Fact-checked by Grok 2 weeks ago

Jordan curve theorem

The Jordan curve theorem asserts that every simple closed in the \mathbb{R}^2—a continuous, non-self-intersecting loop—separates the plane into exactly two connected components: an interior region, which is bounded, and an exterior region, which is unbounded. The itself forms the common of these two regions, and any point not on the lies in exactly one of these regions. This result, while seemingly intuitive for everyday curves like circles, holds for arbitrary continuous embeddings and is a cornerstone of planar . Named after the French mathematician Camille Jordan (1838–1922), the theorem was first stated and sketched in his 1887 treatise Cours d'analyse de l'École polytechnique, where he defined simple closed curves using parametrizations from the unit interval. However, Jordan's proof was incomplete, particularly in handling the general continuous case beyond polygons, and contained gaps that were later identified. The first rigorous proof was provided by American mathematician in 1905, building on Jordan's ideas and employing early topological methods to establish the separation property. Early proofs include those by Max Dehn around 1900 for polygonal cases and by Nels J. Lennes in 1911 using combinatorial and algebraic approaches. Despite its apparent simplicity, the Jordan curve theorem is notoriously non-trivial to prove, requiring advanced tools from point-set topology and often spanning dozens of pages in modern treatments. It underpins key results in , such as the classification of surfaces, and is essential in for theorems like the argument principle, which relies on the separation of domains by closed contours. In , discrete analogs of the theorem enable algorithms for polygon processing and mesh generation, while generalizations like the Jordan–Schönflies theorem extend it to homeomorphisms with the circle. The theorem's influence persists in higher-dimensional topology, inspiring analogs like the Brouwer fixed-point theorem and invariance of domain.

Foundations

Key Definitions

A Jordan curve is defined as the image of a continuous injective mapping \gamma: S^1 \to \mathbb{R}^2, where S^1 denotes the unit circle in the , ensuring the curve is simple and closed without self-intersections. Equivalently, it can be described as the continuous image of the unit interval [0,1] under a map \gamma such that \gamma(0) = \gamma(1) and \gamma is injective on (0,1). The image of such a mapping exhibits key topological properties in the plane: it is compact as the continuous image of the compact set S^1, connected since S^1 is connected, and locally connected, meaning every point has a local basis of connected open neighborhoods within the curve's induced from \mathbb{R}^2. The interior region of a Jordan curve is the bounded of the complement \mathbb{R}^2 \setminus \gamma(S^1), while the exterior region is the unbounded of the same complement. To understand these regions, basic topological concepts are prerequisites. A topological space X is connected if it cannot be expressed as the union of two disjoint nonempty open subsets, and path-connected if for any two points in X, there exists a continuous path (a continuous map from [0,1] to X) joining them. Representative examples include the unit circle, parametrized by \gamma(\theta) = (\cos \theta, \sin \theta) for \theta \in [0, 2\pi), which embeds smoothly in the plane and bounds the unit disk as its interior. Another is the boundary of the unit square, parametrized piecewise linearly along the edges from (0,0) to (1,0), (1,1), (0,1), and back to $(0,0)$, forming a polygonal Jordan curve whose interior is the square itself.

Statement of the Theorem

The Jordan curve theorem asserts that every Jordan curve in the Euclidean plane \mathbb{R}^2 divides the plane into exactly two connected components: one bounded, referred to as the interior, and one unbounded, referred to as the exterior, with the curve itself serving as the topological boundary of both components. As a corollary, no continuous path can connect a point in the interior to a point in the exterior without intersecting the curve, ensuring a strict topological separation. This separation arises from the property of simple closed curves that they enclose a without allowing connection around them in the without crossing. To distinguish the interior from the exterior, one can use the of the curve around a given point, which equals \pm 1 (depending on ) for points in the interior and $0$ for points in the exterior. A classic example is the circle, where the theorem identifies the enclosed disk as the bounded interior and the surrounding plane as the unbounded exterior, with the circle as the common boundary. Similarly, an ellipse, being a smooth simple closed curve, divides the plane in the same manner, with its oval-shaped region as the interior.

Proofs

Jordan's Classical Proof

Camille Jordan presented his proof of the Jordan curve theorem in the 1887 edition of his textbook Cours d'analyse de l'École polytechnique, where he established that a simple closed curve in the plane separates it into a bounded interior region and an unbounded exterior region. His approach relies on analytic techniques, including uniform continuity of the curve's parametrization and geometric approximations, to construct these regions rigorously for the first time beyond polygonal cases. Rather than purely topological methods, Jordan draws on principles from analysis to handle the continuity of the curve, defining the interior through limiting processes that exhaust the plane's components. The proof begins with the assumption that the curve J is the continuous image of the under a parametrization f: [0,1] \to \mathbb{R}^2 with f(0) = f(1). Jordan exploits this uniform continuity to approximate J by a sequence of P_n, where the distance d(J(t), P_n(t)) < \epsilon_n for sufficiently small \epsilon_n \to 0, ensuring the polygons become arbitrarily close to the curve. For each polygonal approximation P_n, the interior and exterior are well-defined via standard geometric constructions, such as inscribed and exscribed polygons that bound the regions tightly. To extend this to the general curve, Jordan employs exhaustion: he constructs nested sequences of regions U_n^+ (approximating the interior) and U_n^- (approximating the exterior) derived from around the polygon edges, where these tubes act as Dirichlet-like regions with radius r chosen generically to avoid tangencies and ensure transverse intersections. The interior of J is then the intersection \bigcap_n U_n^+, and the exterior is the union \bigcup_n U_n^-, demonstrating separation as the limit of these polygonal separations. Central to distinguishing these regions is the definition of an or turning number for points relative to the . Jordan defines a \pi_J(x) for a point x \notin J, which counts the number of intersections of a from x to with the modulo 2; points with odd lie in the interior, and even in the exterior. This aligns with the of the itself, given by the formula I(J) = \frac{1}{2\pi} \int_J d\theta = \pm 1 for a simple closed , where \theta is the angle of the , confirming the curve's simplicity and the 's constancy in each component. For polygonal approximations, the is computed directly via edge crossings, and ensures it transfers to the limit , with potential-theoretic elements implicit in the harmonic-like behavior of the function across regions. Despite its ingenuity, Jordan's proof faced challenges in handling non-smooth curves, as the parametrization might lack differentiability, complicating definitions and the for the rotation index. He assumes but does not fully justify the control of the time parameter in approximations, potentially allowing pathological behaviors where the curve oscillates wildly. Initially considered incomplete or even incorrect by contemporaries like Osgood and Veblen due to gaps in rigor—particularly the unproven separation for polygons and insufficient detail on tube parameters—the proof omitted explicit verification that the limit regions are connected and nonempty. These issues arose especially for pathological continuous curves, such as those with positive area or infinite , where the exhaustion might fail without additional analytic bounds. Later analyses, however, have shown that minor modifications suffice to render the argument fully rigorous.

Modern Proof Techniques

Modern proofs of the Jordan curve theorem leverage to establish the separation property with greater rigor than Jordan's original analytic approach, which relied on approximations and continuity arguments but left some connectivity issues unresolved. One approach uses the and the to compute \pi_1(\mathbb{R}^2 \setminus J) \cong \mathbb{Z}, generated by a winding once around J. This non-trivial fundamental group indicates the presence of a "hole," and when combined with computations, confirms exactly two path components in the complement. More advanced proofs employ to reveal the structure of \mathbb{R}^2 \setminus J. For instance, the curve J induces a non-trivial in the first group H_1(\mathbb{R}^2 \setminus J) \cong \mathbb{Z}, generated by a winding once around J, confirming that J separates the plane. This follows from computing the \tilde{H}_i(\mathbb{R}^2 \setminus J) using the Mayer-Vietoris sequence: decompose \mathbb{R}^2 \setminus J into two open sets covering the exterior and interior approximations, where the sequence yields \tilde{H}_0(\mathbb{R}^2 \setminus J) \cong \mathbb{Z} (indicating two components) and higher groups vanishing appropriately by induction on dimension. Alternatively, Alexander duality provides a direct approach by relating the of J (homeomorphic to S^1) to the of its complement in the one-point compactification of \mathbb{R}^2, showing \tilde{H}^0(S^2 \setminus J) \cong \mathbb{Z} and thus two complementary components. For accessibility, graph-theoretic expositions approximate J by a polygonal and triangulate the , using planarity criteria to prove separation. In Thomassen's , assume more than two components and construct paths between test points that form a K_{3,3} subdivision, contradicting the non-planarity of K_{3,3}; a similar argument bounds components at most two, confirming exact separation after handling the continuous limit via . Computer-assisted formal verification addresses potential subtleties in these proofs by mechanizing them in proof assistants. Hales formalized a version based on Thomassen's approach in the HOL Light system in 2005, comprising 59,000 lines of code, 1,381 lemmas, and over 44,000 proof steps, rigorously confirming the theorem from basic axioms of real analysis. These modern techniques also resolve gaps in Jordan's original proof, particularly in handling potentially wild embeddings where the curve may exhibit pathological local behavior. By establishing that \mathbb{R}^2 \setminus J is uniformly locally connected—meaning for any compact subset, small neighborhoods connect points without crossing J—the proofs ensure the components are path-connected and open, avoiding the analytic ambiguities in Jordan's approximation arguments that failed for certain continuous but non-rectifiable curves.

History

Original Formulation and Early Reception

The Jordan curve theorem was originally formulated by French mathematician Camille Jordan in 1887, appearing as a lemma in the second volume of his influential textbook Cours d'analyse de l'École polytechnique. In this work, Jordan addressed advanced topics in real and complex analysis, where the theorem served to establish a rigorous separation property for simple closed curves in the plane, essential for developments in contour integration and the residue theorem within complex analysis. This formulation arose amid Jordan's broader investigations into functions of a complex variable, including doubly periodic functions like elliptic functions, where distinguishing bounded interior regions from the unbounded exterior was crucial for analyzing integrals over closed paths. Jordan's presentation treated the result as self-evident enough to warrant only a brief, intuitive argument rather than a full proof, it within the chapter on of functions without extensive discussion. However, this approach reflected the era's limited formal development of , known then as analysis situs, and the lemma passed with modest notice in European mathematical literature, overshadowed by Jordan's contributions to and . The theorem gained wider recognition through its early adoption in American mathematics, particularly via Oswald Veblen's 1905 paper "Theory on Plane Curves in Non-Metrical Analysis Situs," published in the Transactions of the . Veblen offered a simplified and more rigorous proof using combinatorial methods, explicitly acknowledging Jordan's foundational insight while highlighting the theorem's importance for non-metric ; this exposition praised the result's elegance and introduced it effectively to U.S. scholars, fostering its integration into emerging topological studies. Despite this positive reception, some contemporaries raised concerns about the completeness of Jordan's original argument, especially as efforts to extend the theorem to higher dimensions uncovered counterexamples to analogous stronger statements, such as the failure of simple closed surfaces to bound regions homeomorphic to balls in three dimensions. These doubts underscored the need for more precise topological foundations, influencing subsequent refinements while affirming the theorem's validity in the plane.

Developments in Proofs and Understanding

In the 1930s, Kazimierz Kuratowski provided a rigorous proof of the Jordan curve theorem using modern set-theoretic methods, building on the concept of prime ends introduced by Constantin Carathéodory in 1913 to analyze the accessibility of boundary points in simply connected domains. This approach clarified the topological separation properties by decomposing the boundary into equivalence classes of accessible arcs, offering a more precise understanding of how simple closed curves divide the plane without relying on metric assumptions. Kuratowski's work, detailed in his 1933 topology textbook Topologie, emphasized the role of connected components and accessibility, resolving ambiguities in earlier arguments about the nature of the interior and exterior regions. Later, Hassler extended these ideas in his 1937 study of regular closed curves, demonstrating that smooth in the satisfy stronger separation properties, such as the existence of well-defined spaces and indices, which align with the theorem's implications for bounded and unbounded components. Whitney's results, including his work on the uniqueness of embeddings for 3-connected planar graphs up to , provided tools to handle more general continuous embeddings while preserving the theorem's core separation assertion. From the to the , Robion Kirby and Laurence Siebenmann addressed subtleties involving wild curves through their comprehensive study of triangulability for topological manifolds, showing that in dimensions greater than or equal to 5, obstructions to structures (measured by the Kirby-Siebenmann invariant) explain pathological embeddings that challenge naive extensions of planar separation results. Their 1977 monograph resolved key questions about when wild embeddings—such as those with non-locally flat points—admit triangulations, indirectly affirming the robustness of the Jordan curve theorem in the plane by contrasting it with higher-dimensional where separation fails without additional structure. Formal verification efforts began in the 1990s with preliminary computer-assisted checks of special cases, but a landmark achievement came in 2005 when Thomas Hales completed a full in the HOL system, comprising over 44,000 proof steps and confirming the theorem's validity from basic axioms of . This machine-checked proof, later published in 2007, highlighted the theorem's reliance on intricate lemmas about connectedness and boundedness, paving the way for in . Post-2021 developments in , such as enhancements in and for general topological statements, have included minor refinements to proof assistants but no major breakthroughs specific to the Jordan curve theorem as of 2025.

Generalizations

Curves with Multiple Components

A Jordan curve with multiple components is defined as a finite disjoint union of n pairwise disjoint simple closed curves in the plane \mathbb{R}^2, where each component is a standard Jordan curve. The generalized Jordan curve theorem extends the classical result to this setting: the complement \mathbb{R}^2 minus such a multi-component curve consists of exactly n+1 connected components, comprising n bounded components (each serving as the interior of one of the individual Jordan curves) and one unbounded exterior component. This holds regardless of whether the curves are positioned separately or nested, as long as they remain disjoint. A proof can be obtained by induction on n. The base case n=1 is the classical Jordan curve theorem. For the inductive step, consider the first n-1 curves, whose complement has n connected components by the inductive hypothesis. The nth curve lies entirely within one of these components (since the curves are disjoint) and, by the single-curve theorem applied within that open connected region (homeomorphic to an open disk or the exterior), splits it into two subcomponents, thereby increasing the total number of components by exactly one. For a example, consider two concentric in the , forming a of two curves. Their complement has three connected components: the open disk bounded by the inner , the bounded annular region between the circles, and the unbounded exterior region outside the outer . Note that a figure-eight shape, while visually suggestive of multiple loops, is not a valid multi-component Jordan curve, as it self-intersects at a point and thus fails to be a of simple closed curves. When the multi-component Jordan curve is polygonal (approximating smooth curves via finite edges and vertices), it embeds as a disconnected consisting of n cycle components. The arrangement divides the plane into n+1 faces (regions), and the of this embedding satisfies V - E + F = n + 1, where V is the number of vertices, E the number of edges, F = n + 1 the number of faces (including the single unbounded face), and n is the number of graph components; this aligns with the general formula for disconnected s, V - E + F = c + 1 with c = n.

Higher-Dimensional Analogues

The Jordan-Brouwer separation generalizes the Jordan curve to higher dimensions, stating that in \mathbb{R}^n, a topological of the (n-1)-sphere S^{n-1} separates \mathbb{R}^n into exactly two connected components: one bounded (the interior) and one unbounded (the exterior). This result, proved by Luitzen Egbertus Jan Brouwer in 1910, extends the planar separation property to hyperspheres in of any dimension n \geq 2. The relies on the and of the embedded sphere, ensuring that paths crossing the sphere connect the two components in a topologically distinct manner. A stronger analogue, the Schoenflies theorem, holds in the plane: if J is a Jordan curve in \mathbb{R}^2, then the closure of its bounded complementary component is homeomorphic to the closed 2-disk D^2. Proved by Arthur Schoenflies in 1906 building on Jordan's work, this asserts that the embedding is "tame," allowing a homeomorphism of the plane that maps the curve to the standard circle. However, this theorem fails in three dimensions: there exist embeddings of S^2 in \mathbb{R}^3 whose bounded complementary component is not homeomorphic to the 3-ball D^3. A seminal counterexample is the , constructed by in 1924, which is a wild of S^2 into \mathbb{R}^3. This sphere consists of interlocking "horns" that branch infinitely, creating a bounded region whose is non-trivial due to linked paths that cannot be contracted without intersecting the sphere. As a result, the embedding is not tamable by any of \mathbb{R}^3 to the standard sphere, highlighting the pathology of three-dimensional compared to the plane. Modern extensions of these separation results employ algebraic topology, particularly homology and cohomology, to establish analogues in more general manifolds. In \mathbb{R}^n, the non-trivial homology group H_{n-1}(S^{n-1}; \mathbb{Z}) \cong \mathbb{Z} implies that an embedded (n-1)-sphere induces a separation, as the generator detects the distinction between the interior and exterior components via cycles that link the sphere. For oriented manifolds, cohomology classes further classify such hypersurfaces, ensuring duality between the homology of the complement and the sphere itself, as formalized in Alexander duality. These tools provide robust proofs for separation in smooth or PL categories, though wild embeddings persist as obstacles in the topological category for n \geq 3.

Discrete and Digital Versions

In the discrete setting, a Jordan curve is typically defined as a simple closed chain of lattice points or pixels in the integer grid \mathbb{Z}^2, without self-intersections, where adjacency is specified by 4-connectivity (sharing an edge) or 8-connectivity (sharing an edge or corner). Such curves arise in digital topology, which studies topological properties of grid-based structures analogous to those in the continuous plane. A analogue of the Jordan theorem states that, under suitable assumptions, a simple closed 8-connected in \mathbb{Z}^2 separates the digital plane into exactly two 4-connected components: an interior and an exterior . In this formulation, every point on the is adjacent (via 4-connectivity) to points in both , ensuring no path between the regions avoids the . This holds in topologies like the Khalimsky topology on \mathbb{Z}^2, where neighborhoods are defined such that even-parity points (both coordinates even) are open, odd-parity points (both odd) are closed, and mixed-parity points are half-open, preserving separation . rules further govern : for instance, 4-connected paths preserve parity in one coordinate, while 8-connected paths can change it, enforcing the duality between and connectivities. However, the theorem fails in certain discrete settings, particularly for 4-connected curves when regions are considered 4-connected, due to "leaks" where diagonal (8-connected) paths connect the supposed interior and exterior without crossing the curve. Neither pure 4-adjacency nor pure 8-adjacency alone supports the separation; paired connectivities (e.g., 8 for curves, 4 for regions) or specialized topologies like Khalimsky's are required to avoid such failures. The Steinhaus chessboard theorem provides a related discrete result: on an m \times n colored in the standard alternating pattern, no simple closed path (using king moves, i.e., 8-connectivity) can enclose an equal number of squares without crossing edges between them. This theorem implies a combinatorial form of separation, as the unequal enclosure reflects the topological distinction between interior and exterior, and it serves as a in proofs of Jordan analogues on finite grids. Discrete versions also connect to the Brouwer fixed-point theorem via games like , where the board forms a hexagonal grid approximating \mathbb{Z}^2. The theorem—that a full game on an n \times n board results in a winning path for one player (connecting opposite sides)—is a finite analogue of Jordan separation, as the winner's path acts as a blocking the loser's connection, provable combinatorially without continuous assumptions. This links to Brouwer's theorem in settings, where no fixed-point-free map exists on the grid, mirroring continuous non-separability.

Applications

Topological and Geometric Applications

The Jordan curve theorem plays a fundamental role in the of surfaces by enabling the identification and removal of handles, which are non-separating annuli that contribute to the of a surface. Specifically, the theorem implies that on a , any simple closed curve separates the surface into two regions, ensuring no handles exist since the complement remains connected after removal. For higher- surfaces, the genus g is defined as the maximum number of disjoint simple closed curves whose removal leaves the complement connected; these curves correspond to two-sided cycles on orientable surfaces, allowing systematic via \chi = 2 - 2g. In , the Jordan curve theorem underpins by rigorously defining the interior of a simple closed curve, which is essential for establishing that holomorphic functions integrate to zero over such boundaries enclosing their domain of analyticity. For a rectifiable simple closed curve \gamma in a \Omega, the theorem guarantees that \gamma divides the into an interior contained in \Omega and an exterior, allowing the statement \int_\gamma f(z) \, dz = 0 for holomorphic f on and inside \gamma. This interior definition extends to residue theory, where computes residues at isolated singularities within the bounded , facilitating techniques. The theorem also supports the , which states that any simply connected domain in the (excluding the whole ) is conformally equivalent to the unit disk; the simply connected interior provided by the theorem is crucial for this result in . Geometrically, the theorem supports polygonization of point sets by ensuring that a simple polygonal chain forms a closed bounding a well-defined interior , which is crucial for constructing simple from scattered points without self-intersections. In simple polygon recognition, it provides the topological foundation for algorithms that verify whether a given polygonal path is simple by checking separation properties, such as point-in-polygon tests that count boundary crossings to determine interior points. These applications rely on the theorem's assurance that a non-self-intersecting divides the plane into interior and exterior components. The Jordan curve theorem connects to the through its role in analyzing planar maps, where boundaries of regions are simple closed curves that separate the plane into distinct faces. In proofs of the five color theorem—a precursor to the result—it ensures that Kempe chains, used to recolor adjacent vertices, cannot intersect without violating planarity, as the theorem partitions the plane to prevent such crossings in graph embeddings. This separation property formalizes the notion of adjacent regions in , confirming that four colors suffice for any . A direct illustration of the theorem's implications is its use in proving that a simple closed bounds a simply connected : the bounded component of the complement, known as the interior, is path-connected and has the as its , with any inside contractible to a point within that , establishing simply connectedness.

Computational Algorithms

The ray-casting algorithm, also known as the crossing number or even-odd rule, provides an efficient method to determine whether a point lies inside a simple closed , leveraging the separation of the Jordan curve theorem. The procedure involves emitting a ray from the query point to and counting the intersections with the polygon's edges; an odd count indicates the point is interior, while an even count signifies exterior. This approach runs in time for a polygon with n vertices and requires careful handling of edge cases, such as ray alignment with vertices or horizontal edges, to ensure correctness for non-convex polygons. An alternative to ray-casting is the algorithm, which computes the net number of counterclockwise revolutions the makes around the point by summing oriented or using cross-products between consecutive edges. For a simple closed curve, a non-zero winding number confirms the point is inside the bounded region. This method also achieves O(n) and is particularly straightforward for convex polygons, where the summation avoids by relying solely on sign tests of cross-products. Before applying these tests, verifying that a given polygonal chain forms a simple closed is essential, as self-intersections violate the theorem's assumptions. Simplicity testing detects such using sweep-line algorithms, notably the Bentley-Ottmann method, which maintains a dynamic set of active segments ordered by their intersection with a sweeping vertical line. The algorithm reports all intersections in O((n + k) log n) time, where k is the number of intersections; for mere existence (to decide simplicity), the worst-case time is O(n log n) since k can be quadratic but is not fully enumerated. Extensions handle weakly simple polygons, allowing touching but non-crossing edges, in O(n log n) time via similar sweep techniques. A deeper computational challenge arises in discretized or approximate settings, where a computational version of the —such as verifying separation properties for grid-based or continuous approximations—is PPAD-complete. This hardness result stems from reductions to approximating . In discrete and digital settings, such as pixel grids approximating continuous curves, the complexity of verifying Jordan properties remains tied to PPAD-hardness for high-precision approximations without additional structure.

Modern Uses in Science and Engineering

In image processing, the discrete analogue of the Jordan curve theorem underpins detection and segmentation algorithms by ensuring that a closed divides the pixel grid into interior and exterior regions, facilitating accurate object isolation in binary or grayscale images. This principle is applied in libraries like , where contour detection functions, such as findContours, identify closed boundaries that separate foreground objects from the background, enabling tasks like analysis and defect detection in . For instance, the theorem's version supports point-in-polygon tests via ray-casting methods, which count crossings to determine region membership, a core step in segmenting irregular shapes without topological ambiguities. In , the Jordan curve theorem informs path planning and obstacle avoidance by guaranteeing that closed obstacle boundaries partition the 2D navigation space into safe and unsafe regions, allowing algorithms to compute collision-free trajectories for autonomous vehicles. For example, in layered path planning for mobile robots, the theorem ensures that visibility graphs or potential fields respect boundary separations, preventing paths from erroneously crossing into bounded obstacle interiors during real-time mapping in unknown environments. This application is critical for ensuring global optimality in environments with polygonal obstacles, as seen in algorithms that evolve junction points along boundaries to find shortest paths while maintaining topological consistency. The Jordan curve theorem contributes to through (TDA), where it provides the foundational separation property for detecting boundaries in feature spaces, particularly in convolutional neural networks (CNNs) for . In TDA-integrated models, computes topological features like loops corresponding to closed curves, enabling robust boundary separation in noisy . Bayesian approaches to boundary detection formalize closed curves via the theorem to model probabilistic separations, enhancing performance in tasks like semantic segmentation. In geographic information systems (GIS) and , the Jordan curve theorem validates the definition of simple closed curves as boundaries for administrative regions and flood zones, ensuring that polygons accurately delineate enclosed areas without self-intersections. This is essential for flood risk mapping, where theorem-based point-in-polygon queries determine whether locations fall inside hazard zones defined by riverine or coastal boundaries, supporting spatial overlay analyses in tools like . The theorem's role extends to maintaining topological integrity in vector data models, preventing errors in boundary representations during raster-to-vector conversions for and . Emerging applications in utilize the Jordan curve theorem to model cell as closed curves that separate intracellular and extracellular spaces, aiding simulations of dynamic processes like blebbing in migrating cells. In geometric analyses of amoeboid motion, the theorem helps identify interior regions bounded by contours, allowing of cellular shapes while excluding exterior artifacts in 2D projections of models. computing paradigms, inspired by biological compartments, explicitly invoke the theorem to define nested regions without intersections, facilitating algorithmic studies of protein and vesicle formation.

References

  1. [1]
    01-simple-polygons
    The theorem states that any simple closed curve partitions the plane into two connected subsets, exactly one of which is bounded.
  2. [2]
    [PDF] The Jordan Curve Theorem, Formally and Informally
    Jan 31, 2018 · The Jordan curve theorem states that every simple closed pla- nar curve separates the plane into a bounded interior region and an unbounded ...
  3. [3]
    [PDF] RESEARCH ARTICLE The Jordan curve theorem is non-trivial
    The first proof, hence name Jordan curve theorem, was given in 1887 by Camille Jordan in his book Cours d'analyse de l'´Ecole Polytechnique [4] but was ...
  4. [4]
    [PDF] invariance of domain and the jordan curve theorem in r2
    Invariance of Domain and the Jordan Curve Theorem are two simply stated but yet very useful and important theorems in topology. They state the following ...
  5. [5]
    Green's theorem and Jordan rectifiable curve
    complicated Jordan curves, it needs proof. This was pointed out by Camille Jordan in 1892. His proof was incomplete, and later Veblen (1905) corrected it.Missing: history | Show results with:history
  6. [6]
    [PDF] The Jordan curve theorem is one of the most fundamental results in ...
    It is an important result in linear algebra, and is a certain form of a matrix, often called the Jordan canonical form [13]. Camille Jordan, one of the most ...
  7. [7]
    [PDF] Jordan's Proof of the Jordan Curve Theorem - School of Mathematics
    The celebrated theorem of Jordan states that every simple closed curve in the plane separates the complement into two connected nonempty sets: an interior.
  8. [8]
    10.5 Jordan measurable sets
    A bounded set is Jordan measurable if and only if the boundary is a measure zero set. Proof. Suppose is a closed rectangle such that is contained in the ...
  9. [9]
    [PDF] 7.11 Appendix: Index theory in two dimensions
    Theorem 7.11. 1. (Jordan Curve Theorem): Every Jordan curve Γ in the plane separates R2 into two disjoint, open, connected sets, both of which have Γ as their ...
  10. [10]
    Spring 2009 - Problem 11
    Indeed, by the Jordan curve theorem, a simple closed curve has winding number in { − 1 , 0 , 1 } \set{-1, 0, 1} {−1,0,1}, depending on the orientation.
  11. [11]
    [PDF] The Jordan-Schönflies Theorem and the Classification of Surfaces
    The Jordan curve theorem says that a simple closed curve in the Euclidean plane partitions the plane into precisely two parts: the interior and the exterior of ...
  12. [12]
    [PDF] global properties of plane and space curves
    We define the rotation index of the curve as the quantity. 1. 2π. ∫. C κg ds. Page 8. 8. KEVIN YAN. Proposition 3.8. The rotation index is always an integer.
  13. [13]
    [PDF] Algebraic Topology - Cornell Mathematics
    This book was written to be a readable introduction to algebraic topology with rather broad coverage of the subject. The viewpoint is quite classical in ...
  14. [14]
    [PDF] Illustrated Proof of the Jordan Curve Theorem by Rich Schwartz This ...
    Illustrated Proof of the Jordan Curve Theorem by Rich Schwartz. This note exposits J. W. Alexander's brilliant proof [A] of the Jordan Curve. Theorem.
  15. [15]
    [PDF] The Jordan Curve Theorem
    Abstract. The Jordan Curve Theorem (JCT) is arguably best-known for stating an apparently obvious fact while reportedly being quite hard to prove.Missing: original | Show results with:original
  16. [16]
    [PDF] The Jordan Curve Theorem, Formally and Informally
    Hales, The formal proof of the Jordan curve theorem (HOL Light code), available at http://www/ math.pitt.edu/~thales/. 5. J. Harrison, HOL Light, available ...
  17. [17]
    [PDF] The Jordan curve theorem revisited
    Thus, giving a simple proof of JCT for standard curves (see (2.1)-(2.6) below) establishes the theorem for all curves one encounters in classical function ...Missing: triangulation | Show results with:triangulation
  18. [18]
    [PDF] Camille Jordan English version - University of St Andrews
    The JORDAN curve theorem: Every closed non-intersecting JORDAN curve in the Euclidean plane divides it into two disjoint regions whose common edge is the ...Missing: infinitaire | Show results with:infinitaire
  19. [19]
    [PDF] theory on plane curves in non-metrical analysis situs
    JORDAN'S † explicit formulation of the fundamental theorem that a simple closed curve lying wholly in a plane decomposes the plane into an inside and an outside ...
  20. [20]
    [PDF] June 10, 2021 THE JORDAN CURVE THEOREM 1. Arc and Jordan ...
    Jun 10, 2021 · (2) Jordan curve theorem was generalized to higher dimensions by Brouwer: Theorem 2.7 (Jordan-Brouwer Separation Theorem). Suppose X ⊂ Rn ...<|control11|><|separator|>
  21. [21]
    (PDF) A Proof of the Jordan Curve Theorem - ResearchGate
    Aug 10, 2025 · Hales proves the JCT for planar rectangular grids with the HOL Light system, following the Kuratowski characterization of planarity (12) . ...
  22. [22]
    [PDF] the jordan-brouwer separation theorem - UChicago Math
    Sep 28, 2009 · Abstract. The Classical Jordan Curve Theorem says that every simple closed curve in R2 divides the plane into two pieces, an “inside” and an ...
  23. [23]
    [PDF] On the Smooth Jordan Brouwer Separation Theorem - Penn Math
    Mar 1, 2016 · Abstract. We give an elementary proof of the Jordan Brouwer separation theorem for smooth hypersurfaces using the divergence theorem and the ...
  24. [24]
    [PDF] The generalized Schoenflies theorem
    The generalized Schoenflies theorem asserts that if ϕ ∶ Sn−1 → Sn is a topological embedding and A is the closure of a component of Sn ∖ϕ(Sn−1), then A ≅ Dn as ...
  25. [25]
    A proof of the generalized Schoenflies theorem - Project Euclid
    There is no online version at this time. Please download the PDF instead. Access the abstract. Morton Brown "A proof of the generalized Schoenflies theorem," ...
  26. [26]
    [PDF] Algebraic Topology - UCR Math Department
    Aug 21, 2003 · Here is a homological theorem which applies to all dimensions and as we shall see easily implies the classical Jordan curve theorem. Theorem 3.2 ...
  27. [27]
    (PDF) Jordan Curves in the Digital Plane - ResearchGate
    Aug 6, 2025 · We present a digital analogue of the Jordan curve theorem for each of the pre-topologies to demonstrate that they can provide background ...Missing: leaks | Show results with:leaks
  28. [28]
    A discrete form of Jordan curve theorem - Uniwersytet Śląski
    The paper includes purely combinatorial proof of a theorem that implies the following theorem, stated by Hugo Steinhaus in [6, p.35]: consider a chessboard ( ...
  29. [29]
    (PDF) Digital Jordan Curve Theorems - ResearchGate
    Aug 7, 2025 · Efim Khalimsky's digital Jordan curve theorem states that the complement of a Jordan curve in the digital plane equipped with the Khalimsky ...
  30. [30]
    [PDF] The Complexity of Hex and the Jordan Curve Theorem - Erik Demaine
    For Brouwer, we give a special case in two dimensions, involving functions from the unit square to itself. (This can be extended to the general two-dimensional ...Missing: doubts | Show results with:doubts
  31. [31]
    [PDF] Surface Classification - Jeff Erickson
    Feb 8, 2013 · The Jordan curve theorem implies that the sphere has. 6 no handles, but this theorem does not extend to other surfaces. To detach the handle ...
  32. [32]
    [PDF] CAUCHY'S THEOREM FOR SIMPLE CURVES Let Ω be a region in ...
    A rectifiable curve is a curve of finite length. The seemingly self-evident fact that a simple closed curve has an interior and an exterior is the Jordan Curve ...
  33. [33]
    [PDF] 1 The Jordan Polygon Theorem - CGVR
    The Jordan Curve Theorem and its generalizations are the formal foundations of many results, if not every result, in surface topology.
  34. [34]
    [PDF] The Four Color Theorem - UChicago Math
    Then, by the. Jordan Curve Theorem, H1 and H2 must intersect someplace. However, this contradicts the planarity of the graph. So, H1 and H2 cannot both exist, ...
  35. [35]
    [PDF] The Jordan-Schönflies Theorem and the Classification of Surfaces
    This can be done by triangulating the two surfaces by the same graph G as above). ... Tverberg, A proof of the Jordan Curve Theorem, Bull. London Math. Soc ...
  36. [36]
    [PDF] The Point in Polygon Problem for Arbitrary Polygons
    The point in polygon problem determines if a point is inside a polygon. Solutions include the even-odd rule and the winding number, which is the number of ...
  37. [37]
    [PDF] Computing intersections in a set of line segments - Carleton University
    Oct 14, 2003 · How can we decide if L and L0 intersect and, if they do, how can we compute their intersection? 3. Page 4. 3.1 The Bentley-Ottmann algorithm. We ...
  38. [38]
    [PDF] Detecting Weakly Simple Polygons∗ - Hsien-Chih Chang
    Jul 7, 2014 · In the first phase, we determine in O(nlogn) time whether any two edges of P cross, using the classical. 18 sweep-line algorithm of Shamos and ...
  39. [39]
    A Jordan Curve Theorem in the Digital Plane | Request PDF
    ... Jordan curve theorem. This enables using the closure operator for structuring the digital plane in order to study and process digital images. An advantage ...
  40. [40]
    PNPOLY - Point Inclusion in Polygon Test - WR Franklin (WRF)
    At each crossing, the ray switches between inside and outside. This is called the Jordan curve theorem. The case of the ray going thru a vertex is handled ...
  41. [41]
    [PDF] Path-Planning Strategies for a Point Mobile Automaton Moving ...
    Essentially, the algorithms described below are exploiting the Jordan Curve Theorem, which states that any closed curve homeomorphic to a circle drawn ...
  42. [42]
    [PDF] Layered Path Planning for an Autonomous Mobile Robot - DTIC
    The Jordan Curve Theorem states that a simple closed curve C in the Euclidean plane separates the plane into two open connected sets with C as their common ...<|separator|>
  43. [43]
    Bayesian detection of image boundaries - Project Euclid
    In view of a converse of the Jordan curve theorem, we represent the closed boundary γ as a function indexed by Sd−1. , that is, γ : Sd−1 → R+ : s → γ (s) ...<|separator|>
  44. [44]
    [PDF] Topology of Deep Neural Networks - Department of Statistics
    By the Jordan Curve Theorem, there is no homeomorphism. R ... Topological methods for the analysis of high dimensional data sets and 3D object recognition.
  45. [45]
    [PDF] Geographic Information Systems in Water Resources Engineering
    ... GIS. One definition of whether a point is inside a region is the Jordan curve theorem (Haines 1994). Essentially, it says that a point is inside a polygon ...
  46. [46]
    [PDF] A triangulation-based approach to automatically repair GIS polygons
    Given one simple and closed boundary in the plane, finding its interior is straightforward since, as the Jordan curve theorem states, the boundary divides ...
  47. [47]
    Advances in geometric techniques for analyzing blebbing in ...
    Feb 14, 2019 · To remove the exterior ones, we use a procedure based on the Jordan curve theorem. For interior triangles, a line joining the center of the ...