Jordan curve theorem

Source: Wikipedia, the free encyclopedia.
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

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.

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

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

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

in 1911, resulting in the Jordan–Brouwer separation theorem.

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

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

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.
  • 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

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

.

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

weak Kőnig's lemma
over the system .

Application

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

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. ^ a b Jordan (1887).
  2. .
  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. 10 (23). University of Białystok.
  4. ^ Maehara (1984), p. 641.
  5. JSTOR 2320146
    .
  6. .
  7. .
  8. ^ Nayar, Shree (Mar 1, 2021). "First Principles of Computer Vision: Segmenting Binary Images | Binary Images". YouTube.
  9. ISSN 0166-218X
    .
  10. .
  11. . See p. 285.
  12. ^ Oswald Veblen (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. ^ Richard Courant (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. .
  20. ^ Maehara (1984).

References

External links