystyle=0= 0

By solving for ω2superscript𝜔2\omega^{2}italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, one obtains:

ω12superscriptsubscript𝜔12\displaystyle\omega_{1}^{2}italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT =ω022βabsentsuperscriptsubscript𝜔022𝛽\displaystyle=\omega_{0}^{2}-2\beta= italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 2 italic_β
ω22superscriptsubscript𝜔22\displaystyle\omega_{2}^{2}italic_ω start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT =ω02+βabsentsuperscriptsubscript𝜔02𝛽\displaystyle=\omega_{0}^{2}+\beta= italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_β
ω32superscriptsubscript𝜔32\displaystyle\omega_{3}^{2}italic_ω start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT =ω02+βabsentsuperscriptsubscript𝜔02𝛽\displaystyle=\omega_{0}^{2}+\beta= italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_β

The above expressions describe the frequencies of the three fundamental modes of oscillation of the system. The mode utilised by the proposed error correction device was determined to be the lowest frequency mode. The frequency shift introduced by coupling for this particular mode is given by:

ω1subscript𝜔1\displaystyle\omega_{1}italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT =ω022βabsentsuperscriptsubscript𝜔022𝛽\displaystyle=\sqrt{\omega_{0}^{2}-2\beta}= square-root start_ARG italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 2 italic_β end_ARG
ω0βω0absentsubscript𝜔0𝛽subscript𝜔0\displaystyle\approx\omega_{0}-\frac{\beta}{\omega_{0}}≈ italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - divide start_ARG italic_β end_ARG start_ARG italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG

For the oscillators to be strongly coupled, the frequency shift must be greater than the dissipation rate (ΓΓ\Gammaroman_Γ, as defined in the main text):

ω0ω1βω0Γsubscript𝜔0subscript𝜔1𝛽subscript𝜔0Γ\displaystyle\omega_{0}-\omega_{1}\approx\frac{\beta}{\omega_{0}}\geq\Gammaitalic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≈ divide start_ARG italic_β end_ARG start_ARG italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ≥ roman_Γ

B Decoupling energy of Three Coupled Duffing Oscillators

In comparison, the equations of motion of three damped and driven coupled Duffing oscillators with mass ’m𝑚mitalic_m’ and dam** rate ’γ𝛾\gammaitalic_γ’ are given by:

x¨1+γx˙1+ω02x1+αx1 3subscript¨𝑥1𝛾subscript˙𝑥1superscriptsubscript𝜔02subscript𝑥1𝛼superscriptsubscript𝑥13\displaystyle\ddot{x}_{1}+\gamma\dot{x}_{1}+\omega_{0}^{2}x_{1}+\alpha x_{1}^{% \ 3}over¨ start_ARG italic_x end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_γ over˙ start_ARG italic_x end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_α italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT =Fmcos(ωt)+βx2+βx3absent𝐹𝑚𝜔𝑡𝛽subscript𝑥2𝛽subscript𝑥3\displaystyle=\frac{F}{m}\cos(\omega t)+\beta x_{2}+\beta x_{3}= divide start_ARG italic_F end_ARG start_ARG italic_m end_ARG roman_cos ( start_ARG italic_ω italic_t end_ARG ) + italic_β italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_β italic_x start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT (8)
x¨2+γx˙2+ω02x2+αx2 3subscript¨𝑥2𝛾subscript˙𝑥2superscriptsubscript𝜔02subscript𝑥2𝛼superscriptsubscript𝑥23\displaystyle\ddot{x}_{2}+\gamma\dot{x}_{2}+\omega_{0}^{2}x_{2}+\alpha x_{2}^{% \ 3}over¨ start_ARG italic_x end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_γ over˙ start_ARG italic_x end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_α italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT =Fmcos(ωt)+βx1+βx3absent𝐹𝑚𝜔𝑡𝛽subscript𝑥1𝛽subscript𝑥3\displaystyle=\frac{F}{m}\cos(\omega t)+\beta x_{1}+\beta x_{3}= divide start_ARG italic_F end_ARG start_ARG italic_m end_ARG roman_cos ( start_ARG italic_ω italic_t end_ARG ) + italic_β italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_β italic_x start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT (9)
x¨3+γx˙3+ω02x3+αx3 3subscript¨𝑥3𝛾subscript˙𝑥3superscriptsubscript𝜔02subscript𝑥3𝛼superscriptsubscript𝑥33\displaystyle\ddot{x}_{3}+\gamma\dot{x}_{3}+\omega_{0}^{2}x_{3}+\alpha x_{3}^{% \ 3}over¨ start_ARG italic_x end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT + italic_γ over˙ start_ARG italic_x end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT + italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_x start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT + italic_α italic_x start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT =Fmcos(ωt)+βx1+βx2absent𝐹𝑚𝜔𝑡𝛽subscript𝑥1𝛽subscript𝑥2\displaystyle=\frac{F}{m}\cos(\omega t)+\beta x_{1}+\beta x_{2}= divide start_ARG italic_F end_ARG start_ARG italic_m end_ARG roman_cos ( start_ARG italic_ω italic_t end_ARG ) + italic_β italic_x start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_β italic_x start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT (10)

For a single Duffing oscillator, the shift in resonant frequency as a function of displacement is given by the backbone equation ([1, 2]):

ωbackbone=ω02(1+3α4ω02x2)subscript𝜔𝑏𝑎𝑐𝑘𝑏𝑜𝑛𝑒superscriptsubscript𝜔0213𝛼4superscriptsubscript𝜔02superscript𝑥2\displaystyle\omega_{backbone}=\sqrt{\omega_{0}^{2}(1+\frac{3\alpha}{4\omega_{% 0}^{2}}x^{2})}italic_ω start_POSTSUBSCRIPT italic_b italic_a italic_c italic_k italic_b italic_o italic_n italic_e end_POSTSUBSCRIPT = square-root start_ARG italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 1 + divide start_ARG 3 italic_α end_ARG start_ARG 4 italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG (11)

where ω0subscript𝜔0\omega_{0}italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the resonant frequency of a single oscillator and α𝛼\alphaitalic_α is the Duffing coefficient, which quantifies the degree of non-linearity. We consider the condition of decoupling to be where the frequency shift induced by the kick is greater than that required to sustain strong coupling (as defined in the previous section). This can be written as:

ωdωsubscript𝜔𝑑𝜔\displaystyle\omega_{d}-\omegaitalic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT - italic_ω β/ω0absent𝛽subscript𝜔0\displaystyle\geq\beta/\omega_{0}≥ italic_β / italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (12)

where ωdsubscript𝜔𝑑\omega_{d}italic_ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT is resonance frequency required for decoupling. That is, when comparing to the other oscillators, the shift in resonance frequency of the kicked oscillator is greater or equal to the coupling rate. Assuming all the oscillators are initialised their ‘0’ states with steady state displacement x0subscript𝑥0x_{0}italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, the displacement required for decoupling, xdsubscript𝑥𝑑x_{d}italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT, is then given by:

ω02+34αxd2ω02+34αx02superscriptsubscript𝜔0234𝛼superscriptsubscript𝑥𝑑2superscriptsubscript𝜔0234𝛼superscriptsubscript𝑥02\displaystyle\sqrt{\omega_{0}^{2}+\frac{3}{4}\alpha x_{d}^{2}}-\sqrt{\omega_{0% }^{2}+\frac{3}{4}\alpha x_{0}^{2}}square-root start_ARG italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + divide start_ARG 3 end_ARG start_ARG 4 end_ARG italic_α italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - square-root start_ARG italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + divide start_ARG 3 end_ARG start_ARG 4 end_ARG italic_α italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG =β/ω0absent𝛽subscript𝜔0\displaystyle=\beta/\omega_{0}= italic_β / italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT
ω02+34αxd2superscriptsubscript𝜔0234𝛼superscriptsubscript𝑥𝑑2\displaystyle\omega_{0}^{2}+\frac{3}{4}\alpha x_{d}^{2}italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + divide start_ARG 3 end_ARG start_ARG 4 end_ARG italic_α italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT =(β/ω0+ω02+34αx02)2absentsuperscript𝛽subscript𝜔0superscriptsubscript𝜔0234𝛼superscriptsubscript𝑥022\displaystyle=\left(\beta/\omega_{0}+\sqrt{\omega_{0}^{2}+\frac{3}{4}\alpha x_% {0}^{2}}\ \right)^{2}= ( italic_β / italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + square-root start_ARG italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + divide start_ARG 3 end_ARG start_ARG 4 end_ARG italic_α italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT
xd2superscriptsubscript𝑥𝑑2\displaystyle x_{d}^{2}italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT =43α[(β/ω0+ω02+34αx02)2ω02]absent43𝛼delimited-[]superscript𝛽subscript𝜔0superscriptsubscript𝜔0234𝛼superscriptsubscript𝑥022superscriptsubscript𝜔02\displaystyle=\frac{4}{3\alpha}\left[\left(\beta/\omega_{0}+\sqrt{\omega_{0}^{% 2}+\frac{3}{4}\alpha x_{0}^{2}}\right)^{2}-\omega_{0}^{2}\right]= divide start_ARG 4 end_ARG start_ARG 3 italic_α end_ARG [ ( italic_β / italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + square-root start_ARG italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + divide start_ARG 3 end_ARG start_ARG 4 end_ARG italic_α italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ]

The energy of each oscillator is given by the sum of its potential and kinetic energy:

E𝐸\displaystyle Eitalic_E =PE+KEabsent𝑃𝐸𝐾𝐸\displaystyle=PE+KE= italic_P italic_E + italic_K italic_E
=12kx2+14mαx4+12mv2absent12𝑘superscript𝑥214𝑚𝛼superscript𝑥412𝑚superscript𝑣2\displaystyle=\frac{1}{2}kx^{2}+\frac{1}{4}m\alpha x^{4}+\frac{1}{2}mv^{2}= divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_k italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + divide start_ARG 1 end_ARG start_ARG 4 end_ARG italic_m italic_α italic_x start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_m italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT

Here, k𝑘kitalic_k is the natural spring constant, given by mω02𝑚superscriptsubscript𝜔02m\omega_{0}^{2}italic_m italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, and xdsubscript𝑥𝑑x_{d}italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT is the steady state displacement required for decoupling. When the oscillator reaches its maximum displacement, it attains its maximum potential energy, but has zero kinetic energy. Therefore, we can calculate the energy required for decoupling by calculating the potential energy at xdsubscript𝑥𝑑x_{d}italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT:

Edsubscript𝐸𝑑\displaystyle E_{d}italic_E start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT =12kxd2+14mαxd4absent12𝑘superscriptsubscript𝑥𝑑214𝑚𝛼superscriptsubscript𝑥𝑑4\displaystyle=\frac{1}{2}kx_{d}^{2}+\frac{1}{4}m\alpha x_{d}^{4}= divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_k italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + divide start_ARG 1 end_ARG start_ARG 4 end_ARG italic_m italic_α italic_x start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT

C Error correction of ‘111’ state

Choosing system parameters within the error correction regime, the time dynamics of the system exposed to a random impulse is illustrated in Fig. 5b). Each oscillator is initially prepared in the ‘1’ state then oscillator 3 is subjected to an impulse δp𝛿𝑝\delta pitalic_δ italic_p. We see that the impulse causes the amplitude of oscillator 3 to temporarily reach the ‘0’ state, however, it does not latch to this state. After an equilibration period, all oscillators settle back into the ‘1’ state. This represents a successful error correction event.

Refer to caption
Figure 5: Time dynamics of error correction device. The coupled oscillators are initialised in their ‘1’ states and after 4 periods of oscillation, an impulse is applied to the third oscillator. The third oscillator temporarily transitions into its ‘0’ state, but quickly equilibrates back. The parameters of the oscillator are given by: m=1012𝑚superscript1012m=10^{-12}italic_m = 10 start_POSTSUPERSCRIPT - 12 end_POSTSUPERSCRIPT kg, Γ=105Γsuperscript105\Gamma=10^{5}roman_Γ = 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT, ω0=106subscript𝜔0superscript106\omega_{0}=10^{6}italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT , α=2×1022𝛼2superscript1022\alpha=2\times 10^{22}italic_α = 2 × 10 start_POSTSUPERSCRIPT 22 end_POSTSUPERSCRIPT m.2s2{}^{-2}.s^{-2}start_FLOATSUPERSCRIPT - 2 end_FLOATSUPERSCRIPT . italic_s start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT, ω=1.152×106𝜔1.152superscript106\omega=1.152\times 10^{6}italic_ω = 1.152 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT, F=1.23×106𝐹1.23superscript106F=1.23\times 10^{-6}italic_F = 1.23 × 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT N, Δp=6×1012Δ𝑝6superscript1012\Delta p=6\times 10^{-12}roman_Δ italic_p = 6 × 10 start_POSTSUPERSCRIPT - 12 end_POSTSUPERSCRIPT kg.m.s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT, β=2×1011𝛽2superscript1011\beta=2\times 10^{11}italic_β = 2 × 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT s22{}^{-2}start_FLOATSUPERSCRIPT - 2 end_FLOATSUPERSCRIPT.

D Simulation of error correction map

FIG. 2 c) was produced using the following two algorithms: First, we determined the boundary between the three main regions (‘1’ bias, ‘0’ bias and retention). To achieve this, the following process was performed:

  1. 1.

    For a given set of parameters (ω0subscript𝜔0\omega_{0}italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, α𝛼\alphaitalic_α, ΓΓ\Gammaroman_Γ and β𝛽\betaitalic_β), choose a particular ω𝜔\omegaitalic_ω and F𝐹Fitalic_F.

  2. 2.

    Initialise the oscillators in the ‘0’ states (low displacement) or ‘1’ states (high displacement).

  3. 3.

    Let the oscillators come to equilibrium and record their equilibrium displacement.

  4. 4.

    If the oscillators equilibrate to a steady state that matches the initial conditions (i.e. ’111’ or ’000’), then the system is in the retention region. Otherwise, if the oscillators equilibrate to their ’1’ (‘0’) state regardless of their initial condition, the system is in the ‘1’ (‘0’) bias region.

  5. 5.

    Repeat steps 1 - 4 for many different choices of ω𝜔\omegaitalic_ω and F𝐹Fitalic_F, to produce a base phase map.

The above steps allowed us to divide the phase space into three regions; ‘0’ bias, ‘1’ bias and retention. The second algorithm was then used to locate the error correction region:

  1. 1.

    For a given set of parameters (ω0subscript𝜔0\omega_{0}italic_ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, α𝛼\alphaitalic_α, ΓΓ\Gammaroman_Γ and β𝛽\betaitalic_β), choose a particular ω𝜔\omegaitalic_ω and F𝐹Fitalic_F.

  2. 2.

    Initialise and equilibrate all the oscillators in the ‘0’ states (low displacement) or in the ‘1’ states (high displacement).

  3. 3.

    At a chosen time, introduce an instantaneous kick by changing the momentum of one of the oscillators. The new momentum, p1subscript𝑝1p_{1}italic_p start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, is given by p0+Δpsubscript𝑝0Δ𝑝p_{0}+\Delta pitalic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + roman_Δ italic_p, where p0subscript𝑝0p_{0}italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the original momentum and ΔpΔ𝑝\Delta proman_Δ italic_p is the momentum change from the introduced kick.

  4. 4.

    Let the oscillators come to equilibrium and observe whether the final states of the oscillators match the initial states.

  5. 5.

    Repeat steps 2 - 4 for many different phases of oscillation. If the system corrects errors at all tested error times, designate the set of ω𝜔\omegaitalic_ω and F𝐹Fitalic_F as error correcting.

  6. 6.

    Repeat steps 1 - 5 for many different choices of ω𝜔\omegaitalic_ω and F𝐹Fitalic_F, to produce an error correction parameter space map.

This produced an error correction map, showing the region where the system can correct for individual single event upsets (SEUs) introduced at any time. Note that within a cycle of oscillation, the momentum of an oscillator varies. As a result, an oscillator is expected to behave differently when a kick of fixed size (Δp)Δ𝑝(\Delta p)( roman_Δ italic_p ) is introduced at different phases in the oscillation. An error correction system must be able to correct errors introduced at any phase. Therefore, we have designed the algorithm to repeat the error correction tests by introducing kicks at different times, corresponding to different phases. Here, we tested 15 linearly spaced phases in our algorithm, with error correction only designated as achieved if the error was corrected at all phases.

Combining the results from the above two algorithms, the phase map of three coupled Duffing oscillators was determined, as shown in FIG. 2 c).

E Simulation (FIG. 4)

We developed a simulation that is able to repeatedly test whether a system fails to error correct with randomly imposed kicks. Here, we assume that there is an underlying probability of an oscillator experiencing a kick, Pkicksubscript𝑃𝑘𝑖𝑐𝑘P_{kick}italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT. We have assumed a kick probability of 0.030.030.030.03 per time interval. Note that given these values, the chance of two or more kicks occurring simultaneously on a single oscillator is negligible. As a result, we have chosen to implement a maximum of one kick per oscillator area in our simulation.

The algorithm used to test the performance of a given system (FIG. 4) is shown below:

  1. 1.

    Choose the parameters of the single and coupled oscillators, and the total number of tests. The coupled oscillators should be in the error correction regime. For fair comparison, the single oscillator should be in its retention region.

  2. 2.

    Initialise and equilibrate both systems in their ‘0’ states, record their equilibrium amplitudes. Similarly, initialise both systems in their ‘1’ states and record their equilibrium amplitudes.

  3. 3.

    Use random number generators to decide the number of oscillators experiencing kicks and impose them onto the system at a randomly generated time. For a single oscillator, one random number generator is used, whereas three random number generators are used for three coupled oscillators. For each generated number, r[0,1]𝑟01r\in[0,1]italic_r ∈ [ 0 , 1 ], a kick is imposed whenever r<Pkick𝑟subscript𝑃𝑘𝑖𝑐𝑘r<P_{kick}italic_r < italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT.

  4. 4.

    Let the system come to equilibrium and record the equilibrium amplitudes. Compare the final and the initial equilibrium amplitudes to determine whether the system has failed. Note that it is considered a failure if either the ‘0’ or ‘1’ states have been altered by the occurrence of kicks.

  5. 5.

    Repeat steps 3 - 4 for the total number of tests. Record the total number of failures.

F Theoretical distribution of system failure (FIG. 4)

Theoretically, we can predict the cumulative distribution function (CDF) of system failure for both systems. At a given time window, the CDF is given by the converse of the system succeeding up until that time window. This can be written as:

CDF(t~)𝐶𝐷𝐹~𝑡\displaystyle CDF(\tilde{t})italic_C italic_D italic_F ( over~ start_ARG italic_t end_ARG ) =1(1Pfail)t~absent1superscript1subscript𝑃𝑓𝑎𝑖𝑙~𝑡\displaystyle=1-(1-P_{fail})^{\tilde{t}}= 1 - ( 1 - italic_P start_POSTSUBSCRIPT italic_f italic_a italic_i italic_l end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT over~ start_ARG italic_t end_ARG end_POSTSUPERSCRIPT (13)

where Pfailsubscript𝑃𝑓𝑎𝑖𝑙P_{fail}italic_P start_POSTSUBSCRIPT italic_f italic_a italic_i italic_l end_POSTSUBSCRIPT is the probability of failure for a given system in a given time window. The subscripts s𝑠sitalic_s and c𝑐citalic_c will be used to denote single and coupled systems, respectively. Now, we will focus on determining Pfailsubscript𝑃𝑓𝑎𝑖𝑙P_{fail}italic_P start_POSTSUBSCRIPT italic_f italic_a italic_i italic_l end_POSTSUBSCRIPT for the two systems of interest.

Within our simulation, we assume that all oscillators are of the same area. In addition, we assume that the oscillator areas are sufficiently small such that the probability of an oscillator experiencing multiple kicks is negligible. That is, in a given period of time (time window), each oscillator must experience one of two outcomes; either kicked or not kicked. If we take the probability of a kick per oscillator area per time window as Pkicksubscript𝑃𝑘𝑖𝑐𝑘P_{kick}italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT, then the probability of failure for a single oscillator in any given time window is given by:

Pfails=Pkicksubscript𝑃𝑓𝑎𝑖subscript𝑙𝑠subscript𝑃𝑘𝑖𝑐𝑘\displaystyle P_{fail_{s}}=P_{kick}italic_P start_POSTSUBSCRIPT italic_f italic_a italic_i italic_l start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_POSTSUBSCRIPT = italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT (14)

This is because we set the kick amplitude such that all kick events result in a single oscillator failure.

In comparison, we expect three coupled oscillators to function as a three way majority voting system when operating in the error correcting regime. As a result, we anticipate this error correction system to fail only when multiple oscillators experience kicks during the same time window. Therefore, to calculate the probability of multiple oscillators experiencing kicks within the same time window, we can apply the Binomial theorem:

Pbinomial(j)subscript𝑃𝑏𝑖𝑛𝑜𝑚𝑖𝑎𝑙𝑗\displaystyle P_{binomial}(j)italic_P start_POSTSUBSCRIPT italic_b italic_i italic_n italic_o italic_m italic_i italic_a italic_l end_POSTSUBSCRIPT ( italic_j ) =(3j)(1Pkick)3jPkickjabsentbinomial3𝑗superscript1subscript𝑃𝑘𝑖𝑐𝑘3𝑗superscriptsubscript𝑃𝑘𝑖𝑐𝑘𝑗\displaystyle={3\choose j}(1-P_{kick})^{3-j}P_{kick}^{j}= ( binomial start_ARG 3 end_ARG start_ARG italic_j end_ARG ) ( 1 - italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 3 - italic_j end_POSTSUPERSCRIPT italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT (15)

where j𝑗jitalic_j is the number of oscillators experiencing kicks. Note that j𝑗jitalic_j is an integer in the range from 0 to 3. Here, the Binomial theorem is used to account for the different combinations of oscillator outcomes (kicked or not kicked). Using equation (15), the predicted probability of failure for three coupled oscillators is given by:

Pfailcsubscript𝑃𝑓𝑎𝑖subscript𝑙𝑐\displaystyle P_{fail_{c}}italic_P start_POSTSUBSCRIPT italic_f italic_a italic_i italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_POSTSUBSCRIPT =j=23(3j)(1Pkick)3j(Pkick)jabsentsuperscriptsubscript𝑗23binomial3𝑗superscript1subscript𝑃𝑘𝑖𝑐𝑘3𝑗superscriptsubscript𝑃𝑘𝑖𝑐𝑘𝑗\displaystyle=\sum_{j=2}^{3}{3\choose j}(1-P_{kick})^{3-j}(P_{kick})^{j}= ∑ start_POSTSUBSCRIPT italic_j = 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ( binomial start_ARG 3 end_ARG start_ARG italic_j end_ARG ) ( 1 - italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 3 - italic_j end_POSTSUPERSCRIPT ( italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT
=3(1Pkick)(Pkick)2+(Pkick)3absent31subscript𝑃𝑘𝑖𝑐𝑘superscriptsubscript𝑃𝑘𝑖𝑐𝑘2superscriptsubscript𝑃𝑘𝑖𝑐𝑘3\displaystyle=3(1-P_{kick})(P_{kick})^{2}+(P_{kick})^{3}= 3 ( 1 - italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT ) ( italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ( italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT (16)

which is the sum of the probability of the three coupled oscillators experiencing two or three kicks within the same time window.

Using equations (14) and (F), we may calculate the values of Pfailsubscript𝑃𝑓𝑎𝑖𝑙P_{fail}italic_P start_POSTSUBSCRIPT italic_f italic_a italic_i italic_l end_POSTSUBSCRIPT for a specific Pkicksubscript𝑃𝑘𝑖𝑐𝑘P_{kick}italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT. For instance, when Pkick=0.03subscript𝑃𝑘𝑖𝑐𝑘0.03P_{kick}=0.03italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT = 0.03, we calculated:

Pfailssubscript𝑃𝑓𝑎𝑖subscript𝑙𝑠\displaystyle P_{fail_{s}}italic_P start_POSTSUBSCRIPT italic_f italic_a italic_i italic_l start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_POSTSUBSCRIPT =0.03absent0.03\displaystyle=0.03= 0.03
Pfailcsubscript𝑃𝑓𝑎𝑖subscript𝑙𝑐\displaystyle P_{fail_{c}}italic_P start_POSTSUBSCRIPT italic_f italic_a italic_i italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_POSTSUBSCRIPT =0.0026absent0.0026\displaystyle=0.0026= 0.0026

Finally, by substituting these numbers into equation (13), we obtained the theoretical CDFs for both single and coupled oscillators.

Similarly, the theoretical prediction of probability of failure per event, P~failsubscript~𝑃𝑓𝑎𝑖𝑙\tilde{P}_{fail}over~ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_f italic_a italic_i italic_l end_POSTSUBSCRIPT, as shown in FIG. 4b) can be obtained using equations (14) and (F). Note that the probability of failure presented in FIG. 4b) is normalised to a per event basis. This was achieved by dividing equations (14) and (F) by the probability of an event occurring in each scenario:

P~failssubscript~𝑃𝑓𝑎𝑖subscript𝑙𝑠\displaystyle\tilde{P}_{fail_{s}}over~ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_f italic_a italic_i italic_l start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_POSTSUBSCRIPT =PkickPkickabsentsubscript𝑃𝑘𝑖𝑐𝑘subscript𝑃𝑘𝑖𝑐𝑘\displaystyle=\frac{P_{kick}}{P_{kick}}= divide start_ARG italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT end_ARG start_ARG italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT end_ARG
=1absent1\displaystyle=1= 1
P~failcsubscript~𝑃𝑓𝑎𝑖subscript𝑙𝑐\displaystyle\tilde{P}_{fail_{c}}over~ start_ARG italic_P end_ARG start_POSTSUBSCRIPT italic_f italic_a italic_i italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_POSTSUBSCRIPT =3(1Pkick)(Pkick)2+(Pkick)33(1Pkick)2(Pkick)+3(1Pkick)(Pkick)2+(Pkick)3absent31subscript𝑃𝑘𝑖𝑐𝑘superscriptsubscript𝑃𝑘𝑖𝑐𝑘2superscriptsubscript𝑃𝑘𝑖𝑐𝑘33superscript1subscript𝑃𝑘𝑖𝑐𝑘2subscript𝑃𝑘𝑖𝑐𝑘31subscript𝑃𝑘𝑖𝑐𝑘superscriptsubscript𝑃𝑘𝑖𝑐𝑘2superscriptsubscript𝑃𝑘𝑖𝑐𝑘3\displaystyle=\frac{3(1-P_{kick})(P_{kick})^{2}+(P_{kick})^{3}}{3(1-P_{kick})^% {2}(P_{kick})+3(1-P_{kick})(P_{kick})^{2}+(P_{kick})^{3}}= divide start_ARG 3 ( 1 - italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT ) ( italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ( italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG start_ARG 3 ( 1 - italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT ) + 3 ( 1 - italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT ) ( italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ( italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG

When using the above equations, we predict a single oscillator to always fail whenever an event occurs, and the coupled systems to only fail whenever two or more kicks are imposed.

G Difference between theoretical and numerical results (FIG. 4)

From FIG. 4, it can be observed that the theoretical predictions of the system failure rate are not consistent with the simulated data. Therefore, our previous prediction of Pfailsubscript𝑃𝑓𝑎𝑖𝑙P_{fail}italic_P start_POSTSUBSCRIPT italic_f italic_a italic_i italic_l end_POSTSUBSCRIPT must not account for all kicks that the systems can correct. To rectify this, we assume that the single system is capable of correcting xssubscript𝑥𝑠x_{s}italic_x start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT proportion of single kicks. Similarly, we assume that the coupled system not only corrects for all single kicks, but also corrects xcsubscript𝑥𝑐x_{c}italic_x start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT proportion of double kicks. With these assumptions, equations (14) and (F) become:

Pfailssubscript𝑃𝑓𝑎𝑖subscript𝑙𝑠\displaystyle P_{fail_{s}}italic_P start_POSTSUBSCRIPT italic_f italic_a italic_i italic_l start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_POSTSUBSCRIPT =(1xs)Pkickabsent1subscript𝑥𝑠subscript𝑃𝑘𝑖𝑐𝑘\displaystyle=(1-x_{s})P_{kick}= ( 1 - italic_x start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT (17)
Pfailcsubscript𝑃𝑓𝑎𝑖subscript𝑙𝑐\displaystyle P_{fail_{c}}italic_P start_POSTSUBSCRIPT italic_f italic_a italic_i italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_POSTSUBSCRIPT =(1xc)(1Pkick)Pkick2+Pkick3absent1subscript𝑥𝑐1subscript𝑃𝑘𝑖𝑐𝑘superscriptsubscript𝑃𝑘𝑖𝑐𝑘2superscriptsubscript𝑃𝑘𝑖𝑐𝑘3\displaystyle=(1-x_{c})\cdot(1-P_{kick})P_{kick}^{2}+P_{kick}^{3}= ( 1 - italic_x start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) ⋅ ( 1 - italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT ) italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_P start_POSTSUBSCRIPT italic_k italic_i italic_c italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT (18)

Using the above expressions, we have performed a non-linear fit to find the values of xssubscript𝑥𝑠x_{s}italic_x start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and xcsubscript𝑥𝑐x_{c}italic_x start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, that give the best agreement with the simulated data. The fitting was done using the lsqnonlin𝑙𝑠𝑞𝑛𝑜𝑛𝑙𝑖𝑛lsqnonlinitalic_l italic_s italic_q italic_n italic_o italic_n italic_l italic_i italic_n MATLAB function. This function performs a non-linear least squares fit of the simulation data over a range of kick probabilities to equations (17) - (18). The results of the non-linear fit (presented in FIG. 4)) are given by:

xssubscript𝑥𝑠\displaystyle x_{s}italic_x start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT =(8.8±0.4)%absentpercentplus-or-minus8.80.4\displaystyle=(8.8\pm 0.4)\%= ( 8.8 ± 0.4 ) %
xcsubscript𝑥𝑐\displaystyle x_{c}italic_x start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT =(64.0±0.5)%absentpercentplus-or-minus64.00.5\displaystyle=(64.0\pm 0.5)\%= ( 64.0 ± 0.5 ) %

This suggests that the single system can correct single kicks approximately 9%percent99\%9 % of the time, and the coupled system can correct approximately 64%percent6464\%64 % of double kicks. With these modified assumptions, the theoretical predictions are observed to agree with the simulated data, as shown in FIG. 4.

References

  • Wawrzynski [2021] W. Wawrzynski, Scientific reports 11, 2889 (2021).
  • Brennan et al. [2008] M. Brennan, I. Kovacic, A. Carrella, and T. Waters, Journal of sound and vibration 318, 1250 (2008).