Fact-checked by Grok 2 weeks ago

Hyperbolic equilibrium point

In the study of dynamical systems, a equilibrium point, also known as a hyperbolic fixed point, is an equilibrium solution of a where the ( ) at that point has no eigenvalues that are purely imaginary, meaning all eigenvalues have non-zero real parts. This condition ensures that the local behavior near the equilibrium is dominated by expansion or contraction in certain directions, without neutral or oscillatory modes on the imaginary axis. Hyperbolic equilibria are classified into three main types based on the signs of the real parts of the eigenvalues: sinks (all negative real parts, attracting trajectories), sources (all positive real parts, repelling trajectories), and saddles (mixed signs, with both attracting and repelling directions). A key property is the splitting of the tangent space at the equilibrium into stable and unstable eigenspaces, corresponding to directions of contraction and expansion, respectively. The Hartman–Grobman theorem guarantees that, near such a point, the nonlinear flow is topologically conjugate to the linear flow, allowing the local phase portrait to be approximated by the linearized system. Associated with hyperbolic equilibria are the W^s, consisting of points whose trajectories approach the equilibrium as time t \to +\infty, and the unstable manifold W^u, comprising points that approach as t \to -\infty. These manifolds are invariant under the , smooth, and tangent to the respective eigenspaces at the . The stable manifold theorem asserts the existence and uniqueness of these local manifolds for hyperbolic equilibria in finite-dimensional systems, which can extend globally under certain conditions. Hyperbolicity provides robustness to perturbations, making these points structurally stable in generic systems.

Discrete Dynamical Systems

Definition

In discrete dynamical systems, the dynamics are governed by an iteration of a smooth map F: \mathbb{R}^n \to \mathbb{R}^n, where the state evolves as x_{n+1} = F(x_n) for n \in \mathbb{Z}. The solutions define an \{x_n\} starting from an x_0 = x, with forward iterations x_n = F^n(x) for n > 0 and backward iterations x_n = F^{-n}(x) assuming F is invertible. This discrete-time evolution contrasts with the continuous flows of differential equations in other dynamical contexts. A fixed point x^* of the map is an where F(x^*) = x^*, ensuring that orbits starting at x^* remain there for all iterations, i.e., F^n(x^*) = x^* for all n \in \mathbb{Z}. Such points represent stationary states under the mapping. A fixed point x^* is defined as one where the matrix DF(x^*) has no eigenvalues with modulus one, meaning |\lambda| \neq 1 for every eigenvalue \lambda of DF(x^*). This condition excludes neutral directions on the unit circle, which would indicate or resonances in the linearized approximation near x^*.

Linearization and Eigenvalues

In discrete dynamical systems governed by smooth maps F: \mathbb{R}^n \to \mathbb{R}^n, a fixed point x^* satisfies F(x^*) = x^*. To analyze the local behavior near x^*, the system is linearized by shifting variables to y = x - x^*, yielding y_{n+1} = DF(x^*) y_n + o(\|y_n\|) as y_n \to 0, where DF(x^*) is the Jacobian matrix of F at x^*. The solutions to this linearized system are given by y_n = [DF(x^*)]^n y_0, which approximate the nonlinear iterations near the fixed point for small initial perturbations. A fixed point x^* is hyperbolic if all eigenvalues \lambda_i of the DF(x^*) satisfy |\lambda_i| \neq 1, excluding any neutral directions on the unit circle. This spectral condition ensures that the linearization captures the essential qualitative without neutral modes that would require higher-order . The eigenspaces of DF(x^*) decompose \mathbb{R}^n into the direct sum of the stable subspace E^s and the unstable subspace E^u, where E^s is the (generalized) eigenspace corresponding to eigenvalues with |\lambda_i| < 1, and E^u for those with |\lambda_i| > 1. Thus, \mathbb{R}^n = E^s \oplus [E^u](/page/Direct_sum), providing a linear splitting that governs the local . In the hyperbolic case, iterations starting in E^s approach the fixed point exponentially as n \to +\infty, with decay rates determined by the eigenvalues inside disk, while those in E^u diverge exponentially as n \to +\infty. Conversely, iterations in E^u approach the exponentially as n \to -\infty, highlighting the time-reversible nature of the near fixed points. For a concrete illustration, consider the x_{n+1} = A x_n in \mathbb{R}^2, where A = \begin{pmatrix} 2 & 0 \\ 0 & \frac{1}{2} \end{pmatrix}. The eigenvalues of A are \lambda_1 = 2 > 1 and \lambda_2 = \frac{1}{2} < 1, both with moduli not equal to 1, confirming hyperbolicity. Here, E^u = \operatorname{span}\{(1,0)\} corresponds to exponential growth along the x-axis, while E^s = \operatorname{span}\{(0,1)\} yields exponential decay along the y-axis, decomposing \mathbb{R}^2 = E^s \oplus E^u. The explicit solution is x(n) = (2^n x_1(0), (\frac{1}{2})^n x_2(0)), demonstrating the distinct behaviors in each subspace.

Continuous Dynamical Systems

Definition

In continuous dynamical systems, the dynamics are governed by an ordinary differential equation (ODE) of the form \dot{x} = f(x), where x \in \mathbb{R}^n and f: \mathbb{R}^n \to \mathbb{R}^n is a sufficiently smooth vector field. The solutions to this ODE define a flow \phi_t(x), which maps an initial condition x(0) = x to the position x(t) at time t, satisfying \frac{d}{dt} \phi_t(x) = f(\phi_t(x)) with \phi_0(x) = x. This flow provides a continuous-time evolution of the system, contrasting with the discrete iterations of maps in other dynamical contexts. An equilibrium point x^* of the flow is a fixed point where f(x^*) = 0, ensuring that trajectories starting at x^* remain there for all time, i.e., \phi_t(x^*) = x^* for all t \in \mathbb{R}. Such points represent stationary states in the system, where the vector field vanishes. A hyperbolic equilibrium point x^* is defined as one where the Jacobian matrix Df(x^*) has no eigenvalues with zero real part, meaning \Re(\lambda) \neq 0 for every eigenvalue \lambda of Df(x^*). This condition excludes cases with purely imaginary eigenvalues or zero eigenvalues, which would indicate marginal stability or resonance in the linearized approximation near x^*.

Linearization and Eigenvalues

In continuous dynamical systems governed by ordinary differential equations (ODEs) of the form \dot{x} = f(x), where x \in \mathbb{R}^n and f is sufficiently smooth, an equilibrium point x^* satisfies f(x^*) = 0. To analyze the local behavior near x^*, the system is linearized by shifting variables to y = x - x^*, yielding \dot{y} = Df(x^*) y + o(\|y\|) as y \to 0, where Df(x^*) is the of f at x^*. The solutions to this linearized system are given by y(t) = e^{Df(x^*) t} y(0), which approximate the nonlinear flow near the equilibrium for small initial perturbations. An equilibrium x^* is hyperbolic if all eigenvalues \lambda_i of the Jacobian Df(x^*) satisfy \Re(\lambda_i) \neq 0, excluding any center directions where \Re(\lambda_i) = 0. This spectral condition ensures that the linearization captures the essential qualitative dynamics without neutral modes that would require higher-order analysis, such as . The eigenspaces of Df(x^*) decompose \mathbb{R}^n into the direct sum of the stable subspace E^s and the unstable subspace E^u, where E^s is the (generalized) eigenspace spanned by eigenvectors corresponding to eigenvalues with \Re(\lambda_i) < 0, and E^u for those with \Re(\lambda_i) > 0. Thus, \mathbb{R}^n = E^s \oplus E^u, providing a linear splitting that governs the local . In the hyperbolic case, trajectories starting in E^s approach the exponentially as t \to \infty, with decay rates determined by the negative real parts of the eigenvalues, while those in E^u diverge exponentially as t \to \infty. Conversely, solutions in E^u approach the origin exponentially as t \to -\infty, highlighting the time-reversible nature of the linear near hyperbolic equilibria. For a concrete illustration, consider the \dot{x} = A x in \mathbb{R}^2, where A = \begin{pmatrix} 1 & 0 \\ 0 & -2 \end{pmatrix}. The eigenvalues of A are \lambda_1 = 1 > 0 and \lambda_2 = -2 < 0, both with non-zero real parts, confirming hyperbolicity. Here, E^u = \operatorname{span}\{(1,0)\} corresponds to exponential growth along the x-axis, while E^s = \operatorname{span}\{(0,1)\} yields exponential decay along the y-axis, decomposing \mathbb{R}^2 = E^s \oplus E^u. The explicit solution is x(t) = (e^{t} x_1(0), e^{-2t} x_2(0)), demonstrating the distinct behaviors in each subspace.

Properties

Stable and Unstable Manifolds

In the context of hyperbolic equilibrium points, the stable and unstable manifolds provide essential geometric insight into the local behavior of trajectories near the equilibrium. For a continuous dynamical system defined by an autonomous vector field \dot{x} = f(x) with a hyperbolic equilibrium at x^*, where f(x^*) = 0 and the Jacobian Df(x^*) has no eigenvalues with zero real part, the local stable manifold W^s(x^*) is defined as the set of points x such that \lim_{t \to \infty} \phi_t(x) = x^*, where \phi_t denotes the flow of the system. Similarly, the local unstable manifold W^u(x^*) consists of points x satisfying \lim_{t \to -\infty} \phi_t(x) = x^*. These manifolds are immersed submanifolds of the phase space, with dimensions equal to the dimensions of the stable and unstable eigenspaces of Df(x^*), respectively. For discrete dynamical systems given by a diffeomorphism f with a hyperbolic fixed point x^*, where Df(x^*) has no eigenvalues on the unit circle, the local stable manifold W^s(x^*) is the set \{ x : \lim_{n \to \infty} f^n(x) = x^* \}, and the local unstable manifold W^u(x^*) is \{ x : \lim_{n \to -\infty} f^n(x) = x^* \}. These definitions align with the continuous case in capturing the directions of attraction and repulsion near the equilibrium. The manifolds satisfy specific tangency conditions at the equilibrium: the tangent space T_{x^*} W^s(x^*) coincides with the stable eigenspace E^s of the linearization, and T_{x^*} W^u(x^*) = E^u, the unstable eigenspace. This ensures that the nonlinear behavior locally mirrors the linear approximation along these directions. Both manifolds are invariant under the dynamics. In the continuous setting, the flow preserves the stable manifold forward in time, so \phi_t(W^s(x^*)) \subset W^s(x^*) for t \geq 0, and the unstable manifold backward in time, \phi_t(W^u(x^*)) \subset W^u(x^*) for t \leq 0. For discrete systems, f(W^s(x^*)) \subset W^s(x^*) and f(W^u(x^*)) \supset W^u(x^*), with the inverse map acting oppositely on the unstable manifold. The stable manifold theorem guarantees the uniqueness of these local manifolds among all invariant submanifolds tangent to the respective eigenspaces at the hyperbolic equilibrium. This uniqueness holds for both flows and maps, ensuring that the geometric structures are uniquely determined by the hyperbolic linearization.

Local Invariant Sets

In hyperbolic dynamical systems, homoclinic points arise as the intersections of the stable and unstable manifolds of a x^*, specifically at points in W^s(x^*) \cap W^u(x^*) \setminus \{x^*\}. These points mark the locations where trajectories approach x^* both in forward and backward time, forming homoclinic orbits whose closures constitute compact invariant hyperbolic sets. Such structures introduce intricate local dynamics, as the tangling of manifolds near these points generates a dense set of periodic orbits and other homoclinic points, complicating the neighborhood of x^*. Heteroclinic orbits, in contrast, connect distinct hyperbolic equilibria x^* and y^*, comprising trajectories that lie in W^u(x^*) \cap W^s(y^*) and approach x^* as time tends to -\infty and y^* as time tends to +\infty. The closure of a heteroclinic orbit, including the equilibria, forms another hyperbolic invariant set, often serving as a building block for chains or cycles that link multiple saddles. These orbits contribute to the formation of local invariant structures, such as heteroclinic networks, which exhibit recurrent behavior and can embed symbolic dynamics akin to subshifts. In higher-dimensional systems, the presence of homoclinic and heteroclinic orbits near hyperbolic equilibria can produce invariant Cantor sets, as exemplified by the Smale horseshoe map where manifold intersections create a hyperbolic Cantor set supporting chaotic dynamics. These sets are totally disconnected, compact, and invariant, with the dynamics thereon conjugate to a full shift on two symbols, yielding positive topological entropy. Extending to continuous flows, such tangles may underpin strange attractors, where the local structure around the equilibrium features sensitive dependence and dense orbits within bounded regions. For generic hyperbolic systems, transversality conditions ensure the robustness of these intersections: the tangent spaces at a homoclinic point satisfy T_q W^s(x^*) \cap T_q W^u(x^*) = \mathbb{R} \dot{h}(q) along the orbit h, where the intersection is one-dimensional and spanned by the velocity vector. This transversality, prevalent under C^1-generic perturbations, implies that homoclinic points accumulate densely in the local neighborhood, fostering ergodic behavior and the shadowing of pseudo-orbits by true trajectories. Similarly, transverse heteroclinic connections persist, guaranteeing the structural stability of the associated invariant sets.

Theorems and Results

Hartman-Grobman Theorem

The Hartman–Grobman theorem establishes that, near a of a dynamical system, the nonlinear flow or map is topologically conjugate to its linear approximation. For continuous dynamical systems, consider a C^1 vector field f: U \to \mathbb{R}^n on an open set U \subset \mathbb{R}^n with a hyperbolic equilibrium at x^* \in U, meaning f(x^*) = 0 and the Jacobian A = Df(x^*) has no eigenvalues with zero real part. Let \phi_t denote the flow generated by \dot{x} = f(x), and let \psi_t(x) = e^{A t} x be the linear flow of the system \dot{x} = A x. The theorem asserts that there exist neighborhoods U' of x^* and V of $0, and a homeomorphism h: U' \to V such that h \circ \phi_t = \psi_t \circ h for all x \in U' and |t| \leq T for some T > 0. For discrete dynamical systems, the result applies analogously to C^1 maps. Let f: U \to \mathbb{R}^n have a fixed point at x^*, so f(x^*) = x^* and the A = Df(x^*) has no eigenvalues on the unit circle. Then there exist neighborhoods U' of x^* and V of $0, and a h: U' \to V such that h \circ f = A \circ h on U', or equivalently, h \circ f^k = A^k \circ h for iterates k. The requires the or to be at least continuously differentiable (C^1) to ensure the is well-defined, and hyperbolicity at the to guarantee the exhibits no directions. The conjugacy holds in sufficiently small neighborhoods where nonlinear terms are dominated by the linear ones, but the size of these neighborhoods depends on the spectrum of A. Proofs of the typically proceed by first establishing the result for the time-one of the (or directly for maps) using fixed-point arguments in appropriate s. One common approach decomposes the into and unstable eigenspaces of A, then constructs the as a transform over these manifolds, solving a equation to "straighten" the nonlinear orbits into linear ones via a h = \mathrm{id} + v where v is small. For flows, the conjugacy for all small times follows by integrating the time-one result. The implies that the qualitative behavior of orbits near a equilibrium—such as to sinks, repulsion from sources, or hyperbolic trajectories in saddles—is preserved under to the linear case, allowing the use of eigenvalue analysis to determine local stability types without solving the full nonlinear .

In , a possessing a equilibrium point is considered structurally stable if small perturbations in the C^1 topology result in a topologically conjugate featuring nearby hyperbolic equilibria that preserve the original qualitative dynamics. This property ensures that the global remains unchanged under minor modifications to the or map, highlighting the robustness of hyperbolic structures. The of such systems is intrinsically linked to the hyperbolicity condition, where no eigenvalues of the linearized system lie on the imaginary axis; this guarantees persistence through the Hartman-Grobman theorem, which establishes local topological equivalence to the , combined with the enduring nature of and unstable manifolds under perturbation. Consequently, dynamical systems composed solely of equilibria exhibit overall , as perturbations cannot alter their topological type without violating hyperbolicity. In contrast, non-hyperbolic equilibria, particularly those with eigenvalues having zero real parts, introduce , as small perturbations can trigger such as the Andronov-Hopf bifurcation, where a stable gives way to a or vice versa. This bifurcation underscores the fragility of non-hyperbolic points, as parameter variations or perturbations shift the eigenvalues across the imaginary axis, fundamentally changing the system's structure and precluding . Historically, the role of hyperbolicity in ensuring structural stability was pivotal in Stephen Smale's development of Morse-Smale systems during the 1960s, where all equilibria and periodic orbits are required to be hyperbolic to achieve global structural stability on manifolds. This framework extended earlier work, emphasizing that only systems free of non-hyperbolic elements satisfy the density and openness criteria for structural stability, particularly on low-dimensional manifolds as formalized by Peixoto's theorem.

Examples

Two-Dimensional Maps

In two-dimensional discrete dynamical systems, a canonical example of a hyperbolic fixed point is the origin in the \begin{pmatrix} x_{n+1} \\ y_{n+1} \end{pmatrix} = \begin{pmatrix} 0.5 & 0 \\ 0 & 2 \end{pmatrix} \begin{pmatrix} x_n \\ y_n \end{pmatrix}. The matrix at this fixed point has eigenvalues \lambda_1 = 0.5 and \lambda_2 = 2, satisfying |\lambda_1| < 1 < |\lambda_2|, confirming its hyperbolic nature. The corresponding eigenvectors are (1, 0) for \lambda_1 and (0, 1) for \lambda_2, defining the stable manifold along the x-axis and the unstable manifold along the y-axis. Near the origin, trajectories contract exponentially toward the origin in the x-direction but expand exponentially away in the y-direction, producing saddle-like local dynamics where most points diverge unless initialized precisely on the stable manifold. A prominent nonlinear example is the Hénon map, x_{n+1} = 1 - a x_n^2 + y_n, \quad y_{n+1} = b x_n, originally proposed to model chaotic behavior in hydrodynamical systems. For the classical parameters a = 1.4 and b = 0.3, the map possesses two fixed points at approximately (0.632, 0.190) and (-1.132, -0.340). Both are hyperbolic, as the Jacobian eigenvalues at the first point are approximately $0.156 and -1.92 (with magnitudes $0.156 < 1 < 1.92), and at the second point approximately -0.092 and $3.26 (with magnitudes $0.092 < 1 < 3.26). The stable and unstable directions at each point align with the respective eigenspaces, leading to local contraction along one eigendirection and expansion along the other, though the quadratic nonlinearity introduces global folding and stretching that contributes to the map's strange attractor.

Two-Dimensional Flows

In two-dimensional continuous dynamical systems, governed by autonomous vector fields of the form \dot{\mathbf{x}} = \mathbf{f}(\mathbf{x}) where \mathbf{x} \in \mathbb{R}^2, a hyperbolic equilibrium point \mathbf{x}^* satisfies \mathbf{f}(\mathbf{x}^*) = \mathbf{0} and the Jacobian matrix D\mathbf{f}(\mathbf{x}^*) has eigenvalues with non-zero real parts. This excludes centers (purely imaginary eigenvalues) and ensures the local phase portrait is topologically equivalent to that of the linearized system, per the Hartman-Grobman theorem. Hyperbolic equilibria in 2D flows are classified as sinks (both eigenvalues have negative real parts, attracting trajectories), sources (both positive real parts, repelling), or saddles (one positive and one negative real part, with mixed stability). A canonical linear example is the saddle point at the origin for the system \dot{x} = x, \dot{y} = -y. The Jacobian is \begin{pmatrix} 1 & 0 \\ 0 & -1 \end{pmatrix}, with eigenvalues \lambda_1 = 1 > 0 and \lambda_2 = -1 < 0. The unstable manifold aligns with the x-axis (trajectories diverge as t \to \infty), while the stable manifold aligns with the y-axis (trajectories converge as t \to \infty). For a sink, consider \dot{x} = -x, \dot{y} = -y; the Jacobian \begin{pmatrix} -1 & 0 \\ 0 & -1 \end{pmatrix} yields \lambda_1 = \lambda_2 = -1 < 0, making the origin globally attracting with the entire plane as its basin. Nonlinear examples illustrate how hyperbolicity persists beyond linearization. In the system \dot{x} = -x + x^3, \dot{y} = -2y (from Strogatz), the origin is a sink with Jacobian \begin{pmatrix} -1 & 0 \\ 0 & -2 \end{pmatrix} at (0,0), eigenvalues -1 and -2 (both negative), and trace p = -3 < 0, q = 2 > 0. Nearby points at (\pm 1, 0) are saddles, each with eigenvalues $2 and -2 (opposite signs), trace p = 0, and q = -4 < 0, featuring local stable and unstable manifolds tangent to the eigenspaces. The Lotka-Volterra competition model provides a biological context: \dot{x_1} = x_1(3 - x_1 - 2x_2), \dot{x_2} = x_2(2 - x_1 - x_2), with coexistence equilibrium at (1,1). The Jacobian \begin{pmatrix} -1 & -2 \\ -1 & -1 \end{pmatrix} has eigenvalues -1 - \sqrt{2} < 0 and -1 + \sqrt{2} > 0, confirming a hyperbolic saddle; the stable eigenspace is spanned by [\sqrt{2}, 1]^T and unstable by [-\sqrt{2}, 1]^T. These examples highlight how hyperbolic equilibria structure phase portraits, with invariant manifolds dictating long-term behavior in 2D flows.

References

  1. [1]
    [PDF] HYPERBOLIC DYNAMICAL SYSTEMS Glossary 1 Definition ... - IMPA
    The general philosophy is that the behavior of the system close to a hyperbolic fixed point very much resembles the dynamics of its linear part.
  2. [2]
  3. [3]
    [PDF] Dynamical systems and ODEs - UC Davis Mathematics
    Thus, for a hyperbolic equilibrium, all solutions of the linearized system grow or decay exponentially in time.
  4. [4]
    [PDF] global stable and unstable manifolds
    We again assume the linearized system has no eigenvalues with real part equal to 0. These equilibrium points are called hyperbolic equilibrium points.
  5. [5]
    [PDF] MATH 552 (2023W1) Lecture 8: Mon Sep 25
    An equilibrium p0 is called hyperbolic if the linearization (2.5) is hy- perbolic, i.e. if Re λj 6= 0 for all eigenvalues λj of the constant real n × n. Page 2 ...
  6. [6]
    [PDF] Introduction to Dynamical Systems John K. Hunter - UC Davis Math
    or, at least, remain bounded — or grow. (See ...
  7. [7]
    [PDF] 3 Linearisation and equilibria
    Definition 3.3 (Linear stability). A stationary point x∗ of an autonomous ODE is linearly stable iff the real parts of every eigenvalue of Df(x∗) is negative.
  8. [8]
    [PDF] Linearization and Invariant Manifolds
    Linearization is obtained by neglecting higher order terms in a Taylor series expansion of a system's vector field about an equilibrium point.
  9. [9]
    [PDF] 1 Stability of equilibrium points by lineariza- tion.
    Definition. An equilibrium point x* of the system (9) is called hyperbolic if for all eigenvalues 2 (A) it is valid that Re 6= 0.
  10. [10]
    [PDF] Part II Dynamical Systems Michaelmas Term 2014 - lecturer - DAMTP
    Oct 9, 2014 · Definition 12 (Hyperbolic fixed point). A fixed point x of a dynamical system is hyper- bolic iff all the eigenvalues of the linearization A of ...
  11. [11]
    Differential Equations and Dynamical Systems - SpringerLink
    Book Title: Differential Equations and Dynamical Systems · Authors: Lawrence Perko · Series Title: Texts in Applied Mathematics · Publisher: Springer New York, NY.
  12. [12]
    A First Course in Dynamics
    A first course in dynamics with a panorama of recent developments. Search within full text. Access. Boris Hasselblatt, Tufts University, Massachusetts.
  13. [13]
    None
    ### Summary of Homoclinic Orbits, Heteroclinic Orbits, and Local Invariant Sets
  14. [14]
    [PDF] Homoclinic and Heteroclinic Bifurcations in vector fields | UvA-DARE ...
    Jun 17, 2010 · Bifurcations of homoclinic orbits from equilibria in local bifurcations are also considered. The main analytic and geometric techniques such as ...
  15. [15]
    [PDF] The Hartman-Grobman Theorem - University of Utah Math Dept.
    Oct 15, 2012 · Proof of the Hartman Grobman Theorem. The proof is to show that the time-one maps are locally topologically conjugate and then recover the ...
  16. [16]
    None
    ### Summary of Hartman-Grobman Theorem from https://www.math.unl.edu/~bdeng1/Teaching/Lecture%20Notes/hartman_grobman.pdf
  17. [17]
    [PDF] Structural stability and bifurcations
    Bifurcations of hyperbolic equilibrium points are not possible. Nonhyperbolicity of xE is necessary condition for the occurrence of local bifurcations and the ...Missing: non- | Show results with:non-
  18. [18]
    [PDF] Math 312 Lecture Notes Linear Two-dimensional Maps
    (Both the examples shown in Figures 1 and 2 are saddles.) A saddle is unstable, because there are trajectories beginning arbitrarily close of 0 that diverge.
  19. [19]
  20. [20]
    Simple Chaos - The Hénon Map - American Mathematical Society
    I'll start by looking at the case that Hénon did, that with a=1.4 and b=0.3. ... More precisely the eigenvalues at the first fixed point are -4.9806 and ...
  21. [21]
    Nonlinear Dynamics: Two Dimensional Flows and Phase Diagrams
    Fixed points with a J matrix such that Re(µ1, 2) ≠ 0 are called hyperbolic fixed points. Otherwise, fixed points are non-hyperbolic fixed points, whose ...
  22. [22]
    [PDF] AM 114/214 Prof. Daniele Venturi Stability analysis of equilibria in ...
    In this section we pro- vide a few examples of stability analysis of a hyperbolic fixed point in two-dimensional nonlinear dynamical systems. 3A manifold can be ...