Jordan curve theorem

Last updated
Illustration of the Jordan curve theorem. The Jordan curve (drawn in black) divides the plane into an "inside" region (light blue) and an "outside" region (pink). Jordan curve theorem.png
Illustration of the Jordan curve theorem. The Jordan curve (drawn in black) divides the plane into an "inside" region (light blue) and an "outside" region (pink).

In topology, the Jordan curve theorem asserts that every Jordan curve (a plane simple closed curve) divides the plane into an "interior" region bounded by the curve and an "exterior" region containing all of the nearby and far away exterior points. Every continuous path connecting a point of one region to a point of the other intersects with the curve somewhere. While the theorem seems intuitively obvious, it takes some ingenuity to prove it by elementary means. "Although the JCT is one of the best known topological theorems, there are many, even among professional mathematicians, who have never read a proof of it." (Tverberg (1980 , Introduction)). More transparent proofs rely on the mathematical machinery of algebraic topology, and these lead to generalizations to higher-dimensional spaces.

Contents

The Jordan curve theorem is named after the mathematician Camille Jordan (1838–1922), who published its first claimed proof in 1887. [1] [2] For decades, mathematicians generally thought that this proof was flawed and that the first rigorous proof was carried out by Oswald Veblen. However, this notion has been overturned by Thomas C. Hales and others. [3]

Definitions and the statement of the Jordan theorem

A Jordan curve or a simple closed curve in the plane R2 is the image C of an injective continuous map of a circle into the plane, φ: S1R2. A Jordan arc in the plane is the image of an injective continuous map of a closed and bounded interval [a, b] into the plane. It is a plane curve that is not necessarily smooth nor algebraic.

Alternatively, a Jordan curve is the image of a continuous map φ: [0,1] → R2 such that φ(0) = φ(1) and the restriction of φ to [0,1) is injective. The first two conditions say that C is a continuous loop, whereas the last condition stipulates that C has no self-intersection points.

With these definitions, the Jordan curve theorem can be stated as follows:

Theorem  Let C be a Jordan curve in the plane R2. Then its complement, R2 \ C, consists of exactly two connected components. One of these components is bounded (the interior) and the other is unbounded (the exterior), and the curve C is the boundary of each component.

In contrast, the complement of a Jordan arc in the plane is connected.

Proof and generalizations

The Jordan curve theorem was independently generalized to higher dimensions by H. Lebesgue and L. E. J. Brouwer in 1911, resulting in the Jordan–Brouwer separation theorem.

Theorem  Let X be an n-dimensional topological sphere in the (n+1)-dimensional Euclidean space Rn+1 (n > 0), i.e. the image of an injective continuous mapping of the n-sphere Sn into Rn+1. Then the complement Y of X in Rn+1 consists of exactly two connected components. One of these components is bounded (the interior) and the other is unbounded (the exterior). The set X is their common boundary.

The proof uses homology theory. It is first established that, more generally, if X is homeomorphic to the k-sphere, then the reduced integral homology groups of Y = Rn+1 \ X are as follows:

This is proved by induction in k using the Mayer–Vietoris sequence. When n = k, the zeroth reduced homology of Y has rank 1, which means that Y has 2 connected components (which are, moreover, path connected), and with a bit of extra work, one shows that their common boundary is X. A further generalization was found by J. W. Alexander, who established the Alexander duality between the reduced homology of a compact subset X of Rn+1 and the reduced cohomology of its complement. If X is an n-dimensional compact connected submanifold of Rn+1 (or Sn+1) without boundary, its complement has 2 connected components.

There is a strengthening of the Jordan curve theorem, called the Jordan–Schönflies theorem, which states that the interior and the exterior planar regions determined by a Jordan curve in R2 are homeomorphic to the interior and exterior of the unit disk. In particular, for any point P in the interior region and a point A on the Jordan curve, there exists a Jordan arc connecting P with A and, with the exception of the endpoint A, completely lying in the interior region. An alternative and equivalent formulation of the Jordan–Schönflies theorem asserts that any Jordan curve φ: S1R2, where S1 is viewed as the unit circle in the plane, can be extended to a homeomorphism ψ: R2R2 of the plane. Unlike Lebesgue's and Brouwer's generalization of the Jordan curve theorem, this statement becomes false in higher dimensions: while the exterior of the unit ball in R3 is simply connected, because it retracts onto the unit sphere, the Alexander horned sphere is a subset of R3 homeomorphic to a sphere, but so twisted in space that the unbounded component of its complement in R3 is not simply connected, and hence not homeomorphic to the exterior of the unit ball.

Discrete version

The Jordan curve theorem can be proved from the Brouwer fixed point theorem (in 2 dimensions), [4] and the Brouwer fixed point theorem can be proved from the Hex theorem: "every game of Hex has at least one winner", from which we obtain a logical implication: Hex theorem implies Brouwer fixed point theorem, which implies Jordan curve theorem. [5]

It is clear that Jordan curve theorem implies the "strong Hex theorem": "every game of Hex ends with exactly one winner, with no possibility of both sides losing or both sides winning", thus the Jordan curve theorem is equivalent to the strong Hex theorem, which is a purely discrete theorem.

The Brouwer fixed point theorem, by being sandwiched between the two equivalent theorems, is also equivalent to both. [6]

In reverse mathematics, and computer-formalized mathematics, the Jordan curve theorem is commonly proved by first converting it to an equivalent discrete version similar to the strong Hex theorem, then proving the discrete version. [7]

Application to image processing

In image processing, a binary picture is a discrete square grid of 0 and 1, or equivalently, a compact subset of . Topological invariants on , such as number of components, might fail to be well-defined for if does not have an appropriately defined graph structure.

There are two obvious graph structures on :

8-neighbor and 4-neighbor square grids. Sasiedztwa 4 8.svg
8-neighbor and 4-neighbor square grids.
  • the "4-neighbor square grid", where each vertex is connected with .
  • the "8-neighbor square grid", where each vertex is connected with iff , and .

Both graph structures fail to satisfy the strong Hex theorem. The 4-neighbor square grid allows a no-winner situation, and the 8-neighbor square grid allows a two-winner situation. Consequently, connectedness properties in , such as the Jordan curve theorem, do not generalize to under either graph structure.

If the "6-neighbor square grid" structure is imposed on , then it is the hexagonal grid, and thus satisfies the strong Hex theorem, allowing the Jordan curve theorem to generalize. For this reason, when computing connected components in a binary image, the 6-neighbor square grid is generally used. [8]

Steinhaus chessboard theorem

The Steinhaus chessboard theorem in some sense shows that the 4-neighbor grid and the 8-neighbor grid "together" implies the Jordan curve theorem, and the 6-neighbor grid is a precise interpolation between them. [9] [10]

The theorem states that: suppose you put bombs on some squares on a chessboard, so that a king cannot move from the bottom side to the top side without stepping on a bomb, then a rook can move from the left side to the right side stepping only on bombs.

History and further proofs

The statement of the Jordan curve theorem may seem obvious at first, but it is a rather difficult theorem to prove. Bernard Bolzano was the first to formulate a precise conjecture, observing that it was not a self-evident statement, but that it required a proof. [11] It is easy to establish this result for polygons, but the problem came in generalizing it to all kinds of badly behaved curves, which include nowhere differentiable curves, such as the Koch snowflake and other fractal curves, or even a Jordan curve of positive area constructed by Osgood (1903).

The first proof of this theorem was given by Camille Jordan in his lectures on real analysis, and was published in his book Cours d'analyse de l'École Polytechnique. [1] There is some controversy about whether Jordan's proof was complete: the majority of commenters on it have claimed that the first complete proof was given later by Oswald Veblen, who said the following about Jordan's proof:

His proof, however, is unsatisfactory to many mathematicians. It assumes the theorem without proof in the important special case of a simple polygon, and of the argument from that point on, one must admit at least that all details are not given. [12]

However, Thomas C. Hales wrote:

Nearly every modern citation that I have found agrees that the first correct proof is due to Veblen... In view of the heavy criticism of Jordan’s proof, I was surprised when I sat down to read his proof to find nothing objectionable about it. Since then, I have contacted a number of the authors who have criticized Jordan, and each case the author has admitted to having no direct knowledge of an error in Jordan’s proof. [13]

Hales also pointed out that the special case of simple polygons is not only an easy exercise, but was not really used by Jordan anyway, and quoted Michael Reeken as saying:

Jordan’s proof is essentially correct... Jordan’s proof does not present the details in a satisfactory way. But the idea is right, and with some polishing the proof would be impeccable. [14]

Earlier, Jordan's proof and another early proof by Charles Jean de la Vallée Poussin had already been critically analyzed and completed by Schoenflies (1924). [15]

Due to the importance of the Jordan curve theorem in low-dimensional topology and complex analysis, it received much attention from prominent mathematicians of the first half of the 20th century. Various proofs of the theorem and its generalizations were constructed by J. W. Alexander, Louis Antoine, Ludwig Bieberbach, Luitzen Brouwer, Arnaud Denjoy, Friedrich Hartogs, Béla Kerékjártó, Alfred Pringsheim, and Arthur Moritz Schoenflies.

New elementary proofs of the Jordan curve theorem, as well as simplifications of the earlier proofs, continue to be carried out.

The root of the difficulty is explained in Tverberg (1980) as follows. It is relatively simple to prove that the Jordan curve theorem holds for every Jordan polygon (Lemma 1), and every Jordan curve can be approximated arbitrarily well by a Jordan polygon (Lemma 2). A Jordan polygon is a polygonal chain, the boundary of a bounded connected open set, call it the open polygon, and its closure, the closed polygon. Consider the diameter of the largest disk contained in the closed polygon. Evidently, is positive. Using a sequence of Jordan polygons (that converge to the given Jordan curve) we have a sequence presumably converging to a positive number, the diameter of the largest disk contained in the closed region bounded by the Jordan curve. However, we have to prove that the sequence does not converge to zero, using only the given Jordan curve, not the region presumably bounded by the curve. This is the point of Tverberg's Lemma 3. Roughly, the closed polygons should not thin to zero everywhere. Moreover, they should not thin to zero somewhere, which is the point of Tverberg's Lemma 4.

The first formal proof of the Jordan curve theorem was created by Hales (2007a) in the HOL Light system, in January 2005, and contained about 60,000 lines. Another rigorous 6,500-line formal proof was produced in 2005 by an international team of mathematicians using the Mizar system. Both the Mizar and the HOL Light proof rely on libraries of previously proved theorems, so these two sizes are not comparable. NobuyukiSakamotoandKeita Yokoyama ( 2007 ) showed that in reverse mathematics the Jordan curve theorem is equivalent to weak Kőnig's lemma over the system .

Application

If the initial point (pa) of a ray (in red) lies outside a simple polygon (region A), the number of intersections of the ray and the polygon is even.
If the initial point (pb) of a ray lies inside the polygon (region B), the number of intersections is odd. Jordan Curve Theorem for Polygons - Proof.svg
If the initial point (pa) of a ray (in red) lies outside a simple polygon (region A), the number of intersections of the ray and the polygon is even.
If the initial point (pb) of a ray lies inside the polygon (region B), the number of intersections is odd.

In computational geometry, the Jordan curve theorem can be used for testing whether a point lies inside or outside a simple polygon. [16] [17] [18]

From a given point, trace a ray that does not pass through any vertex of the polygon (all rays but a finite number are convenient). Then, compute the number n of intersections of the ray with an edge of the polygon. Jordan curve theorem proof implies that the point is inside the polygon if and only if n is odd.

Computational aspects

Adler, Daskalakis and Demaine [19] prove that a computational version of Jordan’s theorem is PPAD-complete. As a corollary, they show that Jordan's theorem implies the Brouwer fixed-point theorem. This complements the earlier result by Maehara, that Brouwer's fixed point theorem implies Jordan's theorem. [20]

See also

Notes

    1. 1 2 Jordan (1887).
    2. Kline, J. R. (1942). "What is the Jordan curve theorem?". American Mathematical Monthly. 49: 281–286. doi:10.2307/2303093. MR   0006516.
    3. Hales, Thomas C. (2007). "Jordan's proof of the Jordan curve theorem" (PDF). From Insight to Proof: Festschrift in Honour of Andrzej Trybulec. Studies in Logic, Grammar and Rhetoric. University of Białystok. 10 (23).
    4. Maehara (1984), p. 641.
    5. Gale, David (December 1979). "The Game of Hex and the Brouwer Fixed-Point Theorem". The American Mathematical Monthly. 86 (10): 818–827. doi:10.2307/2320146. ISSN   0002-9890. JSTOR   2320146.
    6. Nguyen, Phuong; Cook, Stephen A. (2007). "The Complexity of Proving the Discrete Jordan Curve Theorem". 22nd Annual IEEE Symposium on Logic in Computer Science (LICS 2007). IEEE. pp. 245–256. arXiv: 1002.2954 . doi:10.1109/lics.2007.48. ISBN   978-0-7695-2908-0.
    7. Hales, Thomas C. (December 2007). "The Jordan Curve Theorem, Formally and Informally". The American Mathematical Monthly. 114 (10): 882–894. doi:10.1080/00029890.2007.11920481. ISSN   0002-9890. S2CID   887392.
    8. Nayar, Shree (Mar 1, 2021). "First Principles of Computer Vision: Segmenting Binary Images | Binary Images". YouTube .
    9. Šlapal, J (April 2004). "A digital analogue of the Jordan curve theorem". Discrete Applied Mathematics. 139 (1–3): 231–251. doi: 10.1016/j.dam.2002.11.003 . ISSN   0166-218X.
    10. Surówka, Wojciech (1993). "A discrete form of Jordan curve theorem". Annales Mathematicae Silesianae (7): 57–61. MR   1271184.
    11. Johnson, Dale M. (1977). "Prelude to dimension theory: the geometrical investigations of Bernard Bolzano". Archive for History of Exact Sciences. 17 (3): 262–295. doi:10.1007/BF00499625. MR   0446838. See p. 285.
    12. OswaldVeblen  ( 1905 )
    13. Hales (2007b)
    14. Hales (2007b)
    15. A. Schoenflies (1924). "Bemerkungen zu den Beweisen von C. Jordan und Ch. J. de la Vallée Poussin". Jahresber. Deutsch. Math.-Verein. 33: 157–160.
    16. RichardCourant ( 1978 )
    17. "V. Topology". 1. Jordan curve theorem (PDF). Edinburg: University of Edinburgh. 1978. p. 267.
    18. "PNPOLY - Point Inclusion in Polygon Test - WR Franklin (WRF)". wrf.ecse.rpi.edu. Retrieved 2021-07-18.
    19. Adler, Aviv; Daskalakis, Constantinos; Demaine, Erik D. (2016). Chatzigiannakis, Ioannis; Mitzenmacher, Michael; Rabani, Yuval; Sangiorgi, Davide (eds.). "The Complexity of Hex and the Jordan Curve Theorem". 43rd International Colloquium on Automata, Languages, and Programming (ICALP 2016). Leibniz International Proceedings in Informatics (LIPIcs). Dagstuhl, Germany: Schloss Dagstuhl–Leibniz-Zentrum fuer Informatik. 55: 24:1–24:14. doi:10.4230/LIPIcs.ICALP.2016.24. ISBN   978-3-95977-013-2.
    20. Maehara (1984).

    Related Research Articles

    Brouwer's fixed-point theorem is a fixed-point theorem in topology, named after L. E. J. (Bertus) Brouwer. It states that for any continuous function mapping a nonempty compact convex set to itself, there is a point such that . The simplest forms of Brouwer's theorem are for continuous functions from a closed interval in the real numbers to itself or from a closed disk to itself. A more general form than the latter is for continuous functions from a nonempty convex compact subset of Euclidean space to itself.

    <span class="mw-page-title-main">Connected space</span> Topological space that is connected

    In topology and related branches of mathematics, a connected space is a topological space that cannot be represented as the union of two or more disjoint non-empty open subsets. Connectedness is one of the principal topological properties that are used to distinguish topological spaces.

    <span class="mw-page-title-main">Riemann mapping theorem</span>

    In complex analysis, the Riemann mapping theorem states that if is a non-empty simply connected open subset of the complex number plane which is not all of , then there exists a biholomorphic mapping from onto the open unit disk

    <span class="mw-page-title-main">Surface (topology)</span> Two-dimensional manifold

    In the part of mathematics referred to as topology, a surface is a two-dimensional manifold. Some surfaces arise as the boundaries of three-dimensional solid figures; for example, the sphere is the boundary of the solid ball. Other surfaces arise as graphs of functions of two variables; see the figure at right. However, surfaces can also be defined abstractly, without reference to any ambient space. For example, the Klein bottle is a surface that cannot be embedded in three-dimensional Euclidean space.

    <span class="mw-page-title-main">Cauchy's integral theorem</span> Theorem in complex analysis

    In mathematics, the Cauchy integral theorem in complex analysis, named after Augustin-Louis Cauchy, is an important statement about line integrals for holomorphic functions in the complex plane. Essentially, it says that if is holomorphic in a simply connected domain Ω, then for any simply closed contour in Ω, that contour integral is zero.

    <span class="mw-page-title-main">Winding number</span> Number of times a curve wraps around a point in the plane

    In mathematics, the winding number or winding index of a closed curve in the plane around a given point is an integer representing the total number of times that the curve travels counterclockwise around the point, i.e., the curve's number of turns. For certain open plane curves, the number of turns may be a non-integer. The winding number depends on the orientation of the curve, and it is negative if the curve travels around the point clockwise.

    In vector calculus, Green's theorem relates a line integral around a simple closed curve C to a double integral over the plane region D bounded by C. It is the two-dimensional special case of Stokes' theorem.

    <span class="mw-page-title-main">Algebraic curve</span> Curve defined as zeros of polynomials

    In mathematics, an affine algebraic plane curve is the zero set of a polynomial in two variables. A projective algebraic plane curve is the zero set in a projective plane of a homogeneous polynomial in three variables. An affine algebraic plane curve can be completed in a projective algebraic plane curve by homogenizing its defining polynomial. Conversely, a projective algebraic plane curve of homogeneous equation h(x, y, t) = 0 can be restricted to the affine algebraic plane curve of equation h(x, y, 1) = 0. These two operations are each inverse to the other; therefore, the phrase algebraic plane curve is often used without specifying explicitly whether it is the affine or the projective case that is considered.

    In mathematical analysis, a space-filling curve is a curve whose range reaches every point in a higher dimensional region, typically the unit square. Because Giuseppe Peano (1858–1932) was the first to discover one, space-filling curves in the 2-dimensional plane are sometimes called Peano curves, but that phrase also refers to the Peano curve, the specific example of a space-filling curve found by Peano.

    In mathematics, a Cohen–Macaulay ring is a commutative ring with some of the algebro-geometric properties of a smooth variety, such as local equidimensionality. Under mild assumptions, a local ring is Cohen–Macaulay exactly when it is a finitely generated free module over a regular local subring. Cohen–Macaulay rings play a central role in commutative algebra: they form a very broad class, and yet they are well understood in many ways.

    <span class="mw-page-title-main">Minkowski addition</span> Sums vector sets A and B by adding each vector in A to each vector in B

    In geometry, the Minkowski sum of two sets of position vectors A and B in Euclidean space is formed by adding each vector in A to each vector in B:

    <span class="mw-page-title-main">Simple polygon</span> Shape bounded by non-intersecting line segments

    In geometry, a simple polygon is a polygon that does not intersect itself and has no holes. That is, it is a piecewise-linear Jordan curve consisting of finitely many line segments. These polygons include as special cases the convex polygons, star-shaped polygons, and monotone polygons.

    In mathematics, Carathéodory's theorem is a theorem in complex analysis, named after Constantin Carathéodory, which extends the Riemann mapping theorem. The theorem, first proved in 1913, states that any conformal mapping sending the unit disk to some region in the complex plane bounded by a Jordan curve extends continuously to a homeomorphism from the unit circle onto the Jordan curve. The result is one of Carathéodory's results on prime ends and the boundary behaviour of univalent holomorphic functions.

    <span class="mw-page-title-main">Integer lattice</span> Lattice group in Euclidean space whose points are integer n-tuples

    In mathematics, the n-dimensional integer lattice, denoted , is the lattice in the Euclidean space whose lattice points are n-tuples of integers. The two-dimensional integer lattice is also called the square lattice, or grid lattice. is the simplest example of a root lattice. The integer lattice is an odd unimodular lattice.

    <span class="mw-page-title-main">Euclidean plane</span> Geometric model of the planar projection of the physical universe

    In mathematics, a Euclidean plane is a Euclidean space of dimension two, denoted or . It is a geometric space in which two real numbers are required to determine the position of each point. It is an affine space, which includes in particular the concept of parallel lines. It has also metrical properties induced by a distance, which allows to define circles, and angle measurement.

    In mathematics, the Schoenflies problem or Schoenflies theorem, of geometric topology is a sharpening of the Jordan curve theorem by Arthur Schoenflies. For Jordan curves in the plane it is often referred to as the Jordan–Schoenflies theorem.

    <span class="mw-page-title-main">Stokes' theorem</span> Theorem in vector calculus

    Stokes' theorem, also known as the Kelvin–Stokes theorem after Lord Kelvin and George Stokes, the fundamental theorem for curls or simply the curl theorem, is a theorem in vector calculus on . Given a vector field, the theorem relates the integral of the curl of the vector field over some surface, to the line integral of the vector field around the boundary of the surface. The classical theorem of Stokes can be stated in one sentence: The line integral of a vector field over a loop is equal to the surface integral of its curl over the enclosed surface. It is illustrated in the figure, where the direction of positive circulation of the bounding contour ∂Σ, and the direction n of positive flux through the surface Σ, are related by a right-hand-rule. For the right hand the fingers circulate along ∂Σ and the thumb is directed along n.

    <span class="mw-page-title-main">Inscribed square problem</span> Unsolved problem about inscribing a square in a Jordan curve

    The inscribed square problem, also known as the square peg problem or the Toeplitz' conjecture, is an unsolved question in geometry: Does every plane simple closed curve contain all four vertices of some square? This is true if the curve is convex or piecewise smooth and in other special cases. The problem was proposed by Otto Toeplitz in 1911. Some early positive results were obtained by Arnold Emch and Lev Schnirelmann. As of 2020, the general case remains open.

    <span class="mw-page-title-main">Two ears theorem</span> Every simple polygon with more than three vertices has at least two ears

    In geometry, the two ears theorem states that every simple polygon with more than three vertices has at least two ears, vertices that can be removed from the polygon without introducing any crossings. The two ears theorem is equivalent to the existence of polygon triangulations. It is frequently attributed to Gary H. Meisters, but was proved earlier by Max Dehn.

    In mathematics, a planar Riemann surface is a Riemann surface sharing the topological properties of a connected open subset of the Riemann sphere. They are characterized by the topological property that the complement of every closed Jordan curve in the Riemann surface has two connected components. An equivalent characterization is the differential geometric property that every closed differential 1-form of compact support is exact. Every simply connected Riemann surface is planar. The class of planar Riemann surfaces was studied by Koebe who proved in 1910, as a generalization of the uniformization theorem, that every such surface is conformally equivalent to either the Riemann sphere or the complex plane with slits parallel to the real axis removed.

    References