Fact-checked by Grok 2 weeks ago

Linear stability

Linear stability analysis is a mathematical technique used to evaluate the local stability of points in nonlinear dynamical systems by approximating the system's near those points through , typically via the , and determining based on the eigenvalues of that : an is if all eigenvalues have negative real parts, unstable if any have positive real parts, and marginally if the real parts are zero. This approach provides insight into whether small perturbations from will decay or grow, serving as a first-order approximation that reveals the qualitative without solving the full nonlinear equations. The process begins by identifying an equilibrium point x^* where the vector field f(x^*) = 0 in a system described by \dot{x} = f(x). The nonlinear system is then linearized around x^* using the first-order Taylor expansion, yielding \dot{\delta x} \approx J(x^*) \delta x, where \delta x = x - x^* is the perturbation and J(x^*) is the Jacobian matrix evaluated at the equilibrium. The eigenvalues \lambda of J(x^*) dictate the stability: the real parts determine the growth or decay rates of perturbations, with complex eigenvalues indicating possible oscillatory modes such as spirals if the imaginary parts are nonzero. For higher-dimensional systems, the dominant eigenvalue (with the largest real part) governs the overall stability, and tools like the Nyquist criterion can be employed for frequency-domain analysis in control applications. This method finds broad applications across disciplines, including fluid mechanics where it predicts the transition from laminar to turbulent flow by assessing perturbations in velocity and pressure fields around base flows, often parameterized by critical values like the Rayleigh number in thermal convection problems. In chemical engineering, it analyzes reactor stability by linearizing mass and energy balance equations to evaluate sensitivity to temperature or concentration fluctuations. Similarly, in biological systems, linear stability helps model the robustness of steady states in population dynamics or biochemical networks, such as determining whether enzyme-substrate equilibria resist small changes in initial conditions. In control theory and plasma physics, it informs the design of feedback systems and the growth of instabilities in confined plasmas, respectively, by quantifying how infinitesimal disturbances evolve under linearized governing equations. While powerful for local analysis, linear stability may fail for nonlinear effects dominating at larger perturbations, necessitating complementary global methods like Lyapunov functions for complete assessment.

Fundamentals

Definition

Linear stability analysis is a fundamental concept in the study of dynamical systems, used to determine whether small perturbations around an point will grow, decay, or remain bounded over time. In linear dynamical systems, such as those governed by equations of the form \dot{x} = Ax in continuous time or x_{n+1} = Ax_n in discrete time, is assessed by examining the eigenvalues of the system matrix A: the at the is asymptotically stable if all eigenvalues have negative real parts (continuous case) or absolute values less than 1 (discrete case), causing perturbations to decay exponentially, while positive real parts or magnitudes at least 1 lead to growth and instability. For nonlinear dynamical systems, linear stability refers to the stability properties of the linearized near an point, where the system's behavior for small deviations is approximated by its first-order expansion. This approach predicts that perturbations will decay if the linearized system is , thereby indicating local asymptotic stability of the nonlinear , or grow if unstable, suggesting local instability. The key distinction lies in scope: linear stability applies directly to inherently linear systems with exact global behavior governed by the matrix eigenvalues, whereas in nonlinear systems, it provides a local criterion via , which may not capture global dynamics or cases where higher-order terms dominate. The origins of linear stability trace back to the late , rooted in for differential equations developed by and . Poincaré's work on , particularly in analyzing the , highlighted the utility of linear approximations to understand qualitative behavior near equilibria through small perturbations. Complementing this, Lyapunov's seminal 1892 dissertation, The General Problem of the of Motion, established rigorous frameworks for stability, including as a tool to classify equilibria based on perturbation responses. This method's primary motivation is its role as a computationally efficient first-order approximation, enabling predictions of long-term solution behavior near equilibria—such as attraction or repulsion—without solving the often intractable full nonlinear equations, thus serving as an essential preliminary step in broader stability investigations.

Equilibrium Points

In dynamical systems, an equilibrium point, also known as a fixed point, is defined as a state where the system's variables remain constant over time, corresponding to a constant solution of the governing equations. For continuous-time systems described by ordinary differential equations of the form \dot{x} = f(x), where x \in \mathbb{R}^n and f: \mathbb{R}^n \to \mathbb{R}^n, an equilibrium point x^* satisfies f(x^*) = 0, meaning the time derivative \dot{x} = 0 at that point. In discrete-time systems, such as iterations x_{n+1} = g(x_n), equilibria are fixed points where x^* = g(x^*). To identify equilibrium points, one solves the f(x) = 0 for continuous systems, often requiring numerical methods or analytical techniques depending on the nonlinearity of f. For systems, fixed points are found by solving x = g(x), which similarly may involve for complex g. These points represent steady states, such as rest positions in mechanical systems or balanced populations in ecological models. Equilibria are classified as hyperbolic if all eigenvalues of the system's Jacobian matrix at x^* have nonzero real parts, or non-hyperbolic otherwise. equilibria permit linear stability because the local behavior near x^* is topologically equivalent to that of the linearized system, as established by the . Non-hyperbolic cases, with eigenvalues on the imaginary axis, require more advanced techniques beyond linearization due to potential center manifolds. Equilibrium points play a central in understanding dynamical systems, as their determines whether they act as attractors (drawing nearby trajectories toward them), repellers (pushing trajectories away), or saddles (with mixed along and unstable directions). For instance, in predator-prey models, a might represent coexistence, while an unstable one indicates risks. This classification via linear reveals the qualitative long-term of trajectories in .

Linearization

Jacobian Matrix

The Jacobian matrix plays a central role in linear stability analysis by providing the linear approximation of the vector field near an equilibrium point. For an autonomous dynamical system \dot{x} = f(x), where x \in \mathbb{R}^n and f: \mathbb{R}^n \to \mathbb{R}^n is a smooth vector field, the Jacobian matrix at an equilibrium x^* (satisfying f(x^*) = 0) is defined as J(x^*) = \left. \frac{\partial f}{\partial x} \right|_{x = x^*}, the n \times n matrix whose (i,j)-th entry is the partial derivative \frac{\partial f_i}{\partial x_j} evaluated at x^*. This matrix encapsulates the local geometry of the flow near x^*. In the scalar case (n=1), where \dot{x} = f(x), the Jacobian simplifies to the ordinary derivative J(x^*) = f'(x^*), a single value representing the slope of the vector field at the equilibrium. For multivariable systems, computation involves evaluating all relevant partial derivatives; for instance, in a two-dimensional system \dot{x}_1 = f_1(x_1, x_2), \dot{x}_2 = f_2(x_1, x_2), the Jacobian is J(x^*) = \begin{pmatrix} \frac{\partial f_1}{\partial x_1} & \frac{\partial f_1}{\partial x_2} \\ \frac{\partial f_2}{\partial x_1} & \frac{\partial f_2}{\partial x_2} \end{pmatrix}_{x = x^*}. These entries are obtained analytically from the functional form of f or numerically via finite differences if explicit expressions are unavailable. The matrix relates to the at the in the , where it acts as the of the , providing a linear transformation that describes infinitesimal displacements from x^*. This first-order response captures how small perturbations \delta x evolve initially according to \dot{\delta x} \approx J(x^*) \delta x, revealing the local dynamics without higher-order nonlinear effects. For time-dependent systems \dot{x} = f(x, t), the J(x, t) = \frac{\partial f}{\partial x}(x, t) varies with time, complicating direct analysis; a frozen evaluates J at a fixed time (e.g., t=0) to assess instantaneous around an . In periodic cases, an average over one period can approximate the effective for purposes.

Local Approximation

The linearization provides a method to approximate the behavior of a nonlinear near an point x^* by a . Consider the autonomous system \dot{x} = f(x), where f: \mathbb{R}^n \to \mathbb{R}^n is continuously differentiable and x^* satisfies f(x^*) = 0. For a small \delta x = x - x^*, the theorem states that \dot{\delta x} \approx J(x^*) \delta x, where J(x^*) is the matrix of f evaluated at x^*. This approximation arises from the Taylor series expansion of f around x^*: f(x^* + \delta x) = f(x^*) + J(x^*) \delta x + O(|\delta x|^2). Since f(x^*) = 0, the first-order term J(x^*) \delta x dominates for sufficiently small \|\delta x\|, as the higher-order terms become negligible. The validity of this local approximation holds under certain conditions on the equilibrium. Specifically, for hyperbolic equilibria—where all eigenvalues of J(x^*) have nonzero real parts—the linear term accurately captures the qualitative dynamics near x^*, as established by the , which guarantees between the nonlinear flow and its . In non-hyperbolic cases, where at least one eigenvalue has zero real part, the higher-order terms may influence the behavior, necessitating nonlinear analysis beyond the local . The resulting approximated system takes the form \dot{y} = A y, with A = J(x^*) and y = \delta x. The solution to this linear system is y(t) = e^{A t} y(0), where e^{A t} is the matrix exponential, describing exponential growth, decay, or oscillation depending on the spectrum of A.

Stability Criteria

Eigenvalue Method

The eigenvalue method assesses the linear stability of an equilibrium point x^* in a continuous-time dynamical system \dot{x} = f(x) by examining the eigenvalues of the Jacobian matrix J(x^*), which arises from the local linear approximation \delta \dot{x} \approx J(x^*) \delta x near the equilibrium. This approach relies on the fact that the qualitative behavior of trajectories near a hyperbolic equilibrium (where no eigenvalue has zero real part) mirrors that of the linearized system, as established by the Hartman-Grobman theorem. The stability criterion is determined by the real parts of the eigenvalues \lambda of J(x^*): the equilibrium is asymptotically if \operatorname{Re}(\lambda) < 0 for all \lambda; unstable if \operatorname{Re}(\lambda) > 0 for any \lambda; and neutrally (or marginally ) if \operatorname{Re}(\lambda) = 0 for all \lambda, though the latter case is rare and typically requires nonlinear for resolution. In the asymptotically case, perturbations decay exponentially to zero; in the unstable case, at least one perturbation grows exponentially./10:_Dynamical_Systems_Analysis/10.04:_Using_eigenvalues_and_eigenvectors_to_find_stability_and_solve_ODEs) The general solution to the linearized system reveals how perturbations \delta x(t) evolve: assuming J(x^*) is diagonalizable with eigenvalues \lambda_k and corresponding eigenvectors v_k, it takes the form \delta x(t) \approx \sum_k c_k v_k e^{\lambda_k t}, where the coefficients c_k depend on initial conditions. The exponential terms e^{\lambda_k t} govern the growth or decay, with the sign of \operatorname{Re}(\lambda_k) dictating whether the k-th mode amplifies or diminishes over time./10:_Dynamical_Systems_Analysis/10.04:_Using_eigenvalues_and_eigenvectors_to_find_stability_and_solve_ODEs) When eigenvalues are , they occur in conjugate pairs \lambda = \alpha \pm i \beta (with \beta \neq 0). If \alpha < 0, the system exhibits damped oscillatory behavior, as the real part \alpha causes exponential decay while the imaginary part \beta introduces rotations in the phase plane, leading to spiraling trajectories toward the . For \alpha > 0, the oscillations grow (unstable spiral); for \alpha = 0, pure oscillations persist without decay or growth. In marginal cases where some eigenvalues are purely imaginary (zero real part), the equilibrium is non-hyperbolic, and linear alone is inconclusive regarding . Such situations often necessitate higher-order techniques, such as reduction, to capture the along the center eigenspace while and unstable manifolds handle the rest.

Characteristic Equation

In the context of linear for continuous-time dynamical systems, the arises from the linearized form of the system around an point. For a system described by \dot{x} = f(x) linearized as \dot{\delta x} = A \delta x, where A is the matrix evaluated at the , the is given by \det(\lambda I - A) = 0. The roots of this equation are the eigenvalues \lambda of A, and the is asymptotically if all eigenvalues have negative real parts, a condition known as Hurwitz . For discrete-time systems, such as iterative maps x_{n+1} = g(x_n) near an , the linearization yields \delta x_{n+1} = J \delta x_n, where J is the matrix at the . The corresponding is \det(\lambda I - J) = 0, with roots being the eigenvalues \lambda of J. Asymptotic stability requires all eigenvalues to satisfy |\lambda| < 1, ensuring that perturbations decay over iterations. This framework extends to higher-order scalar linear differential equations, such as the second-order form \ddot{x} + a \dot{x} + b x = 0. The associated is \lambda^2 + a \lambda + b = 0, and holds if the roots have negative real parts, which by the Routh-Hurwitz criterion requires a > 0 and b > 0. In discrete systems, the requirement that all roots of the lie inside the unit circle defines Schur , which contrasts with the Hurwitz condition for continuous systems by shifting the stability boundary from the imaginary axis to the unit circle in the .

Examples

Ordinary Differential Equations

Linear stability analysis is commonly applied to ordinary differential equations (ODEs) to determine the behavior of solutions near points. For a scalar ODE of the form \dot{x} = f(x), the equilibria occur where f(x^*) = 0, and the involves computing the J = f'(x^*). If J < 0, the equilibrium is asymptotically stable; if J > 0, it is unstable; and if J = 0, the analysis is inconclusive, requiring higher-order terms. Consider the simple nonlinear scalar \dot{x} = -x + x^2. The equilibria are found by setting -x + x^2 = 0, yielding x^* = 0 and x^* = 1. For , compute f'(x) = -1 + 2x. At x^* = 0, J = -1 < 0, indicating asymptotic stability: small perturbations decay exponentially toward the . At x^* = 1, J = 1 > 0, indicating : perturbations grow away from the . This example illustrates how captures the local dynamics dominated by the linear term near each point. For multivariable systems, the Lotka-Volterra predator-prey model provides a concrete illustration. The equations are \dot{x} = a x - b x y for prey population x and \dot{y} = -c y + d x y for predator population y, where a, b, c, d > 0. The equilibria are the trivial point (0, 0) and the coexistence point (c/d, a/b). The Jacobian matrix is J(x, y) = \begin{pmatrix} a - b y & -b x \\ d y & -c + d x \end{pmatrix}. At (0, 0), J = \begin{pmatrix} a & 0 \\ 0 & -c \end{pmatrix}, with eigenvalues a > 0 and -c < 0. The positive and negative eigenvalues indicate a saddle point: unstable in the prey direction but stable in the predator direction. At the coexistence point (c/d, a/b), J = \begin{pmatrix} 0 & -\frac{b c}{d} \\ \frac{d a}{b} & 0 \end{pmatrix}, with eigenvalues \pm i \sqrt{a c}, which are purely imaginary. This results in oscillatory neutral stability, where trajectories form closed orbits around the equilibrium in the phase plane, representing sustained population cycles without damping or growth. In the phase plane for two-dimensional linear systems \dot{\mathbf{z}} = A \mathbf{z}, the eigenvalues of A determine the qualitative behavior near the origin (an equilibrium). If both eigenvalues are real and negative, trajectories approach along straight lines or curves forming a stable node. If both are real and positive, an unstable node repels trajectories outward. For complex conjugate eigenvalues \alpha \pm i \beta with \alpha < 0, trajectories spiral inward in a stable spiral (or sink); if \alpha > 0, they spiral outward in an unstable spiral (or source). These portraits visualize how linear stability criteria manifest geometrically, aiding intuition for nonlinear approximations. Despite its utility, linear stability analysis has limitations when nonlinear terms dominate, particularly if all eigenvalues have non-positive real parts but some are zero. For the scalar \dot{x} = x^3, the is at x^* = 0, and f'(0) = 0, rendering linearization inconclusive. Higher-order analysis reveals : solutions satisfy x(t) = \frac{x_0}{ (1 - 2 x_0^2 t)^{1/2} } for x_0 \neq 0, which escape to in finite time, showing that the nonlinear cubic term drives despite the neutral linear prediction. Such cases necessitate Lyapunov functions or theory for resolution.

Partial Differential Equations

Linear stability analysis for partial differential equations (PDEs) extends the concepts from ordinary differential equations to infinite-dimensional systems, where spatial variations play a crucial role. Consider a general evolutionary PDE of the form \partial_t u = F(u, \nabla u, \Delta u, \dots), where u(x,t) is the state variable on a spatial domain, and F is a nonlinear operator incorporating reaction and diffusion terms. An equilibrium solution u^* satisfies F(u^*, \nabla u^*, \dots) = 0. To assess its stability, one linearizes around u^* by setting u = u^* + \epsilon v, where \epsilon is small and v is the perturbation. Substituting and expanding to first order yields the linearized equation \partial_t v = L v, where L = \frac{\delta F}{\delta u} \big|_{u^*} is the Fréchet derivative of F at u^*, acting as an unbounded linear operator on a suitable function space such as L^2 or Sobolev spaces. The of the is determined by the of the L. If all eigenvalues of L have negative real parts (or more generally, if the lies in the left half-plane and satisfies resolvent conditions for sectorial operators), small perturbations exponentially, indicating asymptotic in the . This spectral criterion generalizes the eigenvalue method for finite-dimensional systems, but requires careful consideration of the domain and boundary conditions due to the infinite-dimensional nature. For spatially periodic or infinite domains, techniques decompose perturbations into modes v_k(x,t) = \hat{v}_k(t) e^{i k \cdot x}, leading to ordinary differential equations for each mode's amplitude, whose growth rates are the eigenvalues. A canonical example arises in reaction-diffusion systems, \partial_t u = D \nabla^2 u + f(u), where D > 0 is the diffusion coefficient and f is a nonlinear reaction term, such as f(u) = u(1 - u) for logistic growth. At a constant equilibrium u^* where f(u^*) = 0, the linearized operator is L = D \nabla^2 + f'(u^*). Applying Fourier modes e^{i k \cdot x} yields the dispersion relation \lambda(k) = -D |k|^2 + f'(u^*), where k is the wave number. The uniform state is stable if \operatorname{Re} \lambda(k) < 0 for all k, but Turing instability occurs when diffusion destabilizes spatial patterns: if f'(u^*) > 0 (unstable without diffusion) yet \lambda(k) > 0 for some finite k \neq 0 but \lambda(0) < 0, finite-wavelength perturbations grow, leading to pattern formation like spots or stripes. This mechanism, first proposed by Turing, requires at least two interacting species with differing diffusivities for realistic chemical or biological systems. In numerical simulations, spatial discretization via finite differences approximates the continuous L by a large , whose eigenvalues provide an estimate of the PDE's . For instance, central differences for the Laplacian on a uniform grid yield a in periodic settings, with eigenvalues closely matching the -D (2\pi m / N h)^2 for grid size N and spacing h. The further assesses time-stepping schemes by examining the amplification factor of modes, ensuring that discrete eigenvalues remain in the stable half-plane to avoid spurious instabilities. This approach confirms that well-resolved s preserve the qualitative properties of the underlying PDE. As a specific example, consider the Fisher-KPP \partial_t u = \partial_{xx} u + u(1 - u) on the real line, modeling spread. While constant equilibria u=0 (unstable) and u=1 (stable) are analyzed via the above, traveling wave solutions u(x,t) = U(x - c t) connect them, with minimum speed c \geq 2. Linear stability of these pulled fronts—where the wave speed is determined by the at the u \approx 0—involves analyzing perturbations around the front. The linearized operator at the unstable state yields a \lambda(\omega) = - \omega^2 - i c \omega + 1, and marginal stability implies the front selects the speed where the leading eigenvalue has zero real part for a double root, ensuring algebraic convergence of perturbations rather than . This linear analysis at the front explains the universal logarithmic delay in front position observed in simulations and experiments.

References

  1. [1]
    7.5: Linear Stability Analysis of Nonlinear Dynamical Systems
    Apr 30, 2024 · Linear stability analysis of continuous-time nonlinear systems. 1. Find an equilibrium point of the system you are interested in.
  2. [2]
    Linear Stability Analysis - an overview | ScienceDirect Topics
    Linear stability analysis is defined as a method used to assess the sensitivity of a flow to infinitesimal perturbations by linearizing the governing equations ...
  3. [3]
    9b. Linear stability analysis — Biological Circuit Design documentation
    The main idea behind linear stability analysis is to locally approximate a nonlinear dynamical system by its Taylor series to first order near the fixed point.
  4. [4]
    Linear stability analysis | Plasma Physics Class Notes - Fiveable
    Linear stability analysis is crucial for understanding plasma behavior. It examines how small disturbances affect equilibrium states, determining if they grow ...
  5. [5]
    [PDF] Unit 22: Stability
    Lecture. 22.1. A linear dynamical system is either a discrete time dynamical system x(t + 1) = Ax(t) or a continuous time dynamical systems x0(t) = Ax(t).
  6. [6]
    [PDF] 1.4 Stability and Linearization
    Since questions of stability are central in dynamical systems, we will want to define the concept of stability precisely and develop criteria for.
  7. [7]
  8. [8]
    The general problem of the stability of motion - Taylor & Francis Online
    (1992). The general problem of the stability of motion. International Journal of Control: Vol. 55, No. 3, pp. 531-534.
  9. [9]
    Equilibrium - Scholarpedia
    Oct 21, 2011 · An equilibrium is asymptotically stable if all eigenvalues have negative real parts; it is unstable if at least one eigenvalue has positive real part.Jacobian Matrix · Types of Equilibria · Two-Dimensional Space
  10. [10]
    [PDF] Stability and Performance
    Equilibrium points are one of the most important features of a dynami- cal system since they define the states corresponding to constant operating conditions. ...
  11. [11]
    Equilibria in discrete dynamical systems - Math Insight
    An equilibrium is the simplest possible solution to a dynamical system. It is a solution where the state variable is a constant; the variable doesn't change ...
  12. [12]
    [PDF] Stability of Dynamical systems - math.utah.edu
    Unstable equilibrium. The equilibrium point u0 is called unstable provided it is not stable, which means (a) or (b) fails (or both).
  13. [13]
    [PDF] 19 Jacobian Linearizations, equilibrium points
    So, a question arises: “In what limited sense can a nonlinear system be viewed as a linear system?” In this section we develop what is called a “Jacobian ...
  14. [14]
    [PDF] Stability Analysis of Nonlinear Systems using Frozen Stationary ...
    Apr 2, 2003 · In this paper is discussed how to compute stability regions for nonlinear systems with slowly varying parameters using frozen stationary ...
  15. [15]
    [PDF] Nonlinear Dynamics and Chaos
    May 6, 2020 · This book covers nonlinear dynamics and chaos, including chaos, fractals, and dynamics, with applications to physics, biology, chemistry, and ...
  16. [16]
    None
    ### Summary of Hartman-Grobman Theorem for Flows (ODEs)
  17. [17]
  18. [18]
    Stability of equilibria - Scholarpedia
    Mar 15, 2007 · ... purely imaginary eigenvalues that determine the stability of a critical equilibrium. The Center Manifold Theorem allows one to reduce the ...
  19. [19]
    [PDF] 1 Stability of a linear system - Princeton University
    Mar 24, 2016 · A standard result in linear algebra tells us that the origin of the system xk+1 = Axk is GAS if and only if all eigenvalues of A have norm ...
  20. [20]
    [PDF] The Behavior of Dynamic Systems - MIT OpenCourseWare
    3.3 Stability. We test for stability by simply generating the characteristic equations and looking at the roots. Where does the characteristic equation come ...Missing: continuous | Show results with:continuous
  21. [21]
    5.7: 5.7 Linear Stability Analysis of Discrete-Time Nonlinear ...
    Apr 30, 2024 · 1. Find an equilibrium point of the system you are interested in. 2. Calculate the Jacobian matrix of the system at the equilibrium point. 3.
  22. [22]
    [PDF] ECE504: Lecture 8 - spinlab
    Theorem. A continuous-time LTI system is (internally) stable if and only if both of the following conditions are true. 1. Re(λj) ≤ 0 for all j ∈ {1,...,s} ...
  23. [23]
    [PDF] 18.03SCF11 text: Stability - MIT OpenCourseWare
    coefficient form. Assume a0 > 0. a0y. . + a1y. . + a2y = r(t) is stable a0, a1, ...
  24. [24]
    New Stability Criteria for Discrete Linear Systems Based on ... - MDPI
    On the other hand, a discrete linear system is stable if and only if its characteristic polynomial is a Schur polynomial, i.e., all of its zeros are located in ...
  25. [25]
    The Schur stability via the Hurwitz stability analysis using a ...
    In the design of discrete-time control systems, the Schur stability of a discrete-time polynomial should be insured. There are some methods to check the Schur ...
  26. [26]
  27. [27]
    Differential Equations - Phase Plane - Pauls Online Math Notes
    Nov 16, 2022 · In an asymptotically stable node or spiral all the trajectories will move in towards the equilibrium point as t increases, whereas a center ( ...
  28. [28]
    [PDF] A brief introduction to stability theory for linear PDEs
    Jun 5, 2012 · In particular, we will show that for reaction-diffusion equations, linear stability can be determined simply by computing the spectrum of the ...
  29. [29]
    From finite differences to finite elements: A short history of numerical ...
    This is an account of the history of numerical analysis of partial differential equations, starting with the 1928 paper of Courant, Friedrichs, and Lewy.