Fact-checked by Grok 2 weeks ago

Phase plane

In the study of dynamical systems, a phase plane is a two-dimensional graphical representation that depicts the behavior of solutions to a system of two ordinary differential equations, with the axes corresponding to the system's state variables and trajectories showing how solutions evolve over time without explicit dependence on the independent variable. This tool is particularly useful for analyzing autonomous systems of the form \dot{x} = f(x, y), \dot{y} = g(x, y), where trajectories form curves in the plane that do not intersect due to the uniqueness of solutions guaranteed by theorems like Picard's. Phase plane analysis reveals qualitative behaviors such as equilibrium points—where f(x_0, y_0) = 0 and g(x_0, y_0) = 0—which can be stable (attracting nearby trajectories), unstable (repelling them), or centers (indicating periodic motion). For linear systems \vec{x}' = A \vec{x}, the phase portrait near the origin (the typical equilibrium) is classified based on the eigenvalues of A, yielding types like nodes, saddles, spirals, or centers, with stability determined by the real parts of the eigenvalues. In nonlinear systems, the method extends to identify more complex features, including limit cycles—isolated closed trajectories that attract or repel nearby paths—and attractors like spirals where trajectories converge as time approaches infinity. By sketching vector fields or using computational tools to plot these portraits, researchers can infer long-term system dynamics, stability, and bifurcations without solving the equations analytically, making it essential in fields like physics, engineering, and biology.

Fundamentals

Definition and Purpose

A phase plane is a two-dimensional graphical construct employed in the study of second-order autonomous dynamical systems, with each axis representing one of the system's variables—typically and —and trajectories depicted as curves that illustrate the system's temporal evolution without explicit time parameterization. These trajectories arise from solutions to the system's governing equations, providing a visual map of how states transition from initial conditions. Such systems are inherently , meaning their dynamics are governed by ordinary differential equations (ODEs) where the rates of change depend solely on the current state and not on time explicitly: \begin{align*} \frac{dx}{dt} &= f(x, y), \\ \frac{dy}{dt} &= g(x, y), \end{align*} with f and g as smooth functions of the state variables x and y. This autonomy ensures that the phase plane captures invariant behavioral patterns, independent of the starting time. The purpose of the phase plane lies in facilitating qualitative analysis of dynamical behavior, bypassing the often intractable task of obtaining closed-form solutions to the ODEs; instead, it reveals structural properties like the location and of points (where \frac{dx}{dt} = 0 and \frac{dy}{dt} = 0), the presence of attractors drawing nearby trajectories inward, repellors pushing them outward, and periodic orbits forming closed loops indicative of sustained oscillations. This approach is particularly valuable for understanding long-term system tendencies and global flow patterns in the state space. The phase plane concept originated in the late 19th century within Henri Poincaré's foundational work on the qualitative theory of differential equations, where he introduced geometric methods to describe arrangements in without relying on explicit .

Representation of Trajectories

In the phase plane, for a general two-dimensional autonomous system \dot{x} = f(x, y), \dot{y} = g(x, y) are represented as solution curves (x(t), y(t)) parameterized by time t, tracing the evolution of the state variables from initial conditions. These curves illustrate the qualitative behavior of the system, with the direction of motion at each point given by the (f(x, y), g(x, y)), which indicates both the direction and the instantaneous . The magnitude of this vector, \sqrt{f(x, y)^2 + g(x, y)^2}, represents the speed along the , allowing time to be inferred by integrating the divided by this speed, though phase portraits often emphasize geometric flow over explicit temporal scaling. To construct trajectories graphically, a (or direction field) is commonly used, where a grid of points in the phase plane is overlaid with short line segments whose slopes approximate \frac{dy}{dx} = \frac{g(x, y)}{f(x, y)}, provided f(x, y) \neq 0. These segments visualize the local direction of the , enabling trajectories to be sketched by connecting aligned segments while following the indicated directions; or software tools can refine this for precise curves. points, where \dot{x} = 0 and \dot{y} = 0 (i.e., f(x, y) = 0 and g(x, y) = 0), appear as singular points in the field with no defined direction, serving as fixed locations where trajectories may converge or diverge. Trajectories exhibit key properties rooted in the theory of differential equations: by the existence and uniqueness theorem for Lipschitz-continuous right-hand sides, solutions are for given conditions, ensuring that trajectories do not intersect or cross each other in the phase plane. This non-intersection preserves the topological structure of the flow, with trajectories forming a of the plane except at equilibria. Isoclines provide an auxiliary tool for sketching, consisting of curves where the slope \frac{g(x, y)}{f(x, y)} = k is constant for some fixed k, partitioning the plane into regions of uniform direction. Along each isocline, parallel line segments are drawn with slope k, facilitating the approximation of the slope field and the connection of trajectories across these loci.

Linear Dynamical Systems

Eigenvalue Decomposition

In the analysis of linear dynamical systems within the phase plane, the fundamental equation is the autonomous system \dot{x} = Ax, where x \in \mathbb{R}^2 represents the state vector, A is a constant 2×2 matrix, and the origin serves as the equilibrium point. The general solution to this system is given by x(t) = e^{At} x(0), where e^{At} is the matrix exponential, which encapsulates the evolution of initial conditions x(0) over time. To compute this solution explicitly, eigenvalue decomposition is employed by solving the characteristic equation \det(A - \lambda I) = 0, which yields the eigenvalues \lambda_1 and \lambda_2. For distinct real eigenvalues, the corresponding eigenvectors v_1 and v_2 are found by solving (A - \lambda_i I)v_i = 0 for i=1,2, allowing the matrix A to be diagonalized as A = P \Lambda P^{-1}, with \Lambda = \diag(\lambda_1, \lambda_2) and P = [v_1 \, v_2]. The general solution then takes the form x(t) = c_1 v_1 e^{\lambda_1 t} + c_2 v_2 e^{\lambda_2 t}, where c_1 and c_2 are constants determined by initial conditions. In the phase plane, this decomposition reveals the structure of trajectories: solutions along the eigenvector directions v_1 and v_2 are straight lines emanating from or approaching the origin, depending on the signs of \lambda_1 and \lambda_2. General trajectories, as linear combinations of these eigensolutions, curve in the , with their qualitative behavior governed by the eigenvalue signs—negative values draw paths toward the origin, while positive values repel them away. Degenerate cases arise when eigenvalues are repeated or complex, requiring modifications such as generalized eigenvectors for repeated roots or real and imaginary parts for complex conjugates, though full details on these are addressed in subsequent classifications.

Classification of Equilibrium Points

In linear two-dimensional dynamical systems of the form \dot{\mathbf{x}} = A \mathbf{x}, where A is a constant $2 \times 2 matrix, the origin is the sole equilibrium point, as homogeneous linear systems possess no other fixed points. The qualitative behavior of trajectories in the phase plane near this equilibrium is determined by the eigenvalues of A, which dictate the type and stability of the fixed point. Stability is assessed by the real parts of the eigenvalues: the equilibrium is asymptotically stable if all real parts are negative, unstable if any are positive, and neutrally stable (but not asymptotically) if all real parts are zero. Linear systems exhibit no limit cycles, as trajectories either converge to, diverge from, or orbit the origin without periodic attractors disconnected from it. For real and distinct eigenvalues \lambda_1 < \lambda_2, the phase portrait depends on their signs. If both are positive (\lambda_1 > 0), the equilibrium is an unstable node (source), where trajectories radiate outward from the along the directions of the corresponding eigenvectors, diverging to infinity. Conversely, if both are negative (\lambda_2 < 0), it forms a stable node (sink), with trajectories converging toward the , typically approaching faster along the eigenvector for the more negative eigenvalue. When the eigenvalues have opposite signs, the equilibrium is a saddle point, which is unstable; trajectories along the eigenvector for the negative eigenvalue approach the (stable manifold), while those along the positive one diverge (unstable manifold), creating hyperbolic paths that separate regions of inflow and outflow. Repeated real eigenvalues \lambda occur when the characteristic polynomial has a double root, leading to two subcases based on the geometric multiplicity. If the eigenspace is two-dimensional (i.e., A is a scalar multiple of the identity matrix), the equilibrium is a proper node (or star node); trajectories approach or recede radially from the origin along all directions if \lambda < 0 (stable) or \lambda > 0 (unstable), respectively. In the defective case, with only one independent eigenvector (Jordan canonical form), it is an improper node; stability follows the sign of \lambda (stable if \lambda < 0, unstable if \lambda > 0), but trajectories curve toward or away from the origin, aligning asymptotically with the single eigenvector direction rather than radiating uniformly. Complex conjugate eigenvalues \alpha \pm i\beta with \beta \neq 0 produce rotational behavior in the phase plane. If \alpha < 0, the equilibrium is a stable spiral (spiral sink), where trajectories spiral inward toward the origin, combining decay with oscillation. For \alpha > 0, it is an unstable spiral (spiral source), with trajectories spiraling outward. When \alpha = 0 (purely imaginary eigenvalues), the equilibrium is a center, neutrally stable, featuring closed elliptical orbits around the origin that neither grow nor decay, corresponding to periodic solutions. The following table summarizes the classifications for the in planes:
Eigenvalue TypeSubtype/BehaviorPhase Plane Description
Real, distinct, both >0Unstable ()UnstableTrajectories diverge along eigenvectors.
Real, distinct, both <0Stable (sink)Asymptotically stableTrajectories converge along eigenvectors.
Real, distinct, opposite signsSaddleUnstableHyperbolic trajectories; inflow on one axis, outflow on the other.
Real, repeated, full eigenspaceProper (star)Stable if \lambda < 0; unstable if \lambda > 0Radial approach/recession from .
Real, repeated, defectiveImproper Stable if \lambda < 0; unstable if \lambda > 0Curved trajectories aligning with single eigenvector.
, \alpha < 0, \beta \neq 0Stable spiral (sink)Asymptotically stableInward spiraling orbits.
, \alpha > 0, \beta \neq 0Unstable spiral ()UnstableOutward spiraling orbits.
Purely imaginary (\alpha = 0)Neutrally stableClosed elliptical orbits.

Nonlinear Dynamical Systems

Phase Portraits and Fixed Points

In nonlinear dynamical systems described by the autonomous equations \dot{x} = f(x, y) and \dot{y} = g(x, y), the constitutes a comprehensive geometric representation in the phase plane, depicting the (\dot{x}, \dot{y}) = (f(x, y), g(x, y)) along with all trajectories and fixed points where \dot{x} = 0 and \dot{y} = 0, or equivalently f(x^*, y^*) = 0 and g(x^*, y^*) = 0. This portrait reveals the global flow and qualitative long-term behavior, such as attraction to or repulsion from fixed points, without requiring explicit solutions to the equations. Fixed points in nonlinear systems can be multiple and scattered across the phase plane, unlike the single origin typical in linear cases; for instance, the Lotka-Volterra competition model for two species exhibits up to four fixed points, including coexistence and extinction equilibria, depending on parameter values. To analyze local behavior near a fixed point (x^*, y^*), the system undergoes linearization via the Jacobian matrix D\mathbf{F}(x^*, y^*), where \mathbf{F}(x, y) = (f(x, y), g(x, y)) and the entries are partial derivatives: D\mathbf{F} = \begin{pmatrix} \frac{\partial f}{\partial x} & \frac{\partial f}{\partial y} \\ \frac{\partial g}{\partial x} & \frac{\partial g}{\partial y} \end{pmatrix}_{(x^*, y^*)}. The eigenvalues of this matrix determine the local classification as a node, saddle, spiral, or center, mirroring linear system types, provided the fixed point is hyperbolic (no zero real-part eigenvalues). This approximation holds in a neighborhood of the fixed point by the Hartman-Grobman theorem, which guarantees topological conjugacy between the nonlinear flow and its linearization near hyperbolic equilibria. In nonlinear contexts, centers from linear analysis are rare, as small perturbations often destabilize them into spirals; multiple fixed points enable complex interactions, such as separatrices dividing the plane into regions of distinct behaviors. Globally, phase portraits may feature homoclinic orbits, which are trajectories connecting a fixed point to itself, or heteroclinic orbits linking distinct , both of which can bound regions or define sets. Basins of delineate sets of initial conditions converging to specific attractors, such as a stable , with boundaries often formed by stable manifolds of . Qualitative sketching of these portraits involves plotting nullclines (curves where \dot{x} = 0 or \dot{y} = 0), locating fixed points, linearizing locally, and tracing trajectories consistent with the .

Nullclines and Bifurcations

In the analysis of nonlinear dynamical represented in the phase plane, nullclines provide a fundamental tool for understanding the structure of the without solving the differential equations explicitly. For a two-dimensional autonomous \dot{x} = f(x, y), \dot{y} = g(x, y), the x-nullcline is defined as the curve where f(x, y) = 0, along which the horizontal component of the vanishes, resulting in vertical tangents to the trajectories. Similarly, the y-nullcline is the set where g(x, y) = 0, yielding horizontal tangents to the trajectories. The intersections of these nullclines correspond precisely to the fixed points of the , where both \dot{x} = 0 and \dot{y} = 0. Moreover, the nullclines partition the phase plane into regions where the signs of f(x, y) and g(x, y) remain constant, enabling a qualitative determination of the direction and nature of the flow in each sector—for instance, where f > 0 and g > 0, trajectories move rightward and upward. This sign analysis facilitates sketching approximate phase portraits and identifying basins of attraction. To assess the of fixed points located at intersections, is employed around each (x^*, y^*). The matrix of the system is given by J = \begin{pmatrix} \frac{\partial f}{\partial x} & \frac{\partial f}{\partial y} \\ \frac{\partial g}{\partial x} & \frac{\partial g}{\partial y} \end{pmatrix}_{(x^*, y^*)}, with \tau = \frac{\partial f}{\partial x} + \frac{\partial g}{\partial y} and \Delta = \frac{\partial f}{\partial x} \frac{\partial g}{\partial y} - \frac{\partial f}{\partial y} \frac{\partial g}{\partial x}, evaluated at the fixed point. The eigenvalues \lambda satisfy the \lambda^2 - \tau \lambda + \Delta = 0, so \lambda = \frac{\tau \pm \sqrt{\tau^2 - 4\Delta}}{2}. is determined as follows: the fixed point is a if \Delta < 0; asymptotically stable if \tau < 0 and \Delta > 0 (a if \tau^2 - 4\Delta > 0, or a stable spiral if \tau^2 - 4\Delta < 0); and unstable if \tau > 0 and \Delta > 0 (unstable or spiral). For complex eigenvalues with nonzero imaginary part, the real part ( \tau/2 ) dictates : negative for attracting spirals, positive for repelling. This classification holds for hyperbolic fixed points (eigenvalues with nonzero real parts) via the Hartman-Grobman theorem. Bifurcations occur when a \mu is varied, leading to qualitative changes in the portrait's , often at intersections. In a , a pair of fixed points—one stable and one —collide and annihilate as \mu crosses a , altering the number of equilibria from two to zero. A generic normal form in one dimension is \dot{x} = \mu - x^2, where fixed points exist at x = \pm \sqrt{\mu} for \mu > 0 and disappear for \mu < 0; in two dimensions, similar behavior arises in systems like the driven pendulum or autocatalytic reactions, where fold and touch tangentially at the bifurcation point. For instance, consider \dot{x} = \mu - x^2, \dot{y} = -y: the x- consists of vertical lines at x = \pm \sqrt{\mu} (for \mu > 0), and as \mu varies, the fixed points at (\pm \sqrt{\mu}, 0) coalesce along the y- y=0 at \mu = 0. The , conversely, involves a fixed point changing without disappearing, typically giving rise to a . At the critical parameter value, the eigenvalues cross the imaginary axis (purely imaginary with \tau = 0, \Delta > 0), transitioning from a stable spiral to unstable with an emerging periodic orbit. In the supercritical case, a stable forms for \mu > \mu_c, with scaling as \sqrt{\mu - \mu_c}; the subcritical variant produces an unstable cycle that can lead to global jumps in attractors. A normal form is \dot{x} = \mu x - y - x (x^2 + y^2), \quad \dot{y} = x + \mu y - y (x^2 + y^2), where the origin is stable for \mu < 0 and a stable appears for \mu > 0, observable in models like the via cubic nullclines.

Applications

Physical Systems

Phase plane analysis provides valuable insights into the behavior of classical mechanical systems, particularly those exhibiting oscillatory motion with damping. A prototypical example is the damped harmonic oscillator, modeled by the second-order differential equation \ddot{x} + \gamma \dot{x} + \omega^2 x = 0, where \gamma > 0 is the damping coefficient and \omega > 0 is the natural frequency. In the phase plane, with coordinates (x, y) where y = \dot{x}, this system is represented by the first-order equations: \dot{x} = y, \quad \dot{y} = -\omega^2 x - \gamma y. The origin (0, 0) is the sole fixed point, which is asymptotically stable for all \gamma > 0. The qualitative structure of trajectories in the phase plane depends on the damping regime, determined by the discriminant \Delta = \gamma^2 - 4\omega^2. For the undamped case (\gamma = 0), trajectories are closed elliptical orbits centered at the origin, representing periodic oscillations with conserved energy E = \frac{1}{2} \omega^2 x^2 + \frac{1}{2} y^2. As \gamma increases slightly (underdamped regime, $0 < \gamma < 2\omega), the fixed point becomes a stable spiral node, with trajectories spiraling inward toward the origin, illustrating energy dissipation while maintaining oscillatory decay. In the critically damped case (\gamma = 2\omega), trajectories approach the origin along nodal paths without oscillation, and for the overdamped case (\gamma > 2\omega), the fixed point is a stable node with monotonic convergence along separatrices. These energy-like contours, derived from the system's Lyapunov function, highlight the transition from sustained to decaying motion. Another mechanical system amenable to phase plane analysis is the simple pendulum with , governed by \ddot{\theta} + b \dot{\theta} + \sin \theta = 0, where \theta is the , b > 0 is the , and the is normalized by the pendulum length and gravity. In phase plane coordinates (\theta, \omega) with \omega = \dot{\theta}, the equations become: \dot{\theta} = \omega, \quad \dot{\omega} = -\sin \theta - b \omega. Fixed points occur at (\theta, \omega) = (k\pi, 0) for k, with sinks at even multiples (2k\pi, 0) and unstable saddles at odd multiples ((2k+1)\pi, 0). causes all trajectories to spiral toward the nearest fixed point, without forming a ; however, separatrix curves—homoclinic loops to the saddles—enclose regions around each point, bounding librations (oscillations) from rotations (circulations) in the undamped limit. These separatrices resemble contours E = \frac{1}{2} \omega^2 - \cos \theta = constant, with eroding along trajectories, leading to eventual settling at the downward . The Van der Pol oscillator exemplifies nonlinear damping in mechanical contexts, such as self-sustained vibrations in electrical circuits modeling mechanical relays. Its standard form is \ddot{x} - \mu (1 - x^2) \dot{x} + x = 0, where \mu > 0 controls nonlinearity strength. In the phase plane (x, y) with y = \dot{x}, the system is: \dot{x} = y, \quad \dot{y} = \mu (1 - x^2) y - x. The origin is the only fixed point, stable for \mu \leq 0 but undergoing a supercritical Hopf bifurcation at \mu = 0, beyond which it becomes unstable and a unique stable limit cycle emerges, attracting all trajectories regardless of initial conditions. For small \mu > 0, the limit cycle is nearly circular with radius approximately 2; as \mu increases, it distorts into a relaxation oscillation, with slow and fast segments resembling energy dissipation contours in a potential well modulated by negative damping near the origin. This structure underscores self-excitation, where energy input balances nonlinear losses.

Biological and Chemical Models

Phase plane analysis plays a crucial role in understanding the dynamics of interacting s in biological systems, particularly through models like the Lotka-Volterra equations. The classic predator-prey version of this model describes the interaction between a prey x and a predator y, governed by the system \dot{x} = \alpha x - \beta x y, \quad \dot{y} = \delta x y - \gamma y, where \alpha > 0 is the prey growth rate, \beta > 0 is the predation rate, \delta > 0 is the predator growth efficiency from predation, and \gamma > 0 is the predator death rate. This formulation, originally proposed by Lotka, captures oscillatory behavior in populations. The fixed points are at (0,0) and (\gamma/\delta, \alpha/\beta). In the phase plane, the origin (0,0) is a , representing unstable coexistence of for both species, while the interior fixed point (\gamma/\delta, \alpha/\beta) is a surrounded by closed orbits. These closed trajectories indicate neutral , where populations cycle periodically without damping or amplification, reflecting sustained predator-prey oscillations. The nullclines for this model are the vertical line x = \gamma/\delta (where \dot{y} = 0) and the horizontal line y = \alpha/\beta (where \dot{x} = 0), dividing the phase plane into regions of growth and decline. Trajectories follow these isoclines, with prey increasing left of the vertical nullcline and predators increasing below the horizontal one, leading to the characteristic elliptical cycles around the center. This neutral cycling interprets biological coexistence as perpetual fluctuation, where predator reduces prey, subsequently starving predators and allowing prey recovery— a key insight into ecological balance without external forcing. Extending the Lotka-Volterra framework to competitive interactions between two x and y, the model incorporates intraspecific and , typically written in nondimensional form as \dot{u}_1 = u_1 (1 - u_1 - a_{12} u_2), \quad \dot{u}_2 = \rho u_2 (1 - u_2 - a_{21} u_1), where u_1, u_2 are scaled densities, \rho > 0 is the , and a_{12}, a_{21} > 0 are competition coefficients measuring interspecific effects relative to intraspecific ones. The fixed points include (0,0), (1,0), (0,1/\rho), and the interior point u_1^* = (1 - a_{12}) / (1 - a_{12} a_{21}), u_2^* = (1 - a_{21}) / (1 - a_{12} a_{21}) (provided it lies in the positive quadrant). The nullclines are straight lines: u_1 + a_{12} u_2 = 1 for \dot{u}_1 = 0 and a_{21} u_1 + u_2 = 1 for \dot{u}_2 = 0, whose intersection determines the interior . Stability in the phase plane depends on the relative strengths of competition. When a_{12} < 1 and a_{21} < 1 (intraspecific competition dominates), the interior fixed point is a stable node, attracting trajectories to coexistence where both species persist at densities. Conversely, if a_{12} > 1 and a_{21} > 1 (interspecific competition stronger), the interior point becomes a , leading to competitive exclusion: trajectories diverge to one of the axial fixed points, where one species drives the other to based on initial conditions, illustrating . If one coefficient exceeds 1 and the other is less, exclusion favors the species less affected by competition, with the winner's axial stable. These outcomes highlight phase plane trajectories converging to or repelling from equilibria, interpreting ecological principles like the . In chemical systems, phase plane methods reveal oscillatory kinetics and bifurcations, as exemplified by the model, a prototype for autocatalytic reactions far from : \dot{x} = A + x^2 y - (B + 1) x, \quad \dot{y} = B x - x^2 y, where A > 0 and B > 0 are constants related to reaction rates and concentrations. Proposed by Prigogine and Lefever, this model demonstrates symmetry-breaking instabilities. The unique positive fixed point is at (x^*, y^*) = (A, B/A). For B < 1 + A^2, this is stable, with trajectories spiraling inward in the phase plane. However, as B increases beyond $1 + A^2, a supercritical occurs, transforming the fixed point into an unstable focus surrounded by a stable . The nullclines are the parabola y = (B + 1)/x - A/x^2 for \dot{x} = 0 and the line y = B/x for \dot{y} = 0, intersecting at the fixed point. Post-, phase plane trajectories converge to the , representing sustained chemical oscillations, such as in or Belousov-Zhabotinsky reactions. This underscores how parameter changes can shift from steady states to periodic behavior, interpreting as sensitivity to in open systems. In both biological and chemical contexts, phase planes elucidate and coexistence: closed orbits or denote neutral or sustained in interacting populations and reactions, while nodes and saddles reveal pathways to persistence or exclusion.

References

  1. [1]
    Differential Equations - Phase Plane - Pauls Online Math Notes
    Nov 16, 2022 · In this section we will give a brief introduction to the phase plane and phase portraits. We define the equilibrium solution/point for a ...
  2. [2]
    [PDF] Phase Plane Methods
    The equilibrium solutions or equilibria are found by solving the nonlinear equations. F(x0,y0)=0, G(x0,y0)=0. Each such (x0,y0) in D is a trajectory whose ...
  3. [3]
    10.5: Phase Plane Analysis - Attractors, Spirals, and Limit cycles
    Mar 11, 2023 · By definition, at least one trajectory spirals into the limit cycle as time approaches either positive or negative infinity. Like a sink, ...
  4. [4]
    [PDF] Dynamical systems and ODEs - UC Davis Math
    We define the corresponding Poincaré map Φ : Rd → Rd by. Φ : x(0) 7→ x(1). Then Φ defines an autonomous discrete dynamical system of dimension d, which is ...Missing: history | Show results with:history
  5. [5]
    [PDF] 27 Qualitative behavior of linear systems - MIT OpenCourseWare
    Be able to draw the phase portrait of any linear, autonomous, second-order system. ... 27.3 The phase plane: example with definitions. 1. Example 27.1. Let ...
  6. [6]
    [PDF] Chapter 2 Phase Plane Analysis - WordPress.com
    Phase plane analysis is a graphical method for studying second-order systems, which was introduced well before the turn of the century by mathematicians ...
  7. [7]
    [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 ...<|separator|>
  8. [8]
    Qualitative theory of differential equations
    Jun 6, 2020 · The foundations of the qualitative theory of differential equations were laid at the end of the 19th century by H. Poincaré (see , [2]) and A.M. ...<|separator|>
  9. [9]
    [PDF] Nonlinear Dynamics and Chaos
    May 6, 2020 · Welcome to this second edition of Nonlinear Dynamics and Chaos, now avail- able in e-book format as well as traditional print.
  10. [10]
    [PDF] Linear Systems Notes for CDS-140a
    Jul 1, 2011 · With these eigenvalues the phase plots are said to have a “focus” or a “spiral.” We further clasify them as “unstable” if τ > 0 (or equivalently.
  11. [11]
    [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).
  12. [12]
    Phase Portraits of Linear Systems - UBC Math
    The path travelled by the point in a solution is called a trajectory of the system. A picture of the trajectories is called a phase portrait of the system.
  13. [13]
    Phase Portraits for systems with real eigenvalues
    A linear system for which we have both negative, distinct eigenvalues has an equilibrium point that is called a asymptotically stable node (sink) .
  14. [14]
    [PDF] Phase Plane Methods
    A graphic of trajectories drawn as parametric curves in the xy-plane is called a phase portrait and the xy-plane in which it is drawn is called the phase plane.Missing: parameterization arc speed
  15. [15]
    Differential Equations - Repeated Eigenvalues
    Nov 16, 2022 · ... solution to the system. We will also show how to sketch phase portraits associated with real repeated eigenvalues (improper nodes).Missing: proper | Show results with:proper
  16. [16]
    [PDF] Nonlinear Systems
    (b) Find the phase portrait of the nonlinear system near the origin and show that the portrait resembles a stable focus. Hint: Transform the equations into ...
  17. [17]
    a lemma in the theory of structural stability of differential equations1
    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,.
  18. [18]
    [PDF] A quick guide to sketching phase planes - UC Berkeley MCB
    To sketch the phase plane of such a system, at each point (x0,y0) in the xy-plane, we draw a vector starting at (x0,y0) in the direction f(x0,y0)i + g(x0,y0)j.
  19. [19]
    [PDF] Nullclines and phaseplanes
    Sep 25, 2002 · We will introduce three main concepts: (i) fixed points, (ii) stability, and (iii) nullclines. The latter are curves that enable one to break ...
  20. [20]
  21. [21]
    Motion of Driven Damped Oscillator in Phase Space - SpringerLink
    Aug 1, 2023 · In this chapter, we have numerically plotted trajectories of simple harmonic oscillator, damped harmonic oscillator and driven damped harmonic oscillator in ...Missing: plane sources
  22. [22]
    [PDF] Phase Portrait Analysis for Physical Systems
    Jun 2, 1999 · Figure 2: Phase Portrait for the Simple Damped Pendulum. Duffing ... It is interesting to compare the pendulum phase plane to that of the duffing ...
  23. [23]
    [PDF] Chapter Seven: The Pendulum and phase-plane plots
    Now, experimentally, if the length is fixed, the period of the oscillation of a pendulum is found to be very nearly independent of how far it is swinging, that ...Missing: scholarly | Show results with:scholarly
  24. [24]
    Van der Pol oscillator - Scholarpedia
    Jan 8, 2007 · This model was proposed by Balthasar van der Pol (1889-1959) in 1920 ... paper in 1920, which can be formulated as \ddot x - \epsilon ...Analysis · Large Damping · Electrical Circuit · Periodic Forcing and...
  25. [25]
    [PDF] A Bifurcation Analysis of The Forced van der Pol Oscillator
    In this paper we explore the dynamics of a nonlinear damped differential equation, the Forced van der Pol Oscillator. We have done this by constructing.
  26. [26]
    Elements of Physical Biology : Alfred J.Lotka - Internet Archive
    Dec 9, 2006 · 1925. Topics: Biology. Publisher: Williams and Wilkins Company ... PDF download · download 1 file · SINGLE PAGE PROCESSED TIFF ZIP download.
  27. [27]
    [PDF] Mathematical Biology: I. An Introduction, Third Edition
    Consider the basic 2-species Lotka–Volterra competition model with each species ... For each couple we plot the model's null clines in the (Wt ,Ht) phase plane.
  28. [28]
    Alfred J. Lotka and the origins of theoretical population ecology - PMC
    Aug 4, 2015 · The equations describing the predator–prey interaction eventually became known as the “Lotka–Volterra equations,” which served as the ...
  29. [29]
    Symmetry Breaking Instabilities in Dissipative Systems. II
    The thermodynamic theory of symmetry breaking instabilities in dissipative systems is presented. Several kinetic schemes which lead to an unstable behavior ...