Fact-checked by Grok 2 weeks ago

Stable manifold

In , the stable manifold of an point or hyperbolic fixed point is defined as the set of all initial conditions whose trajectories asymptotically approach that point as time progresses to positive infinity. This concept is central to understanding the local and global structure of s near invariant sets, where trajectories on the stable manifold exhibit exponential convergence to the target point. The existence and properties of stable manifolds are rigorously established by the stable manifold theorem, which asserts that for a hyperbolic fixed point of a sufficiently smooth or , there exists a local stable manifold that is a smooth tangent to the stable eigenspace of the linearized system at that point. This theorem, applicable to both discrete and continuous dynamical systems, guarantees that the manifold is invariant under the dynamics and has the same smoothness as the defining or . Stable manifolds play a crucial role in analyzing chaotic behavior, homoclinic tangles, and , such as in fluid mixing or , where they form boundaries separating basins of attraction despite having measure zero in the .

Introduction

Intuitive description

In dynamical systems, the stable manifold associated with a fixed point captures the collection of trajectories that converge toward that point as time progresses forward, effectively defining the "basin of attraction" in a qualitative . This contrasts with the unstable manifold, where trajectories approach the fixed point only when time is reversed, implying divergence under forward evolution. Near a fixed point—one characterized by saddle-like dynamics featuring both contracting and expanding directions—the stable manifold highlights the paths along which initial perturbations diminish over time. To build intuition, consider a ball on a hilly landscape: the stable manifold resembles the downhill slope leading to a valley, where the ball naturally rolls toward the lowest point, embodying contracting behavior. In opposition, the unstable manifold evokes an uphill escape from a hilltop, where even slight nudges cause the ball to roll away indefinitely, representing expansion. This saddle-point analogy underscores the mixed stability inherent in hyperbolic systems, where some directions pull inward while others push outward. Stable manifolds are instrumental in elucidating attractors within physical phenomena, such as the swirling patterns in fluid flows captured by the Lorenz model or the predictable yet chaotic long-term orbits in , like the tumbling of Saturn's moon Hyperion. By delineating these attracting directions, they reveal how systems evolve toward amid underlying hyperbolicity.

Historical development

The concept of the stable manifold emerged in the late 19th century with Henri Poincaré's 1890 prize memoir on the , where he introduced the notions of stable and unstable manifolds in the context of and nonlinear differential equations. Building on this, Aleksandr Lyapunov's 1892 doctoral thesis laid foundational work on , analyzing linear systems near equilibria via eigenvalues with negative real parts, where the stable eigenspace serves as the linear analog of the stable manifold. This provided key criteria for stable behavior in dynamical systems. Oskar Perron extended these ideas to nonlinear systems in 1929, proving the local existence of stable manifolds for fixed points using of solutions. Early in the , contributed significantly by studying flows on negatively curved manifolds in 1901, where he constructed manifolds that foreshadowed the modern stable manifold framework through iterative methods on solution graphs. These ideas influenced subsequent developments in hyperbolic dynamics. In the mid-20th century, incorporated stable and unstable manifolds into the study of chaotic systems, notably through his 1967 , which demonstrated how such manifolds organize complex, structurally behaviors in dynamical systems. Key milestones in the formalization of the stable manifold theorem include the Hartman-Grobman theorem of the late 1950s, which served as a precursor by establishing local topological conjugacy between nonlinear flows and their linearizations near hyperbolic fixed points, thus justifying the manifold structure. Full proofs of the theorem for smooth manifolds built progressively from the early 20th century works of Hadamard and Perron, with significant advancements in the 1930s onward, culminating in Ya. B. Pesin's 1977 extension to nonuniformly hyperbolic systems, where stable manifolds are shown to exist and be smooth based on Lyapunov exponents. The evolution of the stable manifold concept marked a transition from its origins in celestial mechanics and stability analysis to broader applications in ergodic theory and the study of strange attractors during the 1960s and 1980s, integrating geometric insights with measure-theoretic tools to understand long-term dynamics.

Definitions

Stable and unstable sets

In dynamical systems theory, consider a continuous map f: X \to X, where X is a topological space and p \in X is a fixed point satisfying f(p) = p. The stable set of p, denoted W^s(f, p), is defined as the set of all points q \in X whose forward orbits under f converge to p, that is, W^s(f, p) = \{ q \in X : f^n(q) \to p \text{ as } n \to \infty \}. This definition captures the basin of attraction for the fixed point p in the forward direction. Assuming f is invertible, the unstable set of p, denoted W^u(f, p), consists of all points q \in X whose backward orbits under f converge to p: W^u(f, p) = \{ q \in X : f^{-n}(q) \to p \text{ as } n \to \infty \}. In practice, X is often taken to be a to ensure convergence is well-defined via the metric distance. Both the and unstable sets are under f: for the stable set, f(W^s(f, p)) \subseteq W^s(f, p), and if f is invertible, equality holds; similarly, f(W^u(f, p)) = W^u(f, p). This invariance reflects the fact that the sets are unions of entire orbits approaching p in the respective time directions. For continuous-time dynamical systems, consider a \phi_t: X \to X (defined for t \in \mathbb{R}) generated by a , where p \in X is an point satisfying \phi_t(p) = p for all t. The set of p, denoted W^s(\phi, p), is defined as the set of all points q \in X whose forward orbits under the converge to p, that is, W^s(\phi, p) = \{ q \in X : \phi_t(q) \to p \text{ as } t \to +\infty \}. This captures the basin of attraction for the p in forward time. The unstable set of p, denoted W^u(\phi, p), consists of all points q \in X whose backward orbits converge to p: W^u(\phi, p) = \{ q \in X : \phi_{-t}(q) \to p \text{ as } t \to +\infty \}. In practice, X is often a for well-defined convergence. Both sets are invariant under the : \phi_t(W^s(\phi, p)) = W^s(\phi, p) and \phi_t(W^u(\phi, p)) = W^u(\phi, p) for all t \in \mathbb{R}. Local versions of these sets are defined within small neighborhoods of p to focus on behavior near the fixed point. In a metric space (X, d), the local stable set W^s_\epsilon(f, p) for \epsilon > 0 and ball B_\epsilon(p) is the set of points q \in B_\epsilon(p) whose orbits contract toward p at an exponential rate: W^s_\epsilon(f, p) = \{ q \in B_\epsilon(p) : d(f^n(q), p) \leq K \lambda^n d(q, p) \text{ for all } n \geq 0, \text{ some } K > 0, 0 < \lambda < 1 \}. Such contraction properties arise in hyperbolic settings, where the dynamics exhibit expansion and contraction. A corresponding local unstable set can be defined via backward contraction. For flows, the local stable set is analogously W^s_\epsilon(\phi, p) = \{ q \in B_\epsilon(p) : d(\phi_t(q), p) \leq K e^{-\lambda t} d(q, p) \text{ for all } t \geq 0, \text{ some } K > 0, \lambda > 0 \}.

Local and global manifolds

In the context of a f on a smooth manifold M or a \phi_t generated by a C^k vector field on M, the stable set W^s(f,p) or W^s(\phi,p) defined for a hyperbolic fixed point or equilibrium p \in M can be endowed with a manifold structure under appropriate smoothness conditions. The local stable manifold W^s_{\mathrm{loc}}(f,p) or W^s_{\mathrm{loc}}(\phi,p) is a C^k-embedded submanifold (for suitable k) that takes the form of a disk tangent to the stable eigenspace E^s of Df(p) or D\phi_0(p) within a sufficiently small neighborhood of p. This local version is constructed as the graph of a C^k map from the stable subspace to the unstable directions, ensuring it is diffeomorphic to a ball in the stable eigenspace and properly embedded in the ambient manifold. The global stable manifold W^s(f,p) or W^s(\phi,p) extends this structure to the entire space, comprising all points whose forward orbits converge to p, and forms a C^k immersed of M invariant under f or \phi_t. Invariance means that f(W^s(f,p)) \subseteq W^s(f,p) or \phi_t(W^s(\phi,p)) = W^s(\phi,p), so the map the manifold into itself, preserving its structure globally. Unlike case, the global manifold is obtained by taking the of local stable manifolds along the backward of p, but it need not be closed or properly embedded in M. A key distinction arises between immersed and embedded submanifolds: an immersion is a locally injective map with full-rank differential, but globally it may fail to be injective, leading to self-intersections or non-closed sets. In the global stable manifold, this non-injectivity can occur, for instance, when the manifold loops back on itself in certain dynamical systems, such as in examples where trajectories accumulate in a way that causes overlapping branches without embedding properly in the topology of M. Simple constructions in the literature demonstrate such cases, where the global stable manifold is immersed but not embedded, highlighting the topological complexities beyond local neighborhoods.

Stable Manifold Theorem

Statement

The Stable Manifold Theorem provides a fundamental existence result for invariant submanifolds near hyperbolic fixed points in dynamical systems. Consider a C^r- f: U \subset \mathbb{R}^n \to \mathbb{R}^n with r \geq 1, where U is an open neighborhood of a hyperbolic fixed point p \in U, meaning the Df(p) has no eigenvalues of 1 and splits \mathbb{R}^n into stable E^s(p) and unstable E^u(p) eigenspaces with \dim E^s(p) + \dim E^u(p) = n. Then there exist an open neighborhood V \subset U of p and C^r-immersed submanifolds W^s_{\mathrm{loc}}(p), W^u_{\mathrm{loc}}(p) \subset V such that \dim W^s_{\mathrm{loc}}(p) = \dim E^s(p), \dim W^u_{\mathrm{loc}}(p) = \dim E^u(p), W^s_{\mathrm{loc}}(p) is to E^s(p) at p, W^u_{\mathrm{loc}}(p) is to E^u(p) at p, f(W^s_{\mathrm{loc}}(p)) \subset W^s_{\mathrm{loc}}(p), and f(W^u_{\mathrm{loc}}(p)) \subset W^u_{\mathrm{loc}}(p). Moreover, points on the local stable manifold approach the fixed point exponentially under forward iteration: for x \in W^s_{\mathrm{loc}}(p), there exist constants C > 0 and $0 < \lambda < 1 such that \|f^k(x) - p\| \leq C \lambda^k \|x - p\| for all k \geq 0. Analogously, for x \in W^u_{\mathrm{loc}}(p), \|f^{-k}(x) - p\| \leq C \lambda^k \|x - p\| for all k \geq 0. These properties ensure the manifolds capture the asymptotic behavior dictated by the linearization. The theorem originated in Hadamard's 1901 work on asymptotic solutions of differential equations related to geodesics on negatively curved surfaces, where an unstable manifold was constructed geometrically, and was extended by Perron in 1928 to establish stability and asymptotic behavior for integrals of homogeneous linear differential systems. Full rigorous proofs for general nonlinear maps appeared in the 1950s, building on these foundations. The result generalizes to C^r-diffeomorphisms on finite-dimensional , where the manifolds are immersed submanifolds tangent to the respective eigenspaces of Df(p). Further extensions apply to flows generated by C^r-vector fields on such manifolds, yielding immersed stable and unstable manifolds for hyperbolic equilibria. In infinite dimensions, the theorem holds for C^r-diffeomorphisms on under hyperbolicity conditions on the Fréchet derivative, with manifolds as C^r-immersed submanifolds of matching dimensions.

Conditions and assumptions

A fixed point p of a differentiable f: \mathbb{R}^n \to \mathbb{R}^n is hyperbolic if the Jacobian matrix Df(p) has no eigenvalues on the unit circle in the complex plane. The tangent space at p then decomposes into a direct sum of the stable eigenspace E^s(p), spanned by generalized eigenvectors corresponding to eigenvalues \lambda with |\lambda| < 1, and the unstable eigenspace E^u(p), spanned by those with |\lambda| > 1. This ensures exponential along E^s(p) and along E^u(p) under of the linearized map. The map f must satisfy smoothness conditions for the theorem to hold, typically requiring f to be at least continuously differentiable (C^1) in a neighborhood of p. For stronger regularity of the resulting manifolds, f is often assumed to be C^r with r \geq 1, ensuring the stable manifold inherits C^r smoothness. The theorem applies locally in a sufficiently small neighborhood U of p, where the nonlinear behavior is controlled by the linearization, often verified through estimates on higher derivatives. Proofs of the theorem rely on high-level techniques such as cone conditions, which confine candidate to cones around the eigenspaces to preserve the splitting under iteration, or graph transform methods, which iteratively refine over the to converge to an . These approaches ensure the local stable set forms a immersed tangent to E^s(p). Extensions to non-uniform hyperbolicity, as developed in Pesin theory, relax the strict eigenvalue separation by allowing rates of expansion and contraction to vary by , relying instead on Lyapunov exponents to define a measurable splitting into and unstable directions. In this framework, local stable manifolds exist with respect to an invariant measure, without requiring uniform bounds across the .

Properties

Invariance and uniqueness

The stable manifold W^s(p) of a fixed point p for a f is under the , meaning f(W^s(p)) \subset W^s(p). This invariance ensures that trajectories starting on the manifold remain on it and approach p asymptotically, with distances contracting exponentially along the manifold due to the hyperbolic splitting into stable and unstable subspaces. The proof relies on the transform method, where the manifold is represented as a graph over the stable subspace, and the transform preserves this graph structure while enforcing in the stable directions. Uniqueness holds locally: any two local stable manifolds coincide in a sufficiently small neighborhood of p. This follows from applying the contraction mapping theorem to the space of possible functions over the stable eigenspace, where the transform acts as a with Lipschitz constant less than 1 under the hyperbolicity assumptions. For global extensions, under additional conditions such as those in Anosov flows, the stable manifold is a unique immersed of the entire , extending the local version invariantly via the flow. In contrast to hyperbolic cases, for non-hyperbolic fixed points, center manifolds exhibit similar local invariance but lack uniqueness without further restrictions; any two such manifolds are only C^k-conjugate, preserving up to rather than coinciding pointwise.

Tangent spaces and dimensions

The stable manifold W^s(p) of a fixed point p in a is tangent to the stable eigenspace E^s(p) at p, satisfying T_p W^s(p) = E^s(p). The stable eigenspace E^s(p) consists of the generalized eigenvectors corresponding to the contracting eigenvalues of the matrix at p: for discrete-time systems (diffeomorphisms), those with |\lambda| < 1; for continuous-time systems (flows), those with \operatorname{Re}(\lambda) < 0. This is under the assumption of hyperbolicity where the spectrum splits into contracting and expanding parts. This tangency ensures that near p, the nonlinear manifold locally resembles the provided by the eigenspace. The dimension of the stable manifold equals the dimension of the stable eigenspace: \dim W^s(p) = \dim E^s(p), which is the number of contracting eigenvalues, counted with algebraic multiplicity. In finite-dimensional settings, such as on a manifold M of dimension n, if there are k such contracting eigenvalues, then W^s(p) is a k-dimensional immersed in M. This dimensional matching follows directly from the Stable Manifold Theorem and preserves the geometric structure inherited from the linearization. In the linear case, where the dynamical system is given by a linear map Df(p): T_p M \to T_p M for discrete time or the linearization of a vector field for continuous time, the stable manifold coincides exactly with the stable eigenspace E^s(p), as the invariant subspaces are themselves the manifolds. For nonlinear perturbations of such linear systems, the stable manifold remains tangent to E^s(p) and retains the same dimension, deforming continuously from the linear subspace. The codimension of W^s(p) in M is n - k, equal to the dimension of the unstable eigenspace E^u(p). Consequently, the stable manifold W^s(p) and the unstable manifold W^u(p) are transverse at p, since their tangent spaces E^s(p) and E^u(p) span the full T_p M under the splitting.

Examples and Applications

Physical examples

In the of Saturn's , forces exerted by the planet play a crucial role in shaping particle trajectories. The vertical direction perpendicular to the is a direction, where particles experience contracting motions due to these forces, leading them to oscillate harmonically and remain confined near the equatorial . In contrast, the radial direction is unstable, with stretching causing particles to spread outward, while the tangential direction corresponds to neutral orbital motion. Poincaré sections of these reveal the manifold as separatrices guiding particles toward the plane, illustrating how manifolds organize the overall structure and prevent vertical dispersion despite radial instabilities. A classic physical example of a stable manifold appears in the simple near its inverted position, which acts as a in . At this upright configuration, the linearized equations yield hyperbolic , with the stable manifold consisting of trajectories that approach the as time progresses forward along the stable separatrix. These manifolds serve as boundaries separating bounded oscillatory motion from rotational trajectories, observable in laboratory setups where damping causes initial conditions on the stable manifold to decay toward the unstable fixed point. For the , multiple equilibria exhibit similar stable manifolds, consisting of trajectories that converge to specific joints' alignments, contributing to the system's chaotic separatrix structure and visible in high-speed imaging of coupled oscillations. In , stable manifolds manifest near in flows, such as those in boundary layers over airfoils or in behind bluff bodies. For instance, in the flow past a , the stable manifold associated with a on the surface attracts streamlines that approach it asymptotically, forming ridge-like structures that delineate regions of recirculating flow from the main stream. These manifolds can be visualized experimentally using (), where dye or tracer particles follow paths converging to fixed points, highlighting transport barriers and mixing enhancement in unsteady flows like those in ocean currents or atmospheric boundary layers. Computational visualizations of stable manifolds often employ phase portraits to depict their role as separatrices in low-dimensional systems. In simulations of the pendulum or Lorenz attractor analogs, numerical integration traces trajectories backward in time from saddle points, revealing the stable manifold as a curved surface or curve that initial conditions must lie on to reach the equilibrium; software like or Python's libraries generate these plots, emphasizing how small perturbations off the manifold lead to divergence, underscoring the sensitivity in physical realizations like controlled mechanical oscillators.

Applications in dynamical systems

In chaos theory, stable manifolds play a crucial role in delineating the boundaries of chaotic attractors, where their intersections with unstable manifolds lead to the exponential sensitivity to initial conditions characteristic of chaotic dynamics. A paradigmatic example is the Smale horseshoe map, introduced by in 1967, which models the stretching and folding of in hyperbolic systems, resulting in a structure for the whose boundaries are formed by the stable and unstable manifolds of the fixed points. This construction illustrates how stable manifolds organize the intricate geometry of chaotic invariant sets, enabling the symbolic dynamics description of orbits that shadow the manifolds. Homoclinic tangles arise from transverse intersections between stable and unstable manifolds of hyperbolic fixed points in nearly integrable systems, generating complex, non-wandering sets with dense orbits and contributing to the global structure of phase space. First identified by Henri Poincaré in his 1890 analysis of the three-body problem, these tangles produce a labyrinthine web of orbits that fill the region between the manifolds, leading to chaotic scattering and long-lived transients in perturbed Hamiltonian systems. Smale later formalized this phenomenon in 1967 by embedding the horseshoe as a model for such tangles, showing how they induce symbolic dynamics and infinite symbolic sequences for nearby orbits. In bifurcation analysis, stable manifolds of saddle points determine the boundaries of basins of attraction, influencing the qualitative changes in system behavior during transitions such as saddle-node or Hopf bifurcations. During a saddle-node bifurcation, the collision of a stable node and saddle equilibrium results in the disappearance of their manifolds, reshaping basin boundaries and potentially leading to global changes in attractor stability. In Hopf bifurcations, the stable manifold of a nearby saddle can intersect the emerging limit cycle, defining separatrices that separate basins of coexisting attractors, such as in supercritical cases where the manifold acts as a boundary for the cycle's basin. These roles highlight how stable manifolds govern the partitioning of phase space post-bifurcation, affecting multistability and hysteresis in nonlinear systems. Numerical methods for computing stable manifolds in simulations of ordinary and partial differential equations often rely on graph transform techniques and shadowing lemmas to approximate their global structure accurately. The graph transform method, as developed in the context of normally hyperbolic invariant manifolds, iteratively constructs the manifold as a graph over the by solving fixed-point equations for maps, ensuring under hyperbolicity conditions. Shadowing lemmas, originating from Anosov's work on hyperbolic systems, validate these approximations by guaranteeing the existence of true orbits arbitrarily close to pseudotrajectories with small errors, which is essential for validating numerical integrations in regimes of ODEs and PDEs. These tools enable reliable computation of manifold intersections and tangles in high-dimensional simulations, facilitating the study of transient dynamics and boundary crises.

References

  1. [1]
    Stable Manifold - an overview | ScienceDirect Topics
    The stable manifold is defined as the set of initial conditions whose trajectories asymptotically approach the chaotic saddle over time, ...
  2. [2]
    [PDF] the stable manifold theorem and applications - UChicago Math
    Aug 19, 2019 · The stable manifold theorem is a result from dynamical systems theory. ... behavior of more complex dynamical systems, some basic definitions ...
  3. [3]
    Stable Manifold Theory (Chapter 7) - Nonuniform Hyperbolicity
    Summary. In this chapter we present one of the principal results of the nonuniform hyperbolicity theory – the existence of local stable and unstable manifolds.Missing: scholarly | Show results with:scholarly
  4. [4]
    Dynamical Systems Theory - Aerospace Engineering
    Intuitively, the stable manifold spans the directions along which perturbations decay to over time, the unstable manifold spans the directions along which ...
  5. [5]
    [PDF] Invariant Manifolds
    Examples. Examples of Stable and Unstable Manifolds. Example 1. Find the leading two terms in the expansion of the stable manifold for the system. ˙x = −x ...
  6. [6]
    [PDF] Dynamical systems - Harvard Mathematics Department
    For example, if you are on the stable manifold of an unstable pe- riodic point, then the orbit will converge to that periodic orbit. The Poincaré statement ...
  7. [7]
    CHARACTERISTIC LYAPUNOV EXPONENTS AND SMOOTH ...
    This paper discusses characteristic Lyapunov exponents and smooth ergodic theory, covering topics like basic properties of exponents, entropy, and ergodicity.
  8. [8]
    History of dynamical systems - Scholarpedia
    Oct 21, 2011 · This article provides a brief, and perhaps idiosyncratic, introductory review of the early history of the subject, from approximately 1885 through 1965.<|control11|><|separator|>
  9. [9]
    [PDF] Introduction to Dynamical Systems - Ceremade
    ... define the stable set Vs(x) and unstable set Vu(x) by the formulas. Vs(x) ={y ∈ X: d( fn(x), fn(y)) → 0 as n → ∞},. Vu(x) ={y ∈ X: d( fn(x), fn(y)) → 0 ...
  10. [10]
    [PDF] On the global stable manifold - arXiv
    The stable manifold theorem states that Ws(x) is an immersed Ck submanifold of M. A first way to prove such a result is to define the local stable manifold near ...
  11. [11]
    A center-stable manifold theorem for differential equations in ...
    A center-stable manifold theorem for differential equations in Banach spaces. Published: March 1993. Volume 152, pages 249–268, (1993); Cite this article.
  12. [12]
    [PDF] A New Proof of the Stable Manifold Theorem 1 Introduction
    The stable manifold theorem states that for a smooth map, near a hyperbolic xed point, the stable manifold, points whose forward orbit converges to the xed ...Missing: citation | Show results with:citation
  13. [13]
    [PDF] Stable Manifold Theorem: Part 1
    Nov 10, 1999 · The Hadamard approach uses what is known as a graph transform. Here we define a functional not by an integral but by letting the graph of the ...
  14. [14]
    [PDF] stable manifolds and - hyperbolic sets - eScholarship.org
    The proof of the generalized stable manifold theorem proceeds in the following steps: (A) Let E = E₁ × E₂ be a Banach space; TE→ E a hyperbolic linear map.
  15. [15]
    [PDF] Nonuniform hyperbolicity - Yakov Pesin
    Stable manifold theory. 189. 7.1 The Stable Manifold Theorem. 189. 7.2 Nonuniformly hyperbolic sequences of diffeomorphisms. 192. 7.3 The Hadamard–Perron ...
  16. [16]
  17. [17]
  18. [18]
    [PDF] 1 The Stable Manifold Theorem
    The Stable Manifold Theorem states that there exists a k-dimensional manifold S tangent to the stable subspace Es, and an n-k manifold U tangent to the ...
  19. [19]
    Stable and unstable invariant manifolds in a partially chaotic ...
    Oct 22, 2008 · Examples are magnetic field line tracing in toroidal plasmas1 and flow analysis in oscillating fluids. Hyperbolic fixed points and the ...
  20. [20]
    Smale horseshoe - Scholarpedia
    Nov 30, 2007 · The Smale horseshoe is the hallmark of chaos. With striking geometric and analytic clarity it robustly describes the homoclinic dynamics encountered by Poincar ...Missing: seminal | Show results with:seminal
  21. [21]
    [PDF] Dynamical Systems and Their Bifurcations - Fabio Dercole
    The first and most common collision is that between the stable and unstable manifolds of the same saddle, as depicted in Figure 18. The second collision, shown ...
  22. [22]
    [PDF] B. Coomes SHADOWING IN ORDINARY DIFFERENTIAL EQUATIONS
    Abstract. Shadowing deals with the existence of true orbits of dynamical systems near approximate orbits with sufficiently small local errors.Missing: graph | Show results with:graph