Fact-checked by Grok 2 weeks ago

Dynamical system

A dynamical system is a that describes how the state of a system evolves over time, where the state is represented by variables in a , and the evolution follows deterministic rules such as ordinary differential equations for continuous time or iterative maps for discrete time. This framework captures a wide range of behaviors, including growth, decay, oscillation, evolution, collapse, and , by focusing on both quantitative trajectories and qualitative long-term . Dynamical systems are broadly classified into continuous and discrete types, with continuous systems modeled by flows generated by differential equations that describe smooth changes, and discrete systems by iterations of functions that update states in successive steps. Key concepts in the theory include (the set of all possible states), orbits (the paths traced by states over time), fixed points or equilibria (states that do not change), and (whether nearby states converge to or diverge from equilibria). Additional notions such as attractors (sets toward which orbits converge), bifurcations (qualitative changes in behavior as parameters vary), and chaotic dynamics (sensitive dependence on initial conditions) are central to understanding complex evolutions. The field traces its origins to the late , particularly Henri Poincaré's foundational work on the in , which introduced qualitative methods for analyzing nonlinear systems and highlighted the limitations of . This was expanded in the by George Birkhoff's and Stephen Smale's topological approaches, establishing dynamical systems as an independent mathematical discipline with deep ties to analysis, , and probability. The theory's maturation into a core area of reflects its role in addressing global orbit structures and invariant sets, often without requiring explicit solutions. Applications of dynamical systems are vast and interdisciplinary, modeling physical phenomena like planetary motion and fluid in physics, and spread in , control strategies in , and even neural network behaviors in . In complex settings, such as interdependent nonlinear systems, emergent patterns arise from interactions far from equilibrium, influencing fields from climate modeling to . These models not only predict behaviors but also inform decision-making by revealing stability thresholds and chaotic regimes.

Introduction

Core Concepts

A dynamical system formalizes the evolution of a physical, biological, or abstract process over time through a deterministic rule that maps states to subsequent states. At its core, it consists of a state space X, which is the set of all possible configurations or conditions of the system, and an evolution rule \phi, which describes how the system transitions from one state to another, either continuously via a flow or discretely via a map. In the continuous case, \phi is often a flow \Phi: \mathbb{R} \times X \to X that parametrizes the progression along time, while in the discrete case, it is an iterated map T: X \to X applied successively. This pair (X, \phi) captures the essence of predictability in deterministic processes, where the system's behavior is fully specified by its current state without external randomness. Dynamical systems are inherently deterministic, meaning that given the evolution rule and an initial state, the entire future (and often past) path is uniquely determined, contrasting with systems that incorporate probabilistic elements. The initial-value problem plays a central role here: it specifies a starting point x_0 \in X at some initial time, allowing the evolution rule to generate the system's history from that point onward. This setup ensures that solutions to the dynamical system are well-defined and unique under suitable conditions on \phi, such as or , enabling the study of long-term behavior without ambiguity. The path traced by the system from an initial state under the evolution rule is known as a , while the set of all points visited along this path forms the . In discrete systems, the orbit is the sequence \{ \phi^n(x_0) \mid n = 0, 1, 2, \dots \}, where \phi^n denotes n-fold ; in continuous systems, it is the image of the curve \{ \phi(t, x_0) \mid t \in \mathbb{R} \}. Orbits provide the fundamental building blocks for analyzing qualitative features like or periodicity, as multiple trajectories may converge or diverge based on their starting points in the state space. In the phase space, which visualizes the state space X (often with coordinates representing variables like and ), trajectories appear as distinct curves or lines illustrating the system's ; for instance, closed loops indicate periodic orbits, while diverging paths suggest . Such representations highlight how nearby initial conditions can lead to similar or wildly different orbits, a key intuition for understanding . Dynamical systems find applications in modeling physical processes like and biological ones like .

Scope and Applications

Dynamical systems theory encompasses a wide range of applications in the natural and social sciences, providing mathematical frameworks to model and predict the evolution of complex processes over time. In , it is used to describe the motion of celestial bodies, such as planetary orbits under gravitational forces. In , dynamical systems model , capturing interactions between and environmental factors to forecast ecological changes. Similarly, in , these models analyze market fluctuations and resource allocation, simulating how economic variables like evolve. In , particularly systems, dynamical systems theory designs mechanisms to stabilize processes in mechanical and electrical systems. A key strength of dynamical systems lies in its focus on qualitative behaviors, such as —where systems return to after perturbations—and periodicity, where trajectories exhibit repeating cycles. These properties allow researchers to assess long-term outcomes without solving equations explicitly, revealing patterns like attractors in state space that govern system evolution. The theory emerged as a vital bridge between and applied sciences in the mid-20th century, integrating differential equations with interdisciplinary problems to address nonlinear phenomena across fields. Computational simulations, including methods, further enable the approximation of trajectories for systems too complex for analytical solutions.

Historical Context

Early Developments

The study of dynamical systems traces its origins to the foundational work in during the . Galileo Galilei (1564–1642) made pivotal contributions through his experiments on inclined planes and , demonstrating that objects accelerate uniformly under and formulating the principle of , which posits that bodies maintain their state of motion unless acted upon by external forces. These ideas, detailed in his Dialogue Concerning the Two Chief World Systems (1632) and Discourses and Mathematical Demonstrations Relating to Two New Sciences (1638), shifted the analysis of motion from qualitative descriptions to quantitative laws, serving as precursors to systematic studies of evolving physical systems. Galileo's emphasis on empirical observation and mathematical modeling laid the groundwork for deterministic approaches to motion. Building directly on Galileo's insights, synthesized these concepts in his seminal Philosophiæ Naturalis Principia Mathematica (1687), where he articulated the three laws of motion and the law of universal gravitation. Newton's framework established as a , in which the future state of any mechanical configuration evolves predictably from initial conditions and governing forces, enabling precise predictions such as planetary orbits. This Principia marked a cornerstone for dynamical systems by introducing differential equations to describe continuous evolution, influencing subsequent and highlighting the predictability inherent in classical laws. By the late 19th century, the limitations of purely analytical solutions in complex systems prompted qualitative investigations. (1854–1912), in response to a prize competition posed by in 1889, examined the in during the 1880s and 1890s, revealing that small perturbations could lead to intricate, non-periodic behaviors in planetary motions. In his memoir published in Acta Mathematica (1890), Poincaré pioneered qualitative methods, such as recurrence theorems and the analysis of invariant manifolds, to study the global structure of solutions without solving equations explicitly. These approaches, applied to the restricted , demonstrated the inherent instability in certain configurations and foreshadowed the complexities of nonlinear dynamics. A parallel advancement came from Alexander Lyapunov (1857–1918), whose 1892 doctoral thesis The General Problem of the Stability of Motion provided the first comprehensive theory for assessing stability in dynamical systems governed by differential equations. Lyapunov defined as the property where solutions remain bounded near an equilibrium under small perturbations and introduced asymptotic stability, where solutions converge to the equilibrium over time. His direct method, using auxiliary functions to bound solution behavior, offered practical tools for mechanical and celestial applications, bridging physical intuition with rigorous analysis.

20th-Century Foundations

In the early , advanced the abstract study of dynamical systems through his foundational work on topological dynamics, introduced in his 1927 monograph Dynamical Systems, where he analyzed the qualitative behavior of on compact phase spaces using topological methods. This framework emphasized the invariance of topological properties under homeomorphisms, providing a rigorous basis for understanding long-term dynamics without explicit solutions. Complementing this, Birkhoff's 1931 pointwise ergodic theorem established that, for an ergodic measure-preserving transformation on a , the time average of an integrable function along almost every converges to its space average, linking to abstract dynamics. Building on these ideas, Aleksandr Andronov and introduced the concept of in the 1930s, formalized in their 1937 paper on "rough systems" (systèmes grossiers), which defined systems whose qualitative phase portraits remain unchanged under small perturbations of the defining equations. This notion, applied initially to , prioritized robustness in qualitative theory, ensuring that generic dynamical behaviors persist despite imperfections in models derived from physical observations. Their work shifted focus from exact solvability to the topological equivalence of nearby systems, influencing later classifications of stable flows. The mid-20th century saw further maturation of qualitative theory through contributions from and . In the 1950s, Kolmogorov's seminar in fostered developments in ergodic and , including his 1954 on the persistence of quasi-periodic motions in nearly integrable systems under small perturbations, which quantified stability in multi-dimensional phase spaces. By the 1960s, Smale extended these ideas globally, proving in his 1960 paper on Morse inequalities that certain gradient-like systems satisfy topological constraints on fixed points and periodic orbits, and in his seminal 1967 Bulletin article, he unified local and global analyses for differentiable flows on manifolds, establishing as a dense property in the space of smooth vector fields. A pivotal event signaling the field's maturity was the 1963 International Symposium on Nonlinear Differential Equations and Nonlinear Mechanics, organized by Joseph P. LaSalle and , which gathered leading researchers to discuss qualitative methods, , and applications, resulting in proceedings that synthesized progress in abstract formulations.

Definitions and Formulations

Geometric Approach

In the geometric approach, a continuous-time dynamical system, often called a real dynamical system, is defined on a smooth manifold X as a one-parameter group of diffeomorphisms, or flow, denoted \phi_t: \mathbb{R} \times X \to X, where t \in \mathbb{R} parameterizes time. This flow satisfies the group properties: the identity axiom \phi_0(x) = x for all x \in X, and the composition axiom \phi_{t+s}(x) = \phi_t(\phi_s(x)) for all t, s \in \mathbb{R} and x \in X, ensuring invertibility via \phi_{-t} = \phi_t^{-1}. These axioms capture the deterministic evolution of states in the phase space X, where trajectories \{ \phi_t(x) \mid t \in \mathbb{R} \} trace the system's paths under smooth transformations preserving the manifold's topology. The \phi_t is generated by a V: X \to [TX](/page/TX), where [TX](/page/TX) is the of X. Specifically, for each x \in X, the curve t \mapsto \phi_t(x) satisfies the autonomous \frac{d}{dt} \phi_t(x) = V(\phi_t(x)), \quad \phi_0(x) = x. To derive this, consider the composition property: fix x and differentiate \phi_{t+h}(x) = \phi_t(\phi_h(x)) with respect to h at h=0, yielding \frac{d}{dt} \phi_t(x) = \frac{d}{dh} \big|_{h=0} \phi_h(\phi_t(x)), which by the 's equals V(\phi_t(x)). This links the geometric structure to local differential behavior, enabling qualitative analysis of orbits without explicit solutions. In the discrete-time case, a dynamical system is defined by a smooth f: X \to X on the manifold X, with evolution given by : the forward of x \in X is \{ f^n(x) \mid n = 0, 1, 2, \dots \}, where f^0(x) = x and f^{n+1}(x) = f(f^n(x)), extendable backward if f is invertible to \{ f^n(x) \mid n \in \mathbb{Z} \}. This formulation models stroboscopic sections of continuous flows or inherently discrete processes, focusing on and periodic points within the . For multidimensional maps, such as f: \mathbb{R}^n \to \mathbb{R}^n with n > 1, the geometric approach extends to analyzing non-compact phase spaces by compactification, embedding \mathbb{R}^n into a compact manifold like the or, for polynomial vector fields, the . This adds points at to study asymptotic behavior and qualitative features, such as attractors or homoclinic tangles at large norms, transforming unbounded orbits into bounded ones on the compactified space. For instance, in the Poincaré compactification of \mathbb{R}^2, the plane is projected onto the hemisphere S^2, revealing dynamics "at infinity" through invariant circles. Such techniques facilitate global bifurcation analysis without measure-theoretic tools.

Measure-Theoretic Approach

In the measure-theoretic approach, a dynamical system is formalized as a (X, \mu, T), where X is a equipped with a \sigma- \mathcal{B}, \mu is a on (X, \mathcal{B}), and T: X \to X is a that preserves the measure \mu. This framework emphasizes statistical properties and long-term averages over the , extending beyond purely topological considerations by incorporating probabilistic structure. The T acts as either a or the time-1 map of a continuous on the measure space, satisfying the invariance condition \mu(T^{-1}A) = \mu(A) for every measurable set A \in \mathcal{B}. This preservation ensures that the measure remains unchanged under the , allowing the analysis of invariant sets and the distribution of orbits with respect to \mu. In relation to the geometric viewpoint, the (X, \mu) provides a for volumes in the , enabling the quantification of typical behaviors and the study of recurrence in finite-measure settings. A prominent example of an invariant measure is the on the torus \mathbb{T}^d, which is preserved by rotations, such as irrational rotations on the circle or multidimensional toral automorphisms. For an irrational rotation R_\alpha: \mathbb{T} \to \mathbb{T} given by x \mapsto x + \alpha \pmod{1} with \alpha irrational, the \lambda satisfies \lambda(R_\alpha^{-1} A) = \lambda(A) for Borel sets A, reflecting the uniform distribution of orbits. Similarly, the doubling map on the circle preserves despite stretching intervals, as the folding compensates for expansion. These examples illustrate how invariant measures capture the "natural" volume in systems arising from geometric or physical origins. A foundational result in this approach is the , which asserts that if (X, \mu, T) has finite total measure \mu(X) < \infty and T is measure-preserving, then for any measurable set A \subset X with \mu(A) > 0, almost every point x \in A (with respect to \mu) returns to A under iterates of T, i.e., there exist infinitely many n > 0 such that T^n x \in A. This theorem, originally established by Henri Poincaré in 1890, highlights the recurrent nature of finite-measure dynamics and underpins the study of invariant sets and ergodicity, demonstrating that orbits cannot escape bounded regions indefinitely.

Building Dynamical Systems

From Differential Equations

Continuous dynamical systems in \mathbb{R}^n are constructed from autonomous ordinary differential equations (ODEs) of the form \dot{x} = f(x), where x \in \mathbb{R}^n and f: \mathbb{R}^n \to \mathbb{R}^n is a vector field, typically assumed to be continuously differentiable. The solution to the initial value problem (IVP) \dot{x}(t) = f(x(t)), x(0) = x_0, defines a curve in phase space parameterized by time t, known as a trajectory or orbit. The flow \phi_t: \mathbb{R}^n \to \mathbb{R}^n of the dynamical system is the map that evolves initial points via \phi_t(x_0) = x(t), satisfying the initial condition \phi_0(x) = x for all x \in \mathbb{R}^n. This flow captures the time evolution and possesses the semigroup property \phi_{s+t}(x) = \phi_s(\phi_t(x)) for s, t \geq 0, extending to a group under appropriate completeness assumptions. The existence and uniqueness of solutions to the IVP, essential for defining a consistent , follow from the . This asserts that if f is continuous and locally in x (i.e., there exists a constant L > 0 such that \|f(x) - f(y)\| \leq L \|x - y\| for x, y in some neighborhood), then for every x_0 \in \mathbb{R}^n, there exists a unique solution x(t) defined on a maximal time interval (-\alpha, \beta) with \alpha, \beta > 0. The Lipschitz condition prevents solution crossing and ensures the is continuously differentiable in its arguments. Without local , solutions may fail to be unique, as in cases like \dot{x} = |x|^{1/2}. The of the system is the geometric representation in \mathbb{R}^n formed by the family of all curves (trajectories) of the ODE, excluding time parameterization to focus on qualitative structure. These curves foliate the , revealing invariant sets, directions of motion via the f, and asymptotic behaviors without solving the equations explicitly. For n=2, phase portraits are often sketched with arrows indicating flow direction and special orbits highlighted. A representative example is the undamped harmonic oscillator, governed by the linear system \begin{cases} \dot{x} = y, \\ \dot{y} = -x, \end{cases} where x denotes position and y velocity in \mathbb{R}^2. The phase portrait comprises a continuum of closed elliptical orbits centered at the origin (0,0), the sole equilibrium point, each corresponding to periodic motion with conserved energy \frac{1}{2}(x^2 + y^2) = \text{constant}. The explicit solution for initial condition (x(0), y(0)) = (x_0, y_0) is x(t) = x_0 \cos t + y_0 \sin t, \quad y(t) = -x_0 \sin t + y_0 \cos t, yielding orbits that are ellipses scaled by the initial amplitude, with period $2\pi independent of amplitude. This illustrates conservative dynamics where trajectories neither converge nor diverge but cycle indefinitely./03%3A_Linear_Oscillators/3.04%3A_Geometrical_Representations_of_Dynamical_Motion)

From Discrete Maps

Discrete dynamical systems arise from recurrence relations of the form x_{n+1} = f(x_n), where f: X \to X is a defined on a space X, typically a of \mathbb{R}^d or a more general , and n indexes discrete time steps. The f generates the through its iterates f^n, where f^n(x) denotes the composition of f with itself n times, evolving the initial x_0 to subsequent states. This formulation contrasts with continuous-time systems by advancing the state in fixed, jumps rather than smooth evolution. Such discrete maps often originate from continuous dynamical systems. A primary source is the time-one map of a flow, which samples the continuous trajectory at integer time intervals; for a flow \phi_t(x) generated by a differential equation, the map f(x) = \phi_1(x) yields a discrete system whose orbits approximate the original flow's behavior over unit times. Alternatively, discrete maps can result from sampling or discretizing continuous systems, such as using numerical integration schemes like the Euler method on ordinary differential equations to produce iterative updates. Cellular automata provide another construction of discrete dynamical systems, particularly on structures. These consist of a of cells, each in a finite , updated synchronously via a local rule that depends on neighboring cells; the global evolution forms a map on the configuration space of the . For instance, , a one-dimensional with binary states, evolves according to a simple neighborhood rule and exhibits complex, Turing-complete behavior despite its local simplicity. The of an initial point x under the f is \{x, f(x), f^2(x), \dots \}, which traces the system's forward in time. Fixed points, where the state remains unchanged, satisfy f(x) = x, representing equilibria in the dynamics. These elements—orbits and fixed points—form the foundational structures for analyzing long-term behavior in systems.

Key Examples

Continuous Systems

Continuous dynamical systems are typically modeled by ordinary differential equations (ODEs) that describe smooth flows in , evolving continuously over time. These systems arise naturally in physics, such as in and , where trajectories represent the paths of states without discrete jumps. Key examples illustrate nonlinear behaviors like periodic motions, separatrices, and chaotic attractors, providing foundational insights into qualitative dynamics. The simple pendulum serves as a classic example of a continuous dynamical system, governed by the nonlinear second-order \ddot{\theta} + \sin \theta = 0, where \theta denotes the from the downward vertical, and the equation assumes normalized units with over length g/l = 1. In the of coordinates (\theta, \dot{\theta}), the traces closed orbits for small oscillations (librations) around the stable equilibrium at \theta = 0, while larger energies lead to rotational orbits encircling the . Due to the periodicity of \theta $2\pi, the forms a \mathbb{R} \times S^1, with the separatrix—a homoclinic figure-eight curve passing through the saddle points at (\pm \pi, 0)—dividing librational and rotational trajectories. Another prominent example is the , introduced by Balthasar van der Pol in 1926 as a model for self-sustained oscillations originally motivated by circuits, described by the second-order \ddot{x} - \mu (1 - x^2) \dot{x} + x = 0, where \mu > 0 controls the nonlinearity strength. In (x, \dot{x}), trajectories spiral toward a unique stable , which is nearly circular for small \mu and relaxes sharply for large \mu, demonstrating how nonlinear damping drives sustained periodic motion independent of initial conditions. The Lorenz system, derived as a simplified model of atmospheric convection and introduced by Edward Lorenz in 1963, exemplifies chaotic behavior in continuous systems through the coupled first-order ODEs \begin{align*} \dot{x} &= \sigma (y - x), \ \dot{y} &= r x - y - x z, \ \dot{z} &= x y - b z, \end{align*} with typical parameters \sigma = 10, r = 28, b = 8/3. Its trajectories in three-dimensional phase space converge to a strange attractor resembling a butterfly, where solutions exhibit sensitive dependence on initial conditions, folding between two lobes without repeating. This qualitative structure highlights the emergence of deterministic chaos in continuous flows.

Discrete and Cellular Models

Discrete dynamical systems evolve through iterative applications of a map, where the state at each discrete time step is determined by a function applied to the previous state. These models are particularly useful for studying phenomena that occur in distinct steps, such as population growth in generations or digital simulations. A prominent example is the logistic map, a one-dimensional quadratic recurrence relation originally derived from the continuous logistic differential equation but analyzed in its discrete form to reveal complex behaviors. The logistic map is defined by the iteration x_{n+1} = r x_n (1 - x_n), where x_n represents the state (often normalized population) at step n, and r is a parameter controlling growth rate, typically in the interval [0, 4]. For low values of r (e.g., $0 < r < 3), the system converges to a stable fixed point, representing equilibrium. As r increases beyond 3, the dynamics undergo period-doubling bifurcations, leading to cycles of period 2, 4, 8, and higher, until chaos emerges for r > 3.57 approximately, where the behavior becomes highly sensitive to initial conditions. This parameter-dependent transition illustrates how simple nonlinear iterations can produce intricate patterns, making the logistic map a cornerstone for understanding discrete chaos. In higher dimensions, the , introduced by Michel Hénon in 1976, provides a two-dimensional that extends these ideas to coupled variables, often used to model physical processes like atmospheric . It is given by \begin{align*} x_{n+1} &= 1 - a x_n^2 + y_n, \\ y_{n+1} &= b x_n, \end{align*} with typical parameters a = 1.4 and b = 0.3, which generate a strange —a structure where orbits are dense but non-periodic. The map's quadratic nonlinearity in the x-coordinate drives the and folding of trajectories, producing a bounded yet unpredictable confined to an area-filling curve in the . This , numerically observed through long iterations, highlights the geometric complexity possible in iterations beyond one . Cellular automata represent another class of discrete models, where the system consists of a of cells evolving synchronously according to local rules based on neighboring states. John Horton Conway's , introduced in , is a seminal two-dimensional example on an infinite grid, with s in binary states (alive or dead). The evolution follows four rules applied simultaneously: a live cell survives if it has two or three live neighbors; a dead cell becomes alive (birth) with exactly three live neighbors; otherwise, cells die (underpopulation or overpopulation). These simple neighborhood-based updates yield emergent complexity, including stable patterns (still lifes), oscillators (periodic orbits), and gliders (moving structures), demonstrating and universality in discrete spatial dynamics. In one-dimensional maps like the , the long-term behavior is characterized by —the of iterates \{x_n\}—which may converge to fixed points, enter periodic cycles, or wander ergodically. Periodic occur when the repeats after p steps, i.e., x_{n+p} = x_n for all n sufficiently large, with the smallest such p as the . The coexistence of follows a specific ordering: if a continuous map on the interval admits a periodic orbit of m, it also admits orbits of all n where m precedes n in the Sharkovsky ordering (3 ≻ 5 ≻ 7 ≻ ... ≻ 2·3 ≻ 2·5 ≻ ... ≻ 4 ≻ 2 ≻ 1). This theorem encapsulates the hierarchical structure of periodicities in 1D discrete systems, revealing that period-3 implies all other periods, underscoring the richness of even simple iterations.

Linear Systems

Flows in Continuous Time

In continuous-time linear dynamical systems, the evolution of the state vector \mathbf{x}(t) \in \mathbb{R}^n is governed by the ordinary differential equation (ODE) \dot{\mathbf{x}} = A \mathbf{x}, where A is an n \times n constant matrix. This formulation describes the flow of the system, which maps initial conditions \mathbf{x}(0) to states at time t, preserving the linear structure. The unique solution to this initial value problem is given by \mathbf{x}(t) = e^{A t} \mathbf{x}(0), where e^{A t} denotes the matrix exponential, defined as the power series e^{A t} = \sum_{k=0}^{\infty} \frac{(A t)^k}{k!} or, equivalently, through the Laplace transform or other methods when the series is impractical. This exponential map generates the system's flow \phi(t, \mathbf{x}_0) = e^{A t} \mathbf{x}_0, which is a one-parameter group under time addition, ensuring reversibility for t < 0. For non-diagonalizable matrices, the matrix exponential can be computed using the Jordan canonical form, where the behavior along generalized eigenspaces determines the trajectories. To analyze the qualitative behavior, the eigenvalues \lambda of A play a central role. If A is diagonalizable, A = P D P^{-1} with D diagonal containing the eigenvalues, then e^{A t} = P e^{D t} P^{-1}, where e^{D t} has entries e^{\lambda_i t}. The system is asymptotically stable at the origin if all eigenvalues satisfy \operatorname{Re}(\lambda_i) < 0, leading to \lim_{t \to \infty} \mathbf{x}(t) = 0 for any initial condition. It is Lyapunov stable if all \operatorname{Re}(\lambda_i) \leq 0 and eigenvalues with \operatorname{Re}(\lambda_i) = 0 have associated Jordan blocks of size 1 (no larger blocks, which would cause polynomial growth); the system is unstable otherwise. A canonical example is the damped harmonic oscillator, modeled in state-space form as \dot{\mathbf{x}} = \begin{pmatrix} 0 & 1 \\ -\omega^2 & -2\zeta\omega \end{pmatrix} \mathbf{x}, where \mathbf{x} = (q, \dot{q})^T, \omega > 0 is the natural frequency, and \zeta \geq 0 is the ratio. The eigenvalues are \lambda = -\zeta\omega \pm \omega \sqrt{\zeta^2 - 1}, yielding overdamped (\zeta > 1, two real negative roots), critically damped (\zeta = 1, repeated real root), or underdamped (\zeta < 1, complex conjugates with negative real part) behavior, all asymptotically stable for \zeta > 0. Constant-coefficient systems more generally, such as coupled oscillators or electrical circuits, follow similar eigenvalue-driven dynamics. In two dimensions, phase portraits illustrate these flows vividly. For real eigenvalues of the same sign, trajectories approach or depart along straight lines (stable/unstable nodes); opposite signs produce hyperbolic paths crossing at the . Complex eigenvalues with negative real part generate spiraling inward orbits ( focus), while pure imaginary eigenvalues yield closed elliptical orbits (center, Lyapunov but not asymptotically ). These portraits, constructed from the eigenvectors as principal directions, provide geometric insight into the global flow without solving explicitly. For non-diagonalizable cases, such as defective matrices with Jordan blocks larger than 1x1 for eigenvalues with Re(λ)=0, the portraits include shear or deviation terms.

Iterated Maps in Discrete Time

In discrete-time dynamical systems, linear iterated maps describe the evolution of a state vector \mathbf{x}_n \in \mathbb{R}^d through the \mathbf{x}_{n+1} = A \mathbf{x}_n, where A is a constant d \times d . This formulation models systems where each step applies a fixed linear , common in applications like or economic modeling. The general solution is given by \mathbf{x}_n = A^n \mathbf{x}_0, where \mathbf{x}_0 is the initial state and A^n denotes the n-th power. For non-diagonalizable A, A^n is computed via the Jordan canonical form, revealing polynomial terms in n for larger blocks. To compute A^n explicitly, assume A is diagonalizable, so A = P D P^{-1} with D diagonal containing the eigenvalues \lambda_i of A and P the matrix of corresponding eigenvectors. Then, A^n = P D^n P^{-1}, where D^n = \operatorname{diag}(\lambda_1^n, \dots, \lambda_d^n), yielding \mathbf{x}_n = P D^n P^{-1} \mathbf{x}_0. In the scalar case (d=1), the map simplifies to x_{n+1} = \lambda x_n, with solution x_n = \lambda^n x_0, forming a geometric that converges to zero as n \to \infty if |\lambda| < 1. The asymptotic stability of the origin (equilibrium at \mathbf{x} = 0) requires that all trajectories converge to zero regardless of \mathbf{x}_0, which occurs if and only if every eigenvalue \lambda_i of A satisfies |\lambda_i| < 1. The origin is Lyapunov stable if all |\lambda_i| \leq 1 and eigenvalues with |\lambda_i| = 1 have associated Jordan blocks of size 1; otherwise, the system is unstable, for example, when larger blocks cause polynomial growth in n (e.g., for \lambda = 1 with block size >1). If any |\lambda_i| > 1, the system is unstable, with trajectories diverging exponentially. A classic example is the , defined by f_{n+2} = f_{n+1} + f_n with initial conditions f_0 = 0, f_1 = 1, which arises from iterating the A = \begin{pmatrix} 1 & 1 \\ 1 & 0 \end{pmatrix}, so \begin{pmatrix} f_{n+1} \\ f_n \end{pmatrix} = A^n \begin{pmatrix} 1 \\ 0 \end{pmatrix}. The eigenvalues are the \phi \approx 1.618 > 1 and its conjugate \hat{\phi} \approx -0.618, implying instability as the sequence grows exponentially. Another example is the autoregressive process of order one (AR(1)), y_{n+1} = \phi y_n + \epsilon_{n+1} where \epsilon_n is ; the deterministic part follows the scalar map with eigenvalue \phi, and stationarity (corresponding to of the homogeneous system) holds if |\phi| < 1, ensuring bounded variance.

Local Behavior

Linearization Techniques

Linearization techniques approximate the behavior of nonlinear dynamical systems locally near equilibrium points by replacing them with their linear counterparts, facilitating analysis through familiar linear methods. In continuous-time dynamical systems defined by an ordinary differential equation \dot{x} = f(x), where x \in \mathbb{R}^n and x^* is an equilibrium point satisfying f(x^*) = 0, the linearization is obtained via the A = Df(x^*), the matrix of first partial derivatives of f evaluated at x^*. The resulting linear system is \dot{y} = A y, where y = x - x^* represents the deviation from the equilibrium; solutions to this system provide a first-order approximation of the nonlinear flow near x^*. For discrete-time systems given by an iterated map x_{n+1} = g(x_n), with fixed point x^* such that g(x^*) = x^*, the linearization uses the Jacobian matrix B = Dg(x^*). The linearized map becomes z_{n+1} = B z_n, where z_n = x_n - x^*, approximating the nonlinear iteration locally around x^*. The Hartman-Grobman theorem establishes the validity of this approximation under certain conditions: if x^* is a hyperbolic equilibrium—meaning no eigenvalues of A (for flows) have zero real part, or no eigenvalues of B (for maps) have modulus one—then there exists a homeomorphism in a neighborhood of x^* that conjugates the nonlinear system topologically to its linearization, preserving orbits and their directions. This result, originally proved for flows by Hartman and independently for maps by Grobman, ensures that the qualitative phase portrait near hyperbolic points mirrors that of the linear system. A classic example is the logistic map x_{n+1} = r x_n (1 - x_n) for $0 < r \leq 4 and x_n \in [0,1], which models discrete population growth. The fixed points are x^* = 0 and, for r > 1, x^* = (r-1)/r. at these points uses the f'(x) = r(1 - 2x): at x^* = 0, f'(0) = r, indicating if |r| < 1; at x^* = p = (r-1)/r, f'(p) = 2 - r, indicating if |2 - r| < 1 or $1 < r < 3.

Stability Near Equilibria

The local stability of an equilibrium point in a dynamical system is classified based on the behavior of trajectories near that point, typically analyzed through the linearization of the system at the equilibrium. For a hyperbolic equilibrium—where no eigenvalue of the linearized system has zero real part—the guarantees that the nonlinear flow is topologically conjugate to the linear flow in a neighborhood of the equilibrium, allowing stability to be determined from the eigenvalues of the . An equilibrium is asymptotically stable if all trajectories starting sufficiently close to it converge to it as time tends to infinity; in the linearized system, this occurs when all eigenvalues \lambda satisfy \operatorname{Re}(\lambda) < 0. Conversely, the equilibrium is unstable if trajectories diverge from it, corresponding to at least one eigenvalue with \operatorname{Re}(\lambda) > 0. A saddle equilibrium exhibits mixed behavior, with some directions attracting trajectories (stable subspace, \operatorname{Re}(\lambda) < 0) and others repelling them (unstable subspace, \operatorname{Re}(\lambda) > 0); this classification holds for the nonlinear system near hyperbolic saddles. The rates of approach or divergence near the are quantified by the , which for the linearized system coincide with the real parts of the eigenvalues and indicate the or decay along principal directions. In higher dimensions, the largest being negative confirms asymptotic locally, while a positive value signals instability. For non-hyperbolic equilibria where some eigenvalues have \operatorname{Re}(\lambda) = 0 (neutral or center directions), linearization alone is inconclusive, and the theorem provides a tool to reduce the dynamics to a lower-dimensional manifold to the center eigenspace at the . On this invariant center manifold, the is determined by the reduced system, often revealing whether the is stable, unstable, or requires further nonlinear analysis. A classic example is the saddle equilibrium in a two-dimensional , such as for the \dot{x} = x, \dot{y} = -y, where the origin has eigenvalues +1 and -1; trajectories approach along the y-axis () but diverge along the x-axis (unstable manifold), forming curves that separate regions of different qualitative .

Global Dynamics

Invariant Manifolds

In dynamical systems, an invariant manifold is a of the that is mapped into itself under the or defining the , meaning trajectories starting on the manifold remain on it for all time. For a continuous-time dynamical system generated by a \phi_t, the of a point x, denoted W^s(x), consists of all points y such that \phi_t(y) \to x as t \to \infty. Similarly, the unstable manifold W^u(x) comprises points y where \phi_t(y) \to x as t \to -\infty. These manifolds capture the asymptotic of orbits approaching or departing from x, providing essential geometric structures for understanding long-term . Hyperbolic sets form a foundational context for manifolds in nonlinear systems. A compact set \Lambda for a f is if the over \Lambda decomposes into stable and unstable subbundles E^s(z) and E^u(z) at each z \in \Lambda, with uniform along E^s (i.e., \|Df^n(v)\| \leq C \lambda^n \|v\| for v \in E^s, $0 < \lambda < 1) and uniform expansion along E^u (i.e., \|Df^{-n}(w)\| \leq C \lambda^n \|w\| for w \in E^u). This splitting induces local stable and unstable manifolds tangent to these subbundles, which foliate neighborhoods of \Lambda and dictate the local geometry of the dynamics. sets, such as those arising in Smale's horseshoe construction, exhibit structural stability, ensuring that nearby systems preserve the qualitative manifold structure. The stable manifold theorem guarantees the local existence of these structures near hyperbolic points. For a hyperbolic fixed point p of a C^1 diffeomorphism f on a Banach space, where the spectrum of Df(p) splits into parts inside and outside the unit circle, there exists a local stable manifold W^s_\mathrm{loc}(p), which is a C^1 submanifold tangent to the stable eigenspace E^s(p) and invariant under f, such that orbits from points in W^s_\mathrm{loc}(p) converge exponentially to p. An analogous result holds for unstable manifolds and extends to flows via suspension. In the nonuniform hyperbolic case, where expansion and contraction rates vary along orbits, local manifolds still exist but may shrink in size, maintaining injectivity and exponential convergence properties. Invariant manifolds play a critical qualitative role in homoclinic tangles, which emerge when the stable and unstable manifolds of a hyperbolic fixed point intersect transversely, forming a complex interlaced structure. These intersections create a tangle that organizes chaotic orbits into lobes, leading to exponential stretching, symbolic dynamics, and positive topological entropy without measure-theoretic chaos. In Hamiltonian systems, such tangles govern transport barriers and mixing, as seen in Poincaré's analysis of the three-body problem, where manifold intersections produce self-similar patterns of escape and recurrence.

Attractors and Basins

In dynamical systems, an attractor is a compact, invariant set A in the phase space such that there exists an open neighborhood U of A where all trajectories starting in U converge to A as time tends to infinity, meaning the distance from any point in U to A decreases asymptotically. The basin of attraction of such an attractor A, denoted B(A), is the open set consisting of all initial conditions whose forward orbits approach A in the long term. Attractors can take various forms depending on the system's structure. A fixed point attractor is a hyperbolic equilibrium point with all eigenvalues of the Jacobian having negative real parts, attracting nearby trajectories to itself; for example, the origin in the damped harmonic oscillator \dot{x} = -x, \dot{y} = -y. A limit cycle attractor is an isolated periodic orbit that draws in surrounding trajectories, as seen in the Van der Pol oscillator where oscillations stabilize to a unique cycle. A strange attractor is a fractal invariant set with non-integer dimension, exhibiting sensitive dependence on initial conditions while remaining bounded, exemplified by the in the system \dot{x} = \sigma(y - x), \dot{y} = rx - y - xz, \dot{z} = xy - bz with parameters \sigma=10, r=28, b=8/3. Morse-Smale systems represent a class of structurally stable dynamical systems featuring a finite number of hyperbolic fixed points and periodic orbits as attractors, with their stable and unstable invariant manifolds intersecting transversely and no homoclinic tangencies. In these systems, the phase space partitions into disjoint basins of attraction for each attractor, ensuring robust global dynamics without chaotic behavior. The transverse intersections of manifolds in Morse-Smale systems facilitate a complete description of the flow via a finite directed graph of connections between equilibria and orbits.

Bifurcations

Local Bifurcation Types

Local bifurcations occur in dynamical systems when small changes in a parameter lead to qualitative alterations in the local behavior near equilibria, such as the creation, annihilation, or stability exchange of fixed points or periodic orbits. These codimension-one phenomena are analyzed using normal forms derived from center manifold reduction and linear stability analysis, where the Jacobian matrix at the equilibrium has a single eigenvalue crossing the imaginary axis (real part changing sign for real eigenvalues or crossing from negative to positive real part for complex conjugates). The saddle-node bifurcation, also known as a fold bifurcation, involves the coalescence and annihilation of a stable and an unstable equilibrium as the parameter varies. In one-dimensional continuous-time systems, two fixed points approach each other and disappear when the parameter exceeds a critical value, leading to a loss of equilibrium and potentially directing trajectories toward other attractors. The generic normal form for this local bifurcation is given by \dot{x} = \mu + x^2, where \mu is the bifurcation parameter; for \mu < 0, there are two real equilibria at x = \pm \sqrt{-\mu}, with the negative one stable and the positive one unstable, while for \mu > 0, no real equilibria exist. This bifurcation is ubiquitous in models like , where it represents thresholds such as extinction points. Transcritical bifurcations feature an exchange of between two that exist on both sides of the critical value, without their creation or destruction. One , often at the , transitions from to unstable, while the other moves through it and gains . The normal form equation is \dot{x} = \mu x - x^2, with at x = 0 and x = \mu; for \mu < 0, the is and x = \mu unstable, reversing for \mu > 0. This type arises in systems with a or , such as logistic models with harvesting. Pitchfork bifurcations occur in systems with , where a single splits into three upon variation: the original loses , and two new symmetric equilibria emerge. There are two variants—supercritical, where branches appear for \mu > 0, and subcritical, where unstable branches exist for \mu < 0, potentially leading to hysteresis. The normal form is \dot{x} = \mu x - x^3, satisfying odd f(-x, \mu) = -f(x, \mu); at \mu = 0, the origin is the bifurcation point with zero linear term and negative cubic coefficient ensuring supercriticality in this case. Examples include buckling of symmetric structures under load. The Hopf bifurcation marks the birth or death of a small-amplitude limit cycle from an equilibrium as a pair of complex conjugate eigenvalues crosses the imaginary axis, changing the equilibrium's stability. In supercritical cases, a stable limit cycle emerges for \mu > 0, while subcritical yields an unstable cycle for \mu < 0. The normal form in polar coordinates is \dot{r} = \mu r - r^3, \quad \dot{\theta} = \omega + O(r^2), where the first Lyapunov coefficient determines the supercritical (< 0) or subcritical (> 0) nature, with \omega the angular frequency. This local phenomenon is fundamental in oscillatory systems like the .

Global Bifurcation Phenomena

Global bifurcations in dynamical systems involve qualitative changes that affect the overall of , often through the interaction of distant structures like manifolds or orbits, contrasting with local bifurcations that occur near isolated fixed points or periodic orbits. These phenomena can lead to the sudden appearance or disappearance of attractors, the birth of dynamics, or the reconfiguration of basins of attraction, and they are typically analyzed using tools from singularity theory and analysis. Unlike local bifurcations, which can be codimension-one and unfold predictably, ones often require higher and exhibit to initial conditions even before emerges. A prominent example is the , where a trajectory approaches a point as time tends to both plus and minus , forming a that connects the saddle to itself. In continuous-time systems, the creation of such a homoclinic orbit to a hyperbolic saddle can generate a complex structure of invariant manifolds, leading to the formation of horseshoe maps and the onset of strange attractors as the parameter varies. This is responsible for many chaotic regimes in low-dimensional systems, where the return map near the saddle exhibits and folding, amplifying small perturbations into global instability. The period-doubling cascade represents another key global , particularly in one-dimensional iterated maps, where a stable periodic loses through a sequence of period-doubling , with the period increasing exponentially as the parameter approaches a . demonstrated in 1978 that this cascade converges geometrically to , characterized by universal scaling ratios, such as the Feigenbaum constant δ ≈ 4.669, which governs the rate of accumulation and applies across a wide class of unimodal maps. This universality arises from the self-similar structure of the applied to the maps, making the cascade a generic route to in dissipative systems. The blue sky catastrophe describes a global bifurcation where a stable limit cycle disappears without colliding with another orbit or equilibrium, instead expanding indefinitely in amplitude while its period tends to infinity, vanishing "into the blue sky" of phase space. This occurs as a saddle-node bifurcation on a closed invariant curve, where the cycle merges with its unstable counterpart in a manner that alters the global flow topology. An explicit low-dimensional example is provided by a system of ordinary differential equations exhibiting this codimension-one phenomenon, highlighting its role in transitions from periodic to aperiodic behavior in applications like fluid dynamics. The Shilnikov scenario illustrates a specific homoclinic leading to spiral , involving a homoclinic attached to a saddle-focus where the unstable manifold spirals into the stable one. When the saddle value (ratio of eigenvalues) exceeds unity, this generates a of unstable periodic orbits organized in a spiral structure, forming a with Smale horseshoe-like dynamics and positive . Leonid Shilnikov's foundational work in the established that such loops in three-dimensional flows produce complex, non-wandering sets containing homoclinic tangles, providing a mechanism for robust in systems like the Lorenz .

Ergodic Properties

Ergodicity Basics

In measure-preserving dynamical systems, where the or preserves an μ, captures the idea that time averages along typical orbits coincide with spatial averages over the . Specifically, a system (X, μ, φ) is ergodic with respect to μ if every measurable set has μ-measure 0 or 1, which is equivalent to the condition that, for every integrable f : X → ℝ, the time average equals the average for μ-almost every x ∈ X: \lim_{T \to \infty} \frac{1}{T} \int_0^T f(\phi_t(x)) \, dt = \int_X f \, d\mu. This equivalence ensures that long-term behavior observed along individual trajectories reflects the global statistical properties encoded by μ. The foundational result establishing this time-space equivalence is the Birkhoff ergodic theorem, which asserts that in an ergodic measure-preserving system, the limit \lim_{T \to \infty} \frac{1}{T} \int_0^T f(\phi_t(x)) \, dt = \int_X f \, d\mu exists and holds for μ-almost every x, whenever f is integrable with respect to μ. This pointwise convergence theorem, proved by in 1931, underpins much of by linking dynamical orbits to measure-theoretic integrals. A classic example illustrating is the irrational rotation on (one-dimensional ) ℝ/ℤ, defined by φ(x) = (x + α) mod 1, where α is irrational. With respect to μ, this system is ergodic because almost every is dense and equidistributed on , ensuring time averages match spatial integrals. In contrast, a rational rotation, such as α = p/q in lowest terms, is not ergodic, as orbits are periodic and confined to finite sets of measure zero, failing the indecomposability condition. Furthermore, the of an probability measure μ guarantees with respect to μ, as any decomposition into multiple ergodic components would yield distinct measures, contradicting . This highlights how the absence of distributions forces the system to exhibit the full mixing of orbits required for .

Invariant Measures

In dynamical systems, measures are probability measures preserved under the of the system, providing a for understanding long-term statistical behavior. For dissipative systems exhibiting , Sinai-Ruelle-Bowen (SRB) measures play a central role as physically relevant measures. These measures are characterized by being absolutely continuous with respect to on the unstable manifolds of the system, while being singular on stable directions, and they capture the natural statistics of typical orbits in attractors. SRB measures were originally constructed for Axiom A diffeomorphisms, where they coincide with equilibrium states for the geometric potential, and their existence has been extended to broader classes of partially attractors in dissipative settings. A key property strengthening ergodicity—where time averages equal space averages —is mixing, which ensures that correlations between observables over time, indicating asymptotic independence of distant points in the . In a (X, \mu, T), strong mixing holds if for any measurable sets [A, B](/page/List_of_French_composers) \subset X, \mu(T^{-n}A \cap B) \to \mu(A)\mu(B) as n \to \infty, reflecting a uniform spreading of measure that goes beyond mere indecomposability. This of correlations facilitates the study of statistical properties, such as the for Birkhoff sums, in systems like flows on manifolds of negative . In nearly integrable Hamiltonian systems, the Kolmogorov-Arnold-Moser (KAM) theorem guarantees the persistence of invariant tori under small perturbations, each supporting an invariant measure that is the product of Lebesgue measures on the torus angles. These quasi-periodic invariant tori, with Diophantine frequency vectors, survive for perturbations small enough in C^\infty topology, preserving a positive measure set of the original and enabling the construction of KAM tori with associated ergodic measures. The theorem resolves the small divisor problem through iterative normal form transformations, ensuring stability in systems like the for small coupling. To quantify the complexity of invariant measures in dynamical systems, the Kolmogorov-Sinai entropy measures the average rate of information generation or orbit separation. Defined for a measure-preserving transformation T as h_\mu(T) = \sup_{\mathcal{A}} H_\mu(\mathcal{A} | \bigvee_{i=0}^{n-1} T^{-i}\mathcal{A}) / n in the limit n \to \infty over finite partitions \mathcal{A}, it provides a metric invariant distinguishing mixing systems from rigid rotations, with positive values indicating chaotic behavior. This entropy, introduced for Bernoulli shifts and extended to general automorphisms, aligns with thermodynamic entropy in statistical mechanics and upper bounds the topological entropy.

Chaotic Behavior

Defining Chaos

In dynamical systems, chaos refers to a form of complex, aperiodic behavior that arises deterministically yet exhibits profound unpredictability due to extreme sensitivity to initial conditions. This phenomenon occurs in nonlinear systems where small perturbations can lead to exponentially diverging trajectories, distinguishing chaotic dynamics from simple periodic or quasi-periodic motions. While no single universally accepted exists, several rigorous mathematical characterizations capture the essence of , emphasizing topological, metric, and information-theoretic properties. One prominent topological definition, proposed by Robert L. Devaney, describes a continuous f: X \to X on a X as if it satisfies three conditions: (1) topological , meaning there exists a dense —i.e., for some point x \in X, the forward \{f^n(x) \mid n \geq 0\} is dense in X; (2) the set of periodic points is dense in X; and (3) sensitive dependence on initial conditions, where for every point x \in X and every neighborhood U of x, there exists y \in U such that the trajectories of x and y eventually separate by more than any fixed distance. These properties ensure that the system's behavior is both recurrent and unpredictable, with orbits filling the space densely while nearby points diverge rapidly. Devaney's framework, introduced in his seminal 1989 text, provides a clean, purely topological criterion for that applies broadly to maps on intervals, circles, and higher-dimensional spaces. A metric characterization of chaos focuses on Lyapunov exponents, which quantify the average exponential rates of divergence or convergence of nearby trajectories. For a dynamical system, the largest \lambda_{\max} measures the maximal expansion rate; if \lambda_{\max} > 0, trajectories separate exponentially on average, indicating chaotic behavior through sensitive dependence. This condition implies that errors in initial conditions grow as e^{\lambda_{\max} t}, rendering long-term predictions practically impossible despite the system's . Originating from the work of on stability and formalized in the context of via the multiplicative ergodic theorem, positive Lyapunov exponents serve as a quantitative hallmark of in dissipative systems, such as those with strange attractors. Topological entropy offers another measure of chaotic complexity, quantifying the exponential growth rate of the number of distinguishable or periodic points. Defined for a continuous f on a compact as h_{\text{top}}(f) = \lim_{\epsilon \to 0} \lim_{n \to \infty} \frac{1}{n} \log N(n, \epsilon), where N(n, \epsilon) is the minimal number of sets of diameter \epsilon needed to cover the space of n-step itineraries, a positive topological entropy h_{\text{top}}(f) > 0 signifies substantial orbit complexity and is often equated with topological . Introduced by Adler, Konheim, and McAndrew in and refined by Bowen, this captures the "disorder" in the system's , with h_{\text{top}} > 0 implying the existence of dense periodic points and sensitive dependence. Crucially, dynamics differ from processes in their deterministic nature: while both appear unpredictable, stems from nonlinearity and within a fully specified rule set, without inherent , leading to what is termed "deterministic unpredictability." This distinction underscores that systems, though governed by precise equations, mimic in finite-time observations due to , as opposed to probabilistic models where outcomes are fundamentally random.

Routes to Chaos

One prominent route to chaos in dynamical systems involves a sequence of period-doubling bifurcations, where a stable periodic orbit with period n gives rise to a stable orbit with period $2nas a control parameter is varied, leading to an infinite cascade of such doublings that culminates in [chaos](/page/Chaos). This process was first theoretically analyzed by [Mitchell Feigenbaum](/page/Mitchell_Feigenbaum), who demonstrated through [renormalization group](/page/Renormalization_group) methods that the ratios of successive bifurcation intervals converge to a universal [constant](/page/Constant)\delta \approx 4.669$, known as the Feigenbaum constant, independent of the specific as long as it satisfies certain universality conditions. This route has been observed experimentally in diverse systems, such as fluid flows and electronic circuits, confirming the scaling behavior near the accumulation point where the system's becomes positive, marking the onset of . Another mechanism is the quasi-periodic route to , characterized by the successive appearance of tori in , starting from a periodic that bifurcates to a two-dimensional via a , followed by the formation of a three-dimensional under further parameter variation. According to the Ruelle-Takens-Newhouse theorem, generic flows on a two- are unstable and typically lead to the destruction of the and the emergence of a through a sequence involving three incommensurate frequencies, rather than higher-dimensional quasi-periodic motion. This breakdown has been identified in systems like coupled oscillators and fluid convection, where the winding numbers on the deviate from rational values, resulting in chaotic dynamics with a structure. Intermittency provides a third route, where the system alternates between extended laminar phases of nearly periodic behavior and short bursts of chaotic motion, arising near a tangent bifurcation that allows trajectories to linger close to a marginally stable fixed point or manifold. Paul Manneville and Yves Pomeau classified this into types based on the nature of the instability: type I from a saddle-node bifurcation, type II from a subcritical Hopf, and type III from a subcritical flip, each characterized by power-law scaling of the average laminar length with the control parameter distance from the transition. This intermittency has been experimentally verified in nonlinear oscillators and plasma discharges, where the burst frequency increases as the parameter crosses the threshold, leading to fully developed chaos. A canonical example of a system exhibiting chaotic behavior via such routes is the Lorenz attractor, derived from a truncated model of atmospheric , which forms a strange attractor with a butterfly-shaped in three-dimensional . In the Lorenz equations \dot{x} = \sigma(y - x), \dot{y} = x(\rho - z) - y, \dot{z} = xy - \beta z with parameters \sigma=10, \rho=28, \beta=8/3, trajectories are drawn to this fractal set of dimension approximately 2.06, displaying sensitive dependence on initial conditions and positive Lyapunov exponents that confirm . The attractor arises through a subcritical followed by global homoclinic tangencies, illustrating how low-dimensional dissipative systems can produce complex, unpredictable dynamics relevant to weather modeling.

References

  1. [1]
    [PDF] measure theoretic aspects of dynamical systems - UChicago Math
    Definition 1.1. A dynamical system, denoted by (X,f), consists of a non-empty set X called phase space, whose elements represent possible ...<|control11|><|separator|>
  2. [2]
    The idea of a dynamical system - Math Insight
    A dynamical system is all about the evolution of something over time. To create a dynamical system we simply need to decide (1) what is the “something” that ...
  3. [3]
    [PDF] Applied Dynamical Systems - Penn Math - University of Pennsylvania
    YNAMICAL SYSTEMS is the study of behaviors of systems that change over time. Growth, decay, oscillation, evolution, collapse, and chaos are all examples of ...
  4. [4]
    [PDF] Basic Theory of Dynamical Systems - Caltech
    Jan 1, 2011 · Dynamical systems is concerned with both quantitative and qualitative properties of evolution equations, which are often ordinary ...
  5. [5]
    [PDF] Introduction to Dynamical Systems John K. Hunter - UC Davis Math
    (See Definition 1.14 below for a precise definition.) The determination of the stability of equilibria will be an important topic in the following. Other types ...
  6. [6]
    Chapter 13. What are Dynamics and Control?
    A dynamical system is one in which the state of the system changes continuously over time. The notion of state is similar to that of a configuration, although ...
  7. [7]
    [PDF] Lectures on Dynamical Systems - UC Berkeley math
    The main goal of the theory of dynamical system is the study of the global orbit structure of maps and flows. In these notes, we review some ...
  8. [8]
    [PDF] Dynamical Systems Theory - University of California, Santa Barbara
    of 19th century mathematics. It layed the foundation for modern dynamical sys- tems theory along with the work of Birkhoff and Smale and their mathematical ...
  9. [9]
    Chaotic Dynamical Systems (354-0-81)
    The theory of dynamical systems is relatively young in the long history of mathematical discovery and development. Its origins date to the late 19th century ...
  10. [10]
    [PDF] Dynamical systems - Harvard Mathematics Department
    Dynamical system theory has matured into an independent mathematical subject. It is linked to many other areas of mathematics and has its own AMS classification ...
  11. [11]
    Mathematician Studies Dynamical Systems to Find Practical Solutions
    Aug 29, 2017 · Dynamical systems serve as important mathematical models for a wide array of physical phenomena, relating to things such as weather modeling, systems biology, ...
  12. [12]
    Dynamical Systems and Differential Equations | School of Mathematics
    In applications, dynamical systems tools and methods inform modeling in the sciences, they enhance our understanding of phenomena, and they guide decisions in ...
  13. [13]
    Artificial Neural Networks are a New Kind of Dynamical Systems
    Sep 5, 2023 · While applications of dynamical systems are ubiquitous and you will meet them in a variety of science disciplines - from physics to chemistry ...
  14. [14]
    Complex Dynamical Systems
    Jul 24, 2024 · A complex dynamical system is one with interdependent parts that evolve nonlinearly over time. As the system evolves, surprising patterns may emerge.
  15. [15]
    Dynamical Systems and Applications - Barcelona - UB
    Dynamical Systems can be considered as a way to describe evolution problems with respect to time, let them be given by ordinary or partial differential ...
  16. [16]
    Dynamical Systems Applications for Missions to Detect Life in Ocean ...
    Jan 29, 2024 · Dynamical systems theory helps design spacecraft trajectories that optimize fuel consumption and minimize travel time.
  17. [17]
    Introduction to Dynamical Systems
    A dynamical system is a mathematical formalization of a deterministic process, including a state space and a law of state evolution in time.
  18. [18]
    Introduction
    A dynamical system, in this more abstract approach, consists of a set X and a function or trans- formation T defined on X and with values in X. In ergodic ...
  19. [19]
    Linear Dynamical Systems in Rd
    A continuous dynamical system or flow over the 'time set' R with state space X, a metric space, is defined as a continuous map. Φ : R × X −→ X with the ...
  20. [20]
    [PDF] Dynamical Systems and Nonlinear Ordinary Differential Equations
    A deterministic dynamical system is a map T ×M→M, (t, u0) 7→ St(u0) ... the solution of the initial value problem with u(0) = u0 is given by u(t) ...
  21. [21]
    [PDF] DYNAMICAL SYSTEMS - Math-Unipd
    Oct 19, 2021 · ). An important distinction must be made between deterministic dynamical systems ... the Cauchy, or initial value problem associated to the ODE ...
  22. [22]
    Dynamical systems - Scholarpedia
    Feb 9, 2007 · The forward orbit or trajectory of a state s is the time-ordered collection of states that follow from s using the evolution rule. For a ...Definition · Examples · Flows · Iterated function system
  23. [23]
    [PDF] Dynamical systems and ODEs - UC Davis Mathematics
    The subject of dynamical systems concerns the evolution of systems in time. In continuous time, the systems may be modeled by ordinary differential ...Missing: history | Show results with:history
  24. [24]
    [PDF] Introduction to Dynamical Systems - Ceremade
    By analogy with celestial mechanics, the evolution of a particular state of a dynamical system is referred to as an orbit. ... definition of attractors in ...
  25. [25]
    Dynamical Systems: From Classical Mechanics and Astronomy to ...
    Aug 6, 2021 · Dynamical systems arise in several practical real-world situations apart from classical physical systems like astronomy, mechanics, etc.
  26. [26]
    Applications of Dynamical Systems in Biology and Medicine
    This volume highlights problems from a range of biological and medical applications that can be interpreted as questions about system behavior or control.
  27. [27]
    Dynamical Systems: From Classical Mechanics and Astronomy to ...
    Aug 6, 2021 · Dynamical systems arise in several practical real-world situations apart from classical physical systems like astronomy, mechanics, etc.
  28. [28]
    Dynamical Systems in Population Biology - SpringerLink
    In stockThis research monograph provides an introduction to the theory of nonautonomous semiflows with applications to population dynamics.
  29. [29]
    Dynamical System Theory - an overview | ScienceDirect Topics
    The theory of dynamical systems concerns many nonlinear systems and their applications in physics, biology, mathematics, economics, and astronomy (Khalil, 1996 ...
  30. [30]
    [PDF] Applications of Dynamical Systems in Engineering - arXiv
    Dynamical systems with inputs and outputs are sometimes referred to as control systems which is a very important topic in Engineering. Based on the type of ...
  31. [31]
    [PDF] STABILITY IN DYNAMICAL SYSTEMS I +
    A fixed point is stable if initial conditions near it stay near forever; if they blow up exponentially, it is unstable.
  32. [32]
    Some elements for a history of the dynamical systems theory | Chaos
    May 7, 2021 · This theory is a branch of the nonlinear dynamical systems (NDS) theory, which was boosted by Poincaré's works at the late 19th century. It was ...
  33. [33]
    [PDF] Dynamical Systems and Numerical Analysis
    ... numerical simulations are of fundamental importance in gleaning understanding of dynamical systems. Hence it is crucial to understand the behaviour of numerical.<|control11|><|separator|>
  34. [34]
    Galileo: The Telescope & The Laws of Dynamics
    Galileo Galilei (1564-1642) was a pivotal figure ... Galileo made extensive contributions to our understanding of the laws governing the motion of objects.
  35. [35]
    Galileo Galilei - Stanford Encyclopedia of Philosophy
    Jun 4, 2021 · He is renowned for his discoveries: he was the first to report telescopic observations of the mountains on the moon, the moons of Jupiter, the ...Brief Biography · Galileo's Scientific Story · Galileo and the Church · Bibliography
  36. [36]
    [PDF] Basic concepts of dynamical systems theory
    The fundamental step toward the mathematical formalization of reality was taken by Newton and his mechanics, explained in Philosophiae Naturalis Principia ...Missing: Isaac | Show results with:Isaac
  37. [37]
    Principia mathematica (Latin ed.) - Online Library of Liberty
    Newton's most famous work Principia (1687) explains the laws governing the motion of physical objects. Principia rests on the new branch of mathematics that ...
  38. [38]
    [PDF] POINCARÉ'S WORK ON CELESTIAL MECHANICS - arXiv
    The first mathematical tools necessary to understand this kind of real-world phenomena were given by Jules Henri Poincaré (1854-1912) in his work on the ...
  39. [39]
    Henri Poincare, memoir on the three-body problem (1890)
    In the memoir on the three-body problem, Poincaré developed a theory of periodic solutions that opened up an entirely new way of thinking about dynamical ...Missing: methods | Show results with:methods
  40. [40]
    Alexandr Mikhailovich Lyapunov, thesis on the stability of motion ...
    Alexandr Mikhailovich Lyapunov's thesis on the stability of motion memoir is recognized as the first extensive treatise on the stability theory of solutions ...
  41. [41]
    Lyapunov ' s stability theory — 100 years on * | Semantic Scholar
    A brief history of Lyapunov's life and tragic death is given, and following by a section highlighting the important ideas in his thesis of 1892, ...
  42. [42]
    Dynamical systems : Birkhoff, George David, 1884-1944
    Jul 21, 2009 · Dynamical systems. by: Birkhoff, George David, 1884-1944. Publication date: 1927. Topics: Dynamics. Publisher: New York, American Mathematical ...Missing: topological | Show results with:topological
  43. [43]
    [PDF] The Method Topological Roughness of Systems
    Andronov A.A., Pontryagin L.S. (1937) Rough systems//Dokl. Academy of Sciences of the USSR. T.14. No. 5. pp. 247 - 250 ...
  44. [44]
    KAM THEORY: THE LEGACY OF KOLMOGOROV'S 1954 PAPER 1 ...
    Feb 9, 2004 · Kolmogorov-Arnold-Moser (or kam) theory was developed for con- servative dynamical systems that are nearly integrable. Integrable systems in.Missing: 1950s | Show results with:1950s
  45. [45]
    International Symposium on Nonlinear Differential Equations and ...
    International Symposium on Nonlinear Differential Equations and Nonlinear Mechanics. Book • 1963. Edited by: Joseph P. LaSalle and Solomon Lefschetz.
  46. [46]
    [PDF] Ordinary Differential Equations and Dynamical Systems
    This is a preliminary version of the book Ordinary Differential Equations and Dynamical Systems published by the American Mathematical Society (AMS). This ...
  47. [47]
    Geometric Theory of Dynamical Systems - SpringerLink
    In stockGeometric Theory of Dynamical Systems. Book Subtitle: An Introduction. Authors: Jacob Palis, Welington Melo. DOI: https://doi.org/10.1007/978-1-4612-5703-5.
  48. [48]
    [PDF] ERGODIC THEORY AND ENTROPY - UChicago Math
    Definition 2.7. A measure-preserving system is a measure space (X, b,m) that is equipped with a corresponding measure-preserving transformation T:( ...
  49. [49]
    [PDF] measure-preserving dynamical systems and approximation ...
    Aug 29, 2014 · A measure-preserving dynamical system is a measure space and a time rule (T) that describes how points change over time, where T is an ...
  50. [50]
    [PDF] Lecture Notes
    Measure theory is a mature discipline and lies at the heart of ergodic theory ... This means that typical dynamical systems will preserve many invariant measures.
  51. [51]
    [PDF] Measure-preserving dynamical systems on R3 ... - Auburn University
    This satisfies the definition of a dynamical system, but, as above, we cannot guarantee that this preserves Lebesgue measure, or equivalently, Euclidean volume.
  52. [52]
    [PDF] rotations of a torus, the doubling map
    We shall illustrate these two methods by proving that (i) a rotation of a torus, and (ii) the doubling map preserve Lebesgue measure. Let us first recall how ...
  53. [53]
    [PDF] CHAPTER III. MEASURE DYNAMICS. §8. Ergodic Theory
    As examples, the Lebesgue measure on. / is invariant under the angle double map m2 (even though µ(m2(I)) = 2µ(I) for any short interval I ⊂ /. ). If x0 7 ...
  54. [54]
    [PDF] Poincaré, Poincaré Recurrence, and the H-Theorem - PhilArchive
    The second objection used Henri Poincaré's (1854-1912) recurrence theorem resulting in the creation of the so-called recurrence paradox.2 Poincaré's reasoning ( ...
  55. [55]
    [PDF] 3 Discrete Dynamical Systems - University of Bristol
    3.1 Definitions. A first-order discrete dynamical system is a map by which u(n + 1) is determined as a function of u(n), u(n + 1) = f(u(n)),.
  56. [56]
    Dynamical Systems - Scholarpedia
    Feb 9, 2007 · Mathematically, a dynamical system is described by an initial value problem. The implication is that there is a notion of time and that a ...Missing: distinction | Show results with:distinction
  57. [57]
    [PDF] Discrete time dynamics - ChaosBook.org
    Discrete time dynamical systems arise naturally from section 2.1 flows. In general there are two strategies for replacing a continuous-time flow by iterated ...
  58. [58]
    [PDF] 2 Discrete Dynamical Systems: Maps - Complexity Sciences Center
    Given a point xn, the graph of the logistic map provides y D f (xn). To use y as the starting point of the next iteration, we must find the corresponding ...
  59. [59]
    [PDF] Cellular Automata - Dynamical Systems - IFI UZH
    Nov 1, 2013 · • Cellular automata are characterized by a discrete set of states and a discrete time variable. • Identical automata are coupled via a dynamics.
  60. [60]
  61. [61]
    [PDF] Contents 1 Introduction to Dynamics - Evan Dummit
    We study the structure of the orbit of a point under a function, and how to classify the behavior of fixed points and cycles in terms of whether they attract or ...
  62. [62]
    [PDF] 4. The Hamiltonian Formalism - DAMTP
    The phase space of the pendulum is a cylinder R⇥S1, with the R factor corresponding to the momentum. We draw this by flattening out the cylinder. The two ...Missing: source | Show results with:source
  63. [63]
    [PDF] The Dynamics of Pendula: An Introduction to Hamiltonian Systems ...
    Oct 1, 2004 · The phase space is divided into two distinct types of motion by the separatrix: at high energies the pendulum whirls over the top (outside the ...Missing: source | Show results with:source
  64. [64]
    [PDF] Simple mathematical models with very complicated dynamics
    Robert M. May*. First-order difference equations arise in many contexts in the biological, economic and social sciences. Such equations, even though simple ...
  65. [65]
    A two-dimensional mapping with a strange attractor
    Lorenz (1963) has investigated a system of three first-order differential equations, whose solutions tend toward a “strange attractor”.
  66. [66]
    [PDF] The fantastic combinations of John Conway's new solitaire game "life"
    MATHEMATICAL GAMES. The fantastic combinations of John. Conway's new solitaire game "life" by Martin Gardner. Scientific American 223 (October 1970): 120-123.
  67. [67]
    Sharkovsky ordering - Scholarpedia
    Oct 21, 2011 · The Sharkovsky ordering describes the coexistence of cycles with different periods for discrete-time dynamical systems given by maps.Missing: 1D | Show results with:1D
  68. [68]
    [PDF] finite dimensional linear systems - Baillieul Family
    This book is based on a one-semester course on dynamical systems given in the Electrical Engineering Department at the Massachusetts Institute of. Technology ...
  69. [69]
    [PDF] Chapter Five - Linear Systems
    An excellent presentation of linear systems based on the matrix exponential is given in the book by Brockett [Bro70], a more comprehensive treatment is given by.
  70. [70]
    [PDF] The Matrix exponential, Dynamic Systems and Control
    Abstract. The matrix exponential can be found in various connections in analysis and control of dynamic systems. In this short note we are going to list.
  71. [71]
    [PDF] 1 Stability of a linear system - Princeton University
    In this lecture, we consider some applications of SDP: • Stability and stabilizability of linear systems. – The idea of a Lyapunov function. • Eigenvalue and ...
  72. [72]
    [PDF] Linear Oscillations - Physics Courses
    4.1.3 Phase portraits for the damped harmonic oscillator. Expressed as a dynamical system, the equation of motion x + 2β ˙x + ω2. 0x = 0 is written as two ...
  73. [73]
    [PDF] Differential Equations
    FIGURE 4.4.10 Direction field and phase portrait for (a) a critically damped harmonic oscillator. (b) an overdamped harmonic oscillator.
  74. [74]
    [PDF] Nonlinear Systems and Linearization
    To study the behavior of a nonlinear dynamical system near an equilibrium point, we can linearize the system. We will first explain this approach in general ...
  75. [75]
    [PDF] Linearization of Differential Equation Models
    Linearization can be used to give important information about how the system behaves in the neighborhood of equilibrium points. Typically we learn whether the ...
  76. [76]
    [PDF] Math 312 Lecture Notes Linearization - Department of Mathematics
    Mar 23, 2005 · By defining the vector un = un vn , we can write the system in matrix form as un+1 = Jun, (13) where J is the Jacobian matrix given in (9).
  77. [77]
    a lemma in the theory of structural stability of differential equations1
    A LEMMA IN THE THEORY OF STRUCTURAL STABILITY. OF DIFFERENTIAL EQUATIONS1 ... Hartman, On local homeomorphisms of Euclidean spaces, Proceedings of the.
  78. [78]
    A lemma in the theory of structural stability of differential equations
    The object of this note is to answer this question in the affirmative when F(x) is of class C2. (It can be mentioned that, even if F(x) is analytic, ...
  79. [79]
    [PDF] Chapter 9 Stability II: maps and periodic orbits - Full-Time Faculty
    This chapter considers periodic orbits and their stability, using discrete dynamical systems. Stability is defined for fixed points of mappings, and can be ...
  80. [80]
    Stability of equilibria - Scholarpedia
    Mar 15, 2007 · Equilibria can be stable or unstable. Stable equilibria have practical meaning since they correspond to the existence of a certain observable regime.
  81. [81]
    Center manifold - Scholarpedia
    Dec 17, 2006 · A center manifold is an invariant manifold, defined as y=h(x) for small |x| with h(0)=0 and Dh(0)=0, used to simplify dynamical systems.Missing: neutral | Show results with:neutral
  82. [82]
    [PDF] AAS 98-349 TRAJECTORY DESIGN STRATEGIES THAT ...
    An invariant manifold is defined as an m-dimensional surface such that an orbit starting on the surface remains on the surface throughout its dynamical ...<|separator|>
  83. [83]
    LCS Tutorial: Motivation - Shadden Lab
    A fixed point of v is a point xc such that v(xc)=0 . The stable manifolds of a fixed point xc are all trajectories which asymptote to xc when t→∞. Similarly, ...
  84. [84]
    [1708.00480] Hyperbolic dynamics of discrete dynamical systems on ...
    Aug 1, 2017 · We insert a condition in the definition of a hyperbolic set which implies to the unique decomposition of a part of tangent space (at each point ...
  85. [85]
    [PDF] Stable and unstable manifolds. Hyperbolic sets.
    Dynamical Systems and Chaos. Lecture 18: Stable and unstable manifolds ... The local stable manifold of F at p is defined as the local unstable manifold ...
  86. [86]
    [PDF] Hyperbolic Dynamical Systems and Structured stability
    Sep 30, 2024 · Abstract. This paper gives an introduction to discrete dynamical systems, the study of repeated iterations of a function on a space.
  87. [87]
    [PDF] the stable manifold theorem and applications - UChicago Math
    Aug 19, 2019 · This paper presents a proof of the stable manifold theorem, which states that every hyperbolic fixed point has a stable manifold. It also ...
  88. [88]
    Stable Manifold Theory (Chapter 7) - Nonuniform Hyperbolicity
    In this chapter we present one of the principal results of the nonuniform hyperbolicity theory – the existence of local stable and unstable manifolds.
  89. [89]
    Homoclinic tangles-classification and applications - IOPscience
    It is based upon the structure of the homoclinic tangle, and supplies a detailed solution to a transport problem for this class of systems, the characteristics ...
  90. [90]
    [PDF] On the Homoclinic Tangles of Henri Poincaré - Arizona Math
    In this paper, we introduce to the reader a recent theory on the dynamics of homoclinic tangles of equation (1.1). Our objective is to understand and to ...
  91. [91]
    Attractor - Scholarpedia
    Nov 3, 2006 · An attracting set for a dynamical system is a closed subset A of its phase space such that for many choices of initial point the system will evolve towards A.
  92. [92]
    Basin of attraction - Scholarpedia
    Dec 29, 2016 · For each such attractor, its basin of attraction is the set of initial conditions leading to long-time behavior that approaches that attractor.Example · Fractal basin boundaries · Riddled Basins of Attraction
  93. [93]
    [PDF] Ruelle: Strange attractors - Harvard Mathematics Department
    This means that A is attracting. (c) There is sensitive dependence on initial condition when. Xo is in U. This makes A a strange attractor.
  94. [94]
    Morse-Smale systems - Scholarpedia
    Apr 24, 2013 · More-Smale systems are the simplest dynamical systems. They are structurally stable and have intimate connections to the topology of manifolds.Dynamical Systems · Morse-Smale Dynamical... · Morse-Smale Gradient Fields...
  95. [95]
    Bifurcation - Scholarpedia
    Jun 14, 2007 · A bifurcation of a dynamical system is a qualitative change in its dynamics produced by varying parameters.
  96. [96]
    Saddle-node bifurcation - Scholarpedia
    Oct 1, 2015 · A saddle-node bifurcation is a collision and disappearance of two equilibria in dynamical systems. In systems generated by autonomous ODEs, ...Missing: seminal | Show results with:seminal
  97. [97]
    Transcritical Bifurcation -- from Wolfram MathWorld
    A transcritical bifurcation has two branches, one stable and one unstable. It is defined by conditions on a one-parameter family of C^2 maps.Missing: seminal reference
  98. [98]
    Pitchfork Bifurcation -- from Wolfram MathWorld
    This type of bifurcation is called a pitchfork bifurcation. An example of an equation displaying a pitchfork bifurcation is x^.=mux-x^3.
  99. [99]
    Andronov-Hopf bifurcation - Scholarpedia
    Oct 2, 2006 · Andronov-Hopf bifurcation is the birth of a limit cycle from an equilibrium in dynamical systems generated by ODEs, when the equilibrium changes stability.Two-dimensional Case · Multi-dimensional Case · First Lyapunov Coefficient
  100. [100]
    [PDF] Elements of Applied Bifurcation Theory, Second Edition
    This book covers practical bifurcation theory for dynamical systems, including 2D and 3D cases, and is for students and researchers in various fields.<|separator|>
  101. [101]
    [PDF] Homoclinic Bifurcations and Sensitive-Chaotic Dynamics - ICTP
    Sep 9, 1991 · Homoclinic bifurcations, which form the main topic of this monograph, belong to the area of dynamical systems, the theory which describes ...
  102. [102]
    Quantitative universality for a class of nonlinear transformations
    Cite this article. Feigenbaum, M.J. Quantitative universality for a class of nonlinear transformations. ... DOI : https://doi.org/10.1007/BF01020332. Share ...
  103. [103]
    [PDF] Example of a Blue Sky Catastrophe
    Abstract. We present a low-order system of ODEs exhibiting the blue sky catastrophe—a new codimension one bifurcation of periodic orbits.
  104. [104]
    Leonid Shilnikov and mathematical theory of dynamical chaos
    Jan 18, 2022 · As an answer, he proposed a universal scenario of spiral chaos formation24 along the following lines. It is common that the system, at some ...
  105. [105]
    [PDF] Ergodic theory, geometry and dynamics
    Dec 24, 2020 · In topological dynamics, one requires that T is a continuous map in the product topology; in measurable dynamics, that T is a measurable map.
  106. [106]
    [PDF] What are SRB measures, and which dynamical systems have them?
    The original work of Sinai, Ruelle and Bowen was carried out in the context of. Anosov and Axiom A systems. For these dynamical systems they identified and.
  107. [107]
    [PDF] Lecture Notes on Ergodic Theory - Weizmann Institute of Science
    Dynamical systems and ergodic theory. Ergodic theory is a part of the theory of dynamical systems. At its simplest form, a dynamical system is a function T ...
  108. [108]
    Kolmogorov-Arnold-Moser theory - Scholarpedia
    Sep 23, 2010 · Kolmogorov-Arnold-Moser (KAM) theory deals with persistence, under perturbation, of quasi-periodic motions in Hamiltonian dynamical systems.Classical KAM theory · Remarks · Properly degenerate KAM theory
  109. [109]
    [PDF] An Introduction to KAM Theory
    Jan 22, 2008 · Over the past thirty years, the Kolmogorov-Arnold-Moser (KAM) theory has played an important role in increasing our understanding of the ...
  110. [110]
    [PDF] Metric Entropy of Dynamical System - Princeton Math
    Entropy plays an important role in the theory of deterministic chaos or chaos theory because it characterizes the intrinsic instability of dynamics and the ...
  111. [111]
    [PDF] A. N. Kolmogorov's and Y. G. Sinai's papers introducing entropy of ...
    Sinai, On the notion of entropy of a dynamical system, A free translation from the Russian original by the author, Selecta, Volume I: Ergodic Theory and ...
  112. [112]
    Chaos - Stanford Encyclopedia of Philosophy
    Jul 16, 2008 · His definition picks out two key features that are simultaneously present: instability and aperiodicity. Unstable systems are those exhibiting ...
  113. [113]
    Lyapunov exponent - Scholarpedia
    Oct 30, 2013 · A strictly positive maximal Lyapunov exponent is often considered as a definition of deterministic chaos. This makes sense only when the ...
  114. [114]
    Topological entropy - Scholarpedia
    Oct 28, 2013 · Topological dynamical systems of positive entropy are often considered topologically chaotic. Positive entropy always implies Li-Yorke chaos ...
  115. [115]
    Period doubling, Feigenbaum constant and time series prediction in ...
    May 15, 2009 · In this paper, we use a real RLD circuit and follow its route to chaos through period doubling. We consider this work in continuation of the ...
  116. [116]
    Occurrence of strange AxiomA attractors near quasi periodic flows ...
    About this article. Cite this article. Newhouse, S., Ruelle, D. & Takens, F. Occurrence of strange AxiomA attractors near quasi periodic flows onT m,m≧3.Missing: theorem paper
  117. [117]
    Quasiperiodic routes to chaos in confined two-dimensional ...
    Oct 26, 2015 · The Newhouse-Ruelle-Takens [47] theorem later refined this statement by actually allowing T 3 to exist and be structurally stable under more ...
  118. [118]
    Observation of a Pomeau-Manneville intermittent route to chaos in a ...
    Oct 1, 1982 · We report an additional route to chaos: the Pomeau-Manneville intermittency route, characterized by a periodic (laminar) phase interrupted by bursts of ...
  119. [119]
    [PDF] Deterministic nonperiodic flow - Semantic Scholar
    Deterministic nonperiodic flow · E. Lorenz · Published 1 March 1963 · Physics, Environmental Science · Journal of the Atmospheric Sciences.