Fact-checked by Grok 2 weeks ago

Marginal stability

Marginal stability refers to a critical in dynamical where a exhibits neither asymptotic (where trajectories converge to an ) nor (where trajectories diverge), but instead produces bounded responses that persist indefinitely, such as sustained oscillations or constant outputs. In the context of continuous-time linear time-invariant (LTI) described by \dot{x}(t) = Ax(t), marginal stability occurs when all eigenvalues of the matrix A lie in the closed left half of the , with any eigenvalues on the imaginary axis (j\omega-axis) having blocks of size one (i.e., they are simple and non-repeated). This ensures that the 's natural response neither decays to zero nor grows unbounded over time, distinguishing it from asymptotically (all eigenvalues in the open left half-plane) and unstable (any eigenvalue with positive real part). In , marginal stability is often analyzed through the poles of the system's , where non-repeated poles on the imaginary axis lead to step responses that oscillate indefinitely without or , such as in an undamped with G(s) = \frac{4}{s^2 + 4}. Repeated poles on the imaginary axis, however, result in due to in the response. Beyond linear systems, the concept extends to nonlinear dynamical systems, where marginal stability describes equilibria or manifolds at the onset of , organizing complex behaviors like or in . In fields such as and , marginal stability characterizes critical states in jammed packings or protein structures, where systems balance rigidity and flexibility at the edge of . This notion is fundamental for understanding phenomena near stability thresholds, influencing applications from control design to ecological modeling.

Core Concepts

Stability Classifications

In dynamical systems, an equilibrium point is defined as a state where the system's trajectory remains constant over time, satisfying \dot{x} = 0 for continuous-time systems described by \dot{x} = f(x). These points represent steady-state conditions, such as rest positions in systems or balanced states in models, serving as critical references for analyzing system behavior. Stability classifications categorize the behavior of trajectories near an point. Asymptotic stability occurs when all trajectories starting sufficiently close to the converge to it as time approaches , ensuring the returns to this state after small perturbations. In contrast, describes situations where trajectories diverge from the , amplifying even minor disturbances and leading to unbounded growth or departure from the point. These definitions form the foundational in , with marginal stability representing the boundary case where trajectories neither converge nor diverge but remain bounded. The origins of modern stability theory trace back to the work of , who in 1892 defended his doctoral thesis The General Problem of the Stability of Motion, introducing methods to assess stability without solving differential equations explicitly. A key prerequisite for these analyses involves linear approximations of nonlinear systems around equilibrium points, obtained via Taylor expansion to first order, which simplifies the study of local behavior by approximating the dynamics with a . This technique, \dot{x} \approx A(x - x_e) where A = \frac{\partial f}{\partial x}|_{x_e}, enables the application of linear stability criteria while capturing essential qualitative features of the original nonlinear dynamics.

Marginal Stability Criteria

Marginal stability refers to a condition in dynamical systems where the system is neither asymptotically stable nor unstable, meaning that perturbations to the system's state remain bounded over time but do not decay to zero./13%3A_Internal_(Lyapunov)_Stability/13.02%3A_Stability_of_Linear_Systems) In such systems, initial conditions or bounded inputs produce responses that neither grow indefinitely nor converge asymptotically to the equilibrium, often manifesting as sustained oscillations. This boundary state is critical in control theory, as it delineates the edge between stable and unstable behavior without the exponential decay characteristic of asymptotic stability./02%3A_Transfer_Function_Models/2.03%3A_System_Stability) For linear time-invariant systems, marginal stability is precisely defined by the eigenvalues of the system matrix: all eigenvalues must have non-positive real parts, with at least one eigenvalue having a real part exactly equal to zero (either purely imaginary or zero)./13%3A_Internal_(Lyapunov)Stability/13.02%3A_Stability_of_Linear_Systems) This condition ensures that no mode grows exponentially (no positive real parts) while allowing persistent oscillatory or constant components from the zero real-part eigenvalues. Additionally, for continuous-time systems, the eigenvalues with zero real parts must be simple (non-repeated) to avoid unbounded responses from blocks./13%3A_Internal(Lyapunov)_Stability/13.02%3A_Stability_of_Linear_Systems) The Routh-Hurwitz criterion provides a practical algebraic test for marginal stability without computing eigenvalues directly; an entire row of zeros in the Routh array indicates the presence of roots on the imaginary axis, signaling marginal stability after resolving the auxiliary polynomial formed from the previous row. This special case requires replacing the zero row with the derivative of the auxiliary polynomial to continue the array and confirm no right-half-plane roots. Similarly, the Nyquist stability criterion identifies marginal stability when the Nyquist plot of the open-loop transfer function passes through the critical point -1 on the real axis, corresponding to zero gain margin and poles on the jω-axis without net encirclements. A example of marginal stability is the governed by the \ddot{x} + \omega^2 x = 0, where \omega > 0 is the natural frequency./02%3A_Transfer_Function_Models/2.03%3A_System_Stability) The s^2 + \omega^2 = 0 yields eigenvalues \pm i\omega, both with zero real parts, resulting in bounded sinusoidal responses that neither decay nor amplify. This undamped oscillatory behavior exemplifies how marginal stability sustains energy without dissipation or growth./13%3A_Internal_(Lyapunov)_Stability/13.02%3A_Stability_of_Linear_Systems)

Linear Deterministic Systems

Continuous-Time Systems

In continuous-time linear systems, marginal stability is analyzed using the \dot{x}(t) = Ax(t), where x(t) \in \mathbb{R}^n is the and A \in \mathbb{R}^{n \times n} is the . The system is marginally stable if all eigenvalues \lambda_i of A satisfy \operatorname{Re}(\lambda_i) \leq 0, with at least one eigenvalue having \operatorname{Re}(\lambda_i) = 0, and all such eigenvalues having associated Jordan blocks of size one (i.e., simple eigenvalues with geometric multiplicity equal to algebraic multiplicity). This condition ensures that the state trajectories remain bounded for all time but do not necessarily converge to the . The eigenvalues \lambda_i are the roots of the characteristic equation \det(sI - A) = 0, where s is a variable. with negative real parts contribute decaying terms to the response, while those on the imaginary (purely imaginary or zero) produce sustained oscillations or constant terms without growth or decay. Thus, marginal stability is indicated by all roots lying in the closed left half of the , with at least one on the imaginary . The general solution to the state equation is x(t) = e^{At} x(0), where e^{At} is the . This solution remains bounded for all t \geq 0 and initial conditions x(0) if no eigenvalue has a positive real part, but trajectories do not approach zero when imaginary-axis eigenvalues are present. The canonical form of A plays a critical role in determining the precise nature of the response. For marginal stability, any eigenvalue with \operatorname{Re}(\lambda_i) = 0 must have all associated Jordan blocks of size one (i.e., simple eigenvalues with geometric multiplicity equal to algebraic multiplicity). zero eigenvalues yield bounded, non-converging solutions such as constants or sinusoids, whereas repeated zero eigenvalues with larger Jordan blocks introduce growth terms like t^k e^{\lambda_i t}, leading to instability. Physically, marginal stability manifests as sustained oscillations without , as seen in the linearized model of an undamped . The equation \ddot{\theta} + \frac{g}{l} \theta = 0 (for small angles \theta) has state-space form with A = \begin{pmatrix} 0 & 1 \\ -\frac{g}{l} & 0 \end{pmatrix}, yielding eigenvalues \lambda = \pm i \sqrt{g/l} on the imaginary . This results in periodic motion of constant amplitude, bounded but neither growing nor decaying.

Discrete-Time Systems

In discrete-time linear systems, the state evolution is governed by the state-space equation x_{k+1} = A x_k, where x_k is the at time step k and A is the system matrix. Marginal stability occurs when all eigenvalues \lambda_i of A satisfy |\lambda_i| \leq 1, with at least one eigenvalue having |\lambda_i| = 1, and all such eigenvalues having associated blocks of size one (i.e., simple eigenvalues with algebraic multiplicity equal to geometric multiplicity), ensuring that the state remains bounded for any finite but does not necessarily converge to the . This condition parallels the continuous-time case where eigenvalues lie on the imaginary axis, but in discrete time, the relevant boundary is the unit circle in the . The eigenvalues are the roots of the \det(zI - A) = 0, and marginal stability is indicated when all roots lie inside or on the unit circle, with at least one on the boundary. In contrast to Schur stability, which requires all eigenvalues strictly inside the unit circle (|\lambda_i| < 1) for asymptotic stability and exponential decay of the state, unit-modulus eigenvalues in the marginal case lead to persistent oscillations without growth or decay, such as complex conjugate pairs z = e^{\pm i \theta} producing undamped sinusoidal responses. For marginal stability to hold, any eigenvalue on the unit circle must be simple (algebraic multiplicity equals geometric multiplicity, with no associated Jordan blocks larger than 1×1), as higher multiplicity can introduce polynomial growth terms, rendering the system unstable. A representative example is a digital filter with a transfer function whose poles lie on the unit circle, such as a simple resonator H(z) = \frac{1}{1 - 2\cos\theta \, z^{-1} + z^{-2}}, which has poles at e^{\pm i \theta}. For a unit impulse input, the output is an undamped sinusoid at frequency \theta / (2\pi) with constant amplitude, remaining bounded indefinitely without attenuation. The general solution for the unforced system is x_k = A^k x_0, where the powers A^k can be decomposed via the Jordan canonical form. For marginal stability, when all eigenvalues satisfy |\lambda_i| \leq 1 and unit-circle eigenvalues are simple, \|x_k\| remains bounded for all k, as the modal components associated with |\lambda_i| = 1 evolve as constant-amplitude sinusoids or constants, while interior eigenvalues contribute decaying terms.

System Responses

Transient Dynamics

In marginally stable linear systems, the transient response to initial conditions or external impulses features non-decaying oscillations or persistent constant offsets, stemming from eigenvalues on the imaginary axis. These eigenvalues produce undamped sinusoidal components in the time-domain solution, which neither grow nor decay over time, contrasting with asymptotically stable systems where transients eventually vanish. For instance, the general solution to the state equation \dot{x} = Ax includes terms like \sin(\omega t) or \cos(\omega t) for purely imaginary eigenvalues \pm j\omega, ensuring bounded but perpetual motion. The impulse response of a continuous-time marginally stable system, given by h(t) = \mathcal{L}^{-1}\{G(s)\}, exhibits sustained sinusoids when the transfer function G(s) has poles on the j\omega-axis. This response remains bounded for simple poles but does not converge to zero, as the inverse transform yields terms such as \sin(\omega t) or \cos(\omega t) for t \geq 0. In contrast, poles in the open left half-plane would damp these oscillations exponentially. The pole multiplicity plays a pivotal role: simple poles (multiplicity one) confine the response to bounded sinusoids, while higher multiplicity introduces growing polynomial factors, such as t \sin(\omega t), which cause the output to unboundedly increase and tip the system into instability. A classic example is the undamped second-order system, modeled by \ddot{x} + \omega^2 x = 0, with response x(t) = A \cos(\omega t + \phi) to initial conditions x(0) and \dot{x}(0), where A and \phi are determined accordingly. This perpetual oscillation at the natural frequency \omega illustrates the bounded yet non-convergent transient behavior inherent to marginal stability. Such systems are particularly sensitive to perturbations; minor parameter variations, like a slight increase in damping or stiffness, can shift poles off the imaginary axis—either stabilizing the response with decay or destabilizing it with exponential growth.

Frequency Domain Characteristics

In frequency domain analysis, marginal stability is characterized by specific features in graphical representations of system transfer functions, indicating that the system neither grows unbounded nor decays to zero but sustains persistent oscillations. These characteristics are crucial for assessing stability margins in linear systems, where the boundary between stability and instability is probed through frequency responses. Bode plots reveal marginal stability when the phase margin is 0° at the gain crossover frequency (where the magnitude is 0 dB) or the gain margin is 0 dB at the phase crossover frequency (where the phase is -180°). A phase margin of 0° signifies that any additional phase lag would drive the system into instability, resulting in undamped oscillations at the crossover frequency. Similarly, a gain margin of 0 dB indicates that the system is on the verge of instability, as an infinitesimal increase in gain would cause the Nyquist contour to encircle the critical point. These margins quantify the proximity to the stability boundary, with zero values confirming neutral stability. The Nyquist plot provides a complementary view, where marginal stability occurs if the open-loop frequency response contour passes through the critical point -1 + j0 without encircling it. This passage implies that the closed-loop poles lie on the imaginary axis, leading to sustained sinusoidal responses rather than exponential divergence or convergence. The plot's intersection with -1 at a frequency ω_c corresponds to a magnitude of unity and phase of -180°, mirroring the but visualizing the encirclement condition directly. For the underlying transfer function G(s), marginal stability manifests as poles located precisely on the imaginary axis, such as at s = ±jω_0 for a pair of complex conjugate poles. These poles yield an infinite steady-state gain |G(jω_0)| at the resonant frequency ω_0, as the denominator approaches zero while the numerator remains finite, resulting in unbounded amplitude for inputs at that frequency. This infinite gain underscores the system's inability to attenuate oscillations at resonance, distinguishing marginal stability from asymptotic stability (poles in the left half-plane) or instability (poles in the right half-plane). Root locus analysis further illustrates marginal stability when branches of the locus touch or cross the imaginary axis at a critical gain value K_c. At this gain, the closed-loop poles migrate to s = ±jω_c, marking the onset of neutral stability; for K > K_c, poles enter the right half-plane, causing . This crossing point can be found by solving the 1 + KG(s) = 0 for imaginary , providing insight into gain limits for oscillatory . In engineering applications, such as power systems, marginal stability arises when the torque coefficient approaches zero, leading to undamped rotor oscillations between synchronous machines. This condition, often analyzed via frequency-domain small-signal models, highlights the verge of oscillatory , where disturbances result in persistent inter-area modes without .

Stochastic Extensions

Random Perturbations

In stochastic systems, stability under random perturbations is analyzed through linear of the form \dot{x} = A x + \sigma dW_t, where x \in \mathbb{R}^n is the , A is the drift matrix, \sigma is the diffusion matrix, and W_t is a standard . In linear SDEs with additive , eigenvalues of A with zero real parts generally lead to mean-square instability due to diffusive growth in neutral directions, unlike the deterministic case. Marginal stability in such systems requires specific conditions, such as noise orthogonal to the neutral eigenspace, to prevent unbounded variance. In the asymptotic stability regime, where all eigenvalues of A have strictly negative real parts, mean-square stability ensures that the second moment \mathbb{E}[\|x(t)\|^2] remains bounded as t \to \infty and converges exponentially to a positive value determined by the intensity and system parameters. This boundedness arises from the dissipative , where perturbations decay exponentially on average, but the input maintains a non-zero variance without driving the system to a trivial . A representative example is the Ornstein-Uhlenbeck (OU) process modeling velocity in , given by dV = -\gamma V \, dt + \sigma \, dW_t with friction coefficient \gamma > 0. The stationary variance is \mathbb{E}[V^2] = \sigma^2 / (2\gamma), which remains finite and bounded, illustrating sustained fluctuations from . As \gamma \to 0^+, the system approaches the marginal regime, with stationary variance diverging to infinity, but for any \gamma > 0, it is mean-square stable. In the limit \gamma = 0, the process reduces to pure dV = \sigma \, dW_t, where variance \mathbb{E}[V(t)^2] = \sigma^2 t grows linearly, indicating instability. The power spectral density (PSD) provides insight into the frequency content of these fluctuations in stable stochastic systems. For the OU process with small but positive \gamma, the PSD takes the Lorentzian form S(\omega) = \frac{2 \sigma^2 \gamma}{\gamma^2 + \omega^2}, featuring a flat response at low frequencies and resonant peaks near \omega = 0 that broaden as \gamma \to 0, indicating persistent low-frequency . In more general stable systems with near-neutral modes (eigenvalues close to the imaginary axis), the PSD exhibits resonant peaks near the natural frequencies, reflecting accumulation of from at those modes with slow . Unlike deterministic marginal stability, where trajectories remain bounded with persistent oscillations, random perturbations in linear SDEs typically prevent true marginality by causing gradual growth in the second moment when neutral modes are present, as the noise continuously excites those modes and precludes strict boundedness without sufficient damping.

Lyapunov Spectrum Analysis

In the context of stochastic extensions to marginal stability, the Lyapunov spectrum provides a detailed characterization of the average exponential rates of divergence or convergence of nearby trajectories in nonlinear or high-dimensional random dynamical systems, where the largest exponent being zero signifies marginal stability on chaotic attractors. This spectrum consists of a set of real numbers, ordered from largest to smallest, that quantify the local expansion and contraction rates along different directions in the tangent space, with the sum of the exponents relating to the system's phase space volume preservation or dissipation. For marginally stable stochastic systems, the maximum Lyapunov exponent λ₁ = 0 indicates a neutral balance between stretching and folding, preventing both asymptotic stability and explosive divergence, often observed at the boundary between ordered and chaotic behaviors in random environments. The theoretical foundation for computing the Lyapunov spectrum in random dynamical systems is provided by Oseledets' multiplicative ergodic , which guarantees the existence of these exponents under assumptions for almost every and realization. This , applicable to both deterministic and settings, decomposes the into Oseledets subspaces associated with distinct exponents, enabling the spectrum's determination through limits of logarithmic growth rates of cocycle norms. In random dynamical systems driven by , the extends to time-dependent cocycles, ensuring the exponents are invariant under the system's ergodic measure and well-defined for infinite-dimensional cases like those in . A representative example of marginal stability via the occurs in stochastically perturbed coupled oscillators, where emerges when the largest transverse is zero, marking the between desynchronized and coherent phase-locking under . In such systems, weak scales the exponents linearly with strength, and λ₁ = 0 corresponds to the critical point where perturbations neither amplify nor decay on average, facilitating applications in neural networks or climate models with intermittent forcing. Computation of the Lyapunov spectrum typically involves simulations that average over multiple noise realizations, evolving an of tangent vectors via the system's linearized dynamics and Gram-Schmidt orthogonalization to estimate the exponents from rates. This method, initialized with random tangent vectors, iteratively renormalizes the basis after each time step to track expansion rates, with convergence achieved over long integration times and ensemble averages to account for variability. In , the Lyapunov spectrum has been applied to models at the onset of , where the exponents reveal the from laminar to turbulent states, with λ₁ ≈ 0 indicating marginal stability just before the of structures in shear flows. For instance, in homogeneous isotropic , direct numerical simulations show the full spectrum flattening near critical Reynolds numbers, quantifying the dimensional scaling of modes and aiding predictions of energy cascades.

References

  1. [1]
    [PDF] Module 06 Stability of Dynamical Systems - UTSA
    Oct 10, 2017 · Definition of Marginal Stability. The CT LTI system ˙x(t) = Ax(t) is called marginally stable if both of these statements are true:.
  2. [2]
    [PDF] 15 TRANSFER FUNCTIONS & STABILITY
    response never decays or grows in amplitude; this is called marginal stability. If at least one real exponent is positive, then one element of the response ...
  3. [3]
    None
    ### Summary of Marginal Stability from Section 3: Stability
  4. [4]
    [PDF] On the dynamics of turbulent transport near marginal stability
    The marginal stability hypothesis is simply the notion that when a local gradient exceeds the critical value set by a stability criterion, the fluctuation- ...
  5. [5]
    [PDF] Marginal stability in jammed packings: Quasicontacts and weak ...
    Aug 14, 2014 · This marginal stability has traditionally been understood in terms of isostaticity, the equality of the number of mechanical contacts and the ...
  6. [6]
    Functionality and the evolution of marginal stability in proteins - NIH
    If this trade-off holds for all protein sequences, marginal stability is prevalent because it provides the best balance between the sequences that result in ...
  7. [7]
    [PDF] Chapter Four - Graduate Degree in Control + Dynamical Systems
    Equilibrium points are one of the most important features of a dynamical sys- tem since they define the states corresponding to constant operating conditions. A.
  8. [8]
    [PDF] Chapter 7 - Dynamical Systems
    If p is an equilibrium point of the system (7.1) then the set consisting of p alone is an invariant set for that system. Likewise, if γ(p) is a periodic orbit,.
  9. [9]
    Dynamic Behavior - FBSwiki
    Nov 23, 2012 · An equilibrium point is said to be globally asymptotically stable if all initial conditions converge to that equilibrium point. Global stability ...
  10. [10]
    [PDF] Stability of Dynamical systems - math.utah.edu
    Unstable equilibrium. The equilibrium point u0 is called unstable provided it is not stable, which means (a) or (b) fails (or both).Missing: instability | Show results with:instability
  11. [11]
    [PDF] PDF - Lectures on Dynamic Systems and Control
    It is possible to have stability in the sense of Lyapunov without having asymptotic stability, in which case we refer to the equilibrium point as marginally.
  12. [12]
    A. M. Lyapunov's stability theory—100 years on - Oxford Academic
    On 12 October 1892 (by the modern calendar) Alexandr Mikhailovich Lyapunov defended his doctoral thesis The general problem of the stability of motion at Moscow ...Missing: origins | Show results with:origins
  13. [13]
    [PDF] Linear ODEs and Stability - Weill Cornell Medicine
    However, the analysis of sets of linear ODEs is very useful when considering the stability of non-linear systems at equilibrium. For that reason, we will pursue ...
  14. [14]
    [PDF] Chapter 2: Taylor Approximations
    One of the most powerful tools for understanding a nonlinear system is to linearize that system around some point of interest (such as a stationary point).
  15. [15]
    Chapter 5 Stability Analysis | Optimal Control and Estimation
    If the linearized system (5.6) is marginally stable (i.e., all eigenvalues of A A have nonpositive real parts, and at least one eigenvalue has zero real ...
  16. [16]
    [PDF] Explaining the Routh-Hurwitz criterion
    Sep 15, 2019 · Singular case #1: the leading coefficient is zero, but the row is not identically zero. Polynomial division could proceed, but would produce ...
  17. [17]
    [PDF] Nyquist Stability Criterion
    Two important notions can be derived from the Nyquist diagram: phase and gain stability margins. The phase and gain stability margins are presented in Figure 4 ...
  18. [18]
    [PDF] Pendulum and Stability in Two Dimensional Systems
    Sep 20, 2022 · The archetypal oscillatory dynamical system is the pendulum. We begin with the unforced, undamped case. l mg mgsinθ θ. The ...Missing: continuous | Show results with:continuous
  19. [19]
    [PDF] 7.8 Stability of Discrete-Time Linear Systems
    eigenvalues is asymptotically stable if all its eigenvalues are located in the unit. circle of the complex plane; the system is marginally stable if all ...
  20. [20]
    [PDF] System Stability - The Open Repository @ Binghamton (The ORB)
    The symbols. (+) indicate the equilibria. (What would the phase portrait of an undamped pendulum look like?) The graph plots trajectories on theqθ θ.
  21. [21]
    Stability Revisited | Introduction to Digital Filters - DSPRelated.com
    Isolated poles on the unit circle may be called marginally stable. The impulse response component corresponding to a single pole on the unit circle never ...
  22. [22]
    Stability Revisited - Stanford CCRMA
    Isolated poles on the unit circle may be called marginally stable. The impulse response component corresponding to a single pole on the unit circle never decays ...
  23. [23]
    [PDF] asymptotically stable marginally stable unstable
    Relate the transient response of a system to the poles of its transfer ... 3.4 Marginal Stability . . . . . . . . . . . . . . . . . . . . . . . . . 12.
  24. [24]
    7.1 Second Order Underdamped Systems
    A second order underdamped system has complex conjugate roots, an oscillatory response, and a fast response time, with less damping.
  25. [25]
    Control System Stability - Roy Mech
    The transient response, and as a result the stability, of such a system depends on the coefficients a 0, a 1 , a 2. Assuming a 0 >0 then provided that a 1 >0 ...
  26. [26]
    [PDF] Applied Stochastic Differential Equations
    May 3, 2019 · The main motivation for the book is the application of stochastic differential equations (SDEs) in domains such as target tracking and medical ...
  27. [27]
    On the Theory of the Brownian Motion | Phys. Rev.
    With a method first indicated by Ornstein the mean values of all the powers of the velocity u and the displacement s of a free particle in Brownian motion ...
  28. [28]
    (PDF) Marginal Instability and Intermittency in Stochastic Systems
    Aug 10, 2025 · ... power spectral density (PSD) of g(t) so that. λ. is a mean frequency of g(t). Thus, according to the equation (1) random process g(t) is ...
  29. [29]
    Stability analysis of chaotic systems from data - PMC - NIH
    If at least a Lyapunov exponent is positive, λ 1 > 0 , the attractor is chaotic. ... In particular, the Oseledets' theorem enables the extension of Lyapunov ...Missing: marginal | Show results with:marginal<|control11|><|separator|>
  30. [30]
    [PDF] Chapter 6 - Lyapunov exponents - ChaosBook.org
    In the limit of infinite time the Lyapunov exponent is a global measure of the rate at which nearby trajectories diverge, averaged over the strange attractor.Missing: Oseledets | Show results with:Oseledets
  31. [31]
  32. [32]
    [PDF] Notes on the Multiplicative Ergodic Theorem - UChicago Math
    (iii) The Oseledets theorem holds for both real and complex vector bundles (with hermitian metrics). The Lyapunov exponents are real numbers in both cases. Page ...
  33. [33]
    [PDF] Lyapunov Exponents for Random Dynamical Systems
    Jan 15, 2009 · Thanks to the multiplicative ergodic theorem of Oseledets [109], the Lyapunov spec- ... the theory of random dynamical systems (see Arnold [3]).
  34. [34]
    Generalized synchrony of coupled stochastic processes with ...
    Nov 18, 2016 · We study the effect of multiplicative noise in dynamical flows arising from the coupling of stochastic processes with intrinsic noise.
  35. [35]
    [PDF] Scaling and Synchronization in Deterministic and Stochastic ...
    First, a continuous-time stochastic model for weakly coupled chaotic systems is introduced to study the scaling of the Lyapunov exponents with the coupling ...
  36. [36]
    [PDF] A practical method for calculating largest Lyapunov exponents from ...
    Nov 20, 1992 · Lyapunov exponents quantify the exponential divergence of initially close state-space trajectories and estimate the amount of chaos in a system.
  37. [37]
    [PDF] arXiv:1802.10482v1 [cond-mat.stat-mech] 28 Feb 2018
    Feb 28, 2018 · We revisit the numerical calculation of generalized Lyapunov exponents, L(q), in deterministic dynamical systems. The standard method consists ...Missing: simulation | Show results with:simulation
  38. [38]
    [PDF] Entropy production and Lyapunov instability at the onset of turbulent ...
    The horizontal sweeping near the opposite thermal boundary causes fluid flow in a direction counter to the net flow of ... Lyapunov spectrum, and it is also not ...
  39. [39]
    Lyapunov spectrum of forced homogeneous isotropic turbulent flows
    Nov 25, 2019 · In this work, the Lyapunov spectrum of forced homogeneous isotropic turbulent flows is computed. Using the Lyapunov exponents of a flow at different Reynolds ...Missing: onset | Show results with:onset