Jump to content
 







Main menu
   


Navigation  



Main page
Contents
Current events
Random article
About Wikipedia
Contact us
Donate
 




Contribute  



Help
Learn to edit
Community portal
Recent changes
Upload file
 








Search  

































Create account

Log in
 









Create account
 Log in
 




Pages for logged out editors learn more  



Contributions
Talk
 



















Contents

   



(Top)
 


1 Definitions  





2 Properties  





3 Special cases  





4 Computational problems  





5 Related constructions  





6 See also  





7 References  





8 External links  














Simple polygon






العربية
Català
Deutsch
Ελληνικά
Español
فارسی
Français
ि
Bahasa Indonesia
Magyar

Polski
Português
Română
Русский
Slovenščina
Suomi
ி
Українська
Tiếng Vit


 

Edit links
 









Article
Talk
 

















Read
Edit
View history
 








Tools
   


Actions  



Read
Edit
View history
 




General  



What links here
Related changes
Upload file
Special pages
Permanent link
Page information
Cite this page
Get shortened URL
Download QR code
Wikidata item
 




Print/export  



Download as PDF
Printable version
 




In other projects  



Wikimedia Commons
 
















Appearance
   

 





This is a good article. Click here for more information.

From Wikipedia, the free encyclopedia
 


Two simple polygons (green and blue) and a self-intersecting polygon (red, in the lower right, not simple)

Ingeometry, 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.

The sum of external angles of a simple polygon is . Every simple polygon with sides can be triangulatedby of its diagonals, and by the art gallery theorem its interior is visible from some of its vertices.

Simple polygons are commonly seen as the input to computational geometry problems, including point in polygon testing, area computation, the convex hull of a simple polygon, triangulation, and Euclidean shortest paths.

Other constructions in geometry related to simple polygons include Schwarz–Christoffel mapping, used to find conformal maps involving simple polygons, polygonalization of point sets, constructive solid geometry formulas for polygons, and visibility graphs of polygons.

Definitions[edit]

Parts of a simple polygon

A simple polygon is a closed curve in the Euclidean plane consisting of straight line segments, meeting end-to-end to form a polygonal chain.[1] Two line segments meet at every endpoint, and there are no other points of intersection between the line segments. No proper subset of the line segments has the same properties.[2] The qualifier simple is sometimes omitted, with the word polygon assumed to mean a simple polygon.[3]

The line segments that form a polygon are called its edgesorsides. An endpoint of a segment is called a vertex (plural: vertices)[2] or a corner. Edges and vertices are more formal, but may be ambiguous in contexts that also involve the edges and vertices of a graph; the more colloquial terms sides and corners can be used to avoid this ambiguity.[4] The number of edges always equals the number of vertices.[2] Some sources allow two line segments to form a straight angle (180°),[5] while others disallow this, instead requiring collinear segments of a closed polygonal chain to be merged into a single longer side.[6] Two vertices are neighbors if they are the two endpoints of one of the sides of the polygon.[7]

Simple polygons are sometimes called Jordan polygons, because they are Jordan curves; the Jordan curve theorem can be used to prove that such a polygon divides the plane into two regions.[8] Indeed, Camille Jordan's original proof of this theorem took the special case of simple polygons (stated without proof) as its starting point.[9] The region inside the polygon (its interior) forms a bounded set[2] topologically equivalent to an open disk by the Jordan–Schönflies theorem,[10] with a finite but nonzero area.[11] The polygon itself is topologically equivalent to a circle,[12] and the region outside (the exterior) is an unbounded connected open set, with infinite area.[11] Although the formal definition of a simple polygon is typically as a system of line segments, it is also possible (and common in informal usage) to define a simple polygon as a closed set in the plane, the union of these line segments with the interior of the polygon.[2]

Adiagonal of a simple polygon is any line segment that has two polygon vertices as its endpoints, and that otherwise is entirely interior to the polygon.[13]

Properties[edit]

The internal angle of a simple polygon, at one of its vertices, is the angle spanned by the interior of the polygon at that vertex. A vertex is convex if its internal angle is less than (a straight angle, 180°) and concave if the internal angle is greater than . If the internal angle is , the external angle at the same vertex is defined to be its supplement , the turning angle from one directed side to the next. The external angle is positive at a convex vertex or negative at a concave vertex. For every simple polygon, the sum of the external angles is (one full turn, 360°). Thus the sum of the internal angles, for a simple polygon with sides is .[14]

A triangulated polygon with 11 vertices: 11 sides and 8 diagonals form 9 triangles.

Every simple polygon can be partitioned into non-overlapping triangles by a subset of its diagonals. When the polygon has sides, this produces triangles, separated by diagonals. The resulting partition is called a polygon triangulation.[8] The shape of a triangulated simple polygon can be uniquely determined by the internal angles of the polygon and by the cross-ratios of the quadrilaterals formed by pairs of triangles that share a diagonal.[15]

According to the two ears theorem, every simple polygon that is not a triangle has at least two ears, vertices whose two neighbors are the endpoints of a diagonal.[8] A related theorem states that every simple polygon that is not a convex polygon has a mouth, a vertex whose two neighbors are the endpoints of a line segment that is otherwise entirely exterior to the polygon. The polygons that have exactly two ears and one mouth are called anthropomorphic polygons.[16]

This 42-vertex polygonal art gallery is entirely visible from cameras placed at the 4 marked vertices.

According to the art gallery theorem, in a simple polygon with vertices, it is always possible to find a subset of at most of the vertices with the property that every point in the polygon is visible from one of the selected vertices. This means that, for each point in the polygon, there exists a line segment connecting to a selected vertex, passing only through interior points of the polygon. One way to prove this is to use graph coloring on a triangulation of the polygon: it is always possible to color the vertices with three colors, so that each side or diagonal in the triangulation has two endpoints of different colors. Each point of the polygon is visible to a vertex of each color, for instance one of the three vertices of the triangle containing that point in the chosen triangulation. One of the colors is used by at most of the vertices, proving the theorem.[17]

Special cases[edit]

Every convex polygon is a simple polygon. Another important class of simple polygons are the star-shaped polygons, the polygons that have a point (interior or on their boundary) from which every point is visible.[2]

Amonotone polygon, with respect to a straight line , is a polygon for which every line perpendicular to intersects the interior of the polygon in a connected set. Equivalently, it is a polygon whose boundary can be partitioned into two monotone polygonal chains, subsequences of edges whose vertices, when projected perpendicularly onto , have the same order along as they do in the chain.[18]

Computational problems[edit]

To test whether a point is inside the polygon, construct any ray emanating from the point and count its intersections with the polygon. If it crosses only interior points of edges, an odd number of times, the point lies inside the polygon; if even, the point lies outside. Rays through polygon vertices or containing its edges need special care.[19]
A simple polygon (interior shaded blue) and its convex hull (surrounding both blue and yellow regions)

Incomputational geometry, several important computational tasks involve inputs in the form of a simple polygon.

Other computational problems studied for simple polygons include constructions of the longest diagonal or the longest line segment interior to a polygon,[13] of the convex skull (the largest convex polygon within the given simple polygon),[29][30] and of various one-dimensional skeletons approximating its shape, including the medial axis[31] and straight skeleton.[32] Researchers have also studied producing other polygons from simple polygons using their offset curves,[33] unions and intersections,[11] and Minkowski sums,[34] but these operations do not always produce a simple polygon as their result. They can be defined in a way that always produces a two-dimensional region, but this requires careful definitions of the intersection and difference operations in order to avoid creating one-dimensional features or isolated points.[11]

Related constructions[edit]

According to the Riemann mapping theorem, any simply connected open subset of the plane can be conformally mapped onto a disk. Schwarz–Christoffel mapping provides a method to explicitly construct a map from a disk to any simple polygon using specified vertex angles and pre-images of the polygon vertices on the boundary of the disk. These pre-vertices are typically computed numerically.[35]

The black polygon is the shortest loop connecting every red dot, a solution to the traveling salesperson problem.

Every finite set of points in the plane that does not lie on a single line can be connected to form the vertices of a simple polygon (allowing 180° angles); for instance, one such polygon is the solution to the traveling salesperson problem.[36] Connecting points to form a polygon in this way is called polygonalization.[37]

Every simple polygon can be represented by a formula in constructive solid geometry that constructs the polygon (as a closed set including the interior) from unions and intersections of half-planes, with each side of the polygon appearing once as a half-plane in the formula. Converting an -sided polygon into this representation can be performed in time .[38]

The visibility graph of a simple polygon connects its vertices by edges representing the sides and diagonals of the polygon.[3] It always contains a Hamiltonian cycle, formed by the polygon sides. The computational complexity of reconstructing a polygon that has a given graph as its visibility graph, with a specified Hamiltonian cycle as its cycle of sides, remains an open problem.[39]

See also[edit]

References[edit]

  1. ^ Milnor, John W. (1950). "On the total curvature of knots". Annals of Mathematics. 2nd Series. 52: 248–257. doi:10.2307/1969467.
  • ^ a b c d e f Preparata, Franco P.; Shamos, Michael Ian (1985). Computational Geometry: An Introduction. Texts and Monographs in Computer Science. Springer-Verlag. p. 18. doi:10.1007/978-1-4612-1098-6. ISBN 978-1-4612-1098-6.
  • ^ a b Everett, Hazel; Corneil, Derek (1995). "Negative results on characterizing visibility graphs". Computational Geometry: Theory & Applications. 5 (2): 51–63. doi:10.1016/0925-7721(95)00021-Z. MR 1353288.
  • ^ Aronov, Boris; Seidel, Raimund; Souvaine, Diane (1993). "On compatible triangulations of simple polygons". Computational Geometry: Theory & Applications. 3 (1): 27–35. doi:10.1016/0925-7721(93)90028-5. MR 1222755.
  • ^ Malkevitch, Joseph (2016). "Are precise definitions a good idea?". AMS Feature Column. American Mathematical Society.
  • ^ a b McCallum, Duncan; Avis, David (1979). "A linear algorithm for finding the convex hull of a simple polygon". Information Processing Letters. 9 (5): 201–206. doi:10.1016/0020-0190(79)90069-3. MR 0552534.
  • ^ de Berg, M.; van Kreveld, M.; Overmars, Mark; Schwarzkopf, O. (2008). Computational Geometry: Algorithms and Applications (3rd ed.). Springer. p. 58. doi:10.1007/978-3-540-77974-2.
  • ^ a b c Meisters, G. H. (1975). "Polygons have ears". The American Mathematical Monthly. 82 (6): 648–651. doi:10.2307/2319703. JSTOR 2319703. MR 0367792.
  • ^ 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.
  • ^ Thomassen, Carsten (1992). "The Jordan-Schönflies theorem and the classification of surfaces". The American Mathematical Monthly. 99 (2): 116–130. doi:10.1080/00029890.1992.11995820. JSTOR 2324180. MR 1144352.
  • ^ a b c d Margalit, Avraham; Knott, Gary D. (1989). "An algorithm for computing the union, intersection or difference of two polygons". Computers & Graphics. 13 (2): 167–183. doi:10.1016/0097-8493(89)90059-9.
  • ^ a b Niven, Ivan; Zuckerman, H. S. (1967). "Lattice points and polygonal area". The American Mathematical Monthly. 74 (10): 1195–1200. doi:10.1080/00029890.1967.12000095. JSTOR 2315660. MR 0225216.
  • ^ a b Aggarwal, Alok; Suri, Subhash (1990). "Computing the longest diagonal of a simple polygon". Information Processing Letters. 35 (1): 13–18. doi:10.1016/0020-0190(90)90167-V. MR 1069001.
  • ^ Richmond, Bettina; Richmond, Thomas (2023). A Discrete Transition to Advanced Mathematics. Pure and Applied Undergraduate Texts. Vol. 63 (2nd ed.). American Mathematical Society. p. 421. ISBN 9781470472047.
  • ^ Snoeyink, Jack (1999). "Cross-ratios and angles determine a polygon". Discrete & Computational Geometry. 22 (4): 619–631. doi:10.1007/PL00009481. MR 1721028.
  • ^ Toussaint, Godfried (1991). "Anthropomorphic polygons". The American Mathematical Monthly. 98 (1): 31–35. doi:10.2307/2324033. JSTOR 2324033. MR 1083611.
  • ^ Fisk, S. (1978). "A short proof of Chvátal's watchman theorem". Journal of Combinatorial Theory, Series B. 24 (3): 374. doi:10.1016/0095-8956(78)90059-X.
  • ^ Preparata, Franco P.; Supowit, Kenneth J. (1981). "Testing a simple polygon for monotonicity". Information Processing Letters. 12 (4): 161–164. doi:10.1016/0020-0190(81)90091-0.
  • ^ Schirra, Stefan (2008). "How reliable are practical point-in-polygon strategies?" (PDF). In Halperin, Dan; Mehlhorn, Kurt (eds.). Algorithms – ESA 2008, 16th Annual European Symposium, Karlsruhe, Germany, September 15–17, 2008. Proceedings. Lecture Notes in Computer Science. Vol. 5193. Springer. pp. 744–755. doi:10.1007/978-3-540-87744-8_62.
  • ^ Snoeyink, Jack (2017). "Point Location" (PDF). In Toth, Csaba D.; O'Rourke, Joseph; Goodman, Jacob E. (eds.). Handbook of Discrete and Computational Geometry (3rd ed.). Chapman and Hall/CRC. pp. 1005–1023.
  • ^ Braden, Bart (1986). "The surveyor's area formula" (PDF). The College Mathematics Journal. 17 (4): 326–337. doi:10.2307/2686282. JSTOR 2686282. Archived from the original (PDF) on 2012-11-07.
  • ^ Grünbaum, Branko; Shephard, G. C. (February 1993). "Pick's theorem". The American Mathematical Monthly. 100 (2): 150–161. doi:10.2307/2323771. JSTOR 2323771. MR 1212401.
  • ^ Chazelle, Bernard (1991). "Triangulating a simple polygon in linear time". Discrete & Computational Geometry. 6 (5): 485–524. doi:10.1007/BF02574703. MR 1115104.
  • ^ Urrutia, Jorge (2000). "Art gallery and illumination problems". In Sack, Jörg-Rüdiger; Urrutia, Jorge (eds.). Handbook of Computational Geometry. Amsterdam: North-Holland. pp. 973–1027. doi:10.1016/B978-044482537-7/50023-1. ISBN 0-444-82537-1. MR 1746693.
  • ^ Aichholzer, Oswin; Mulzer, Wolfgang; Pilz, Alexander (2015). "Flip distance between triangulations of a simple polygon is NP-complete". Discrete & Computational Geometry. 54 (2): 368–389. arXiv:1209.0579. doi:10.1007/s00454-015-9709-7. MR 3372115.
  • ^ a b Ahn, Hee-Kap; Barba, Luis; Bose, Prosenjit; De Carufel, Jean-Lou; Korman, Matias; Oh, Eunjin (2016). "A linear-time algorithm for the geodesic center of a simple polygon". Discrete & Computational Geometry. 56 (4): 836–859. arXiv:1501.00561. doi:10.1007/s00454-016-9796-0. MR 3561791.
  • ^ a b Guibas, Leonidas; Hershberger, John; Leven, Daniel; Sharir, Micha; Tarjan, Robert E. (1987). "Linear-time algorithms for visibility and shortest path problems inside triangulated simple polygons". Algorithmica. 2 (2): 209–233. doi:10.1007/BF01840360. MR 0895445.
  • ^ El Gindy, Hossam; Avis, David (1981). "A linear algorithm for computing the visibility polygon from a point". Journal of Algorithms. 2 (2): 186–197. doi:10.1016/0196-6774(81)90019-5.
  • ^ Chang, J. S.; Yap, C.-K. (1986). "A polynomial solution for the potato-peeling problem". Discrete & Computational Geometry. 1 (2): 155–182. doi:10.1007/BF02187692. MR 0834056.
  • ^ Cabello, Sergio; Cibulka, Josef; Kynčl, Jan; Saumell, Maria; Valtr, Pavel (2017). "Peeling potatoes near-optimally in near-linear time". SIAM Journal on Computing. 46 (5): 1574–1602. arXiv:1406.1368. doi:10.1137/16M1079695. MR 3708542.
  • ^ Chin, Francis Y. L.; Snoeyink, Jack; Wang, Cao An (1999). "Finding the medial axis of a simple polygon in linear time". Discrete & Computational Geometry. 21 (3): 405–420. doi:10.1007/PL00009429. MR 1672988.
  • ^ Cheng, Siu-Wing; Mencel, Liam; Vigneron, Antoine (2016). "A faster algorithm for computing straight skeletons". ACM Transactions on Algorithms. 12 (3): 44:1–44:21. arXiv:1405.4691. doi:10.1145/2898961.
  • ^ Palfrader, Peter; Held, Martin (February 2015). "Computing mitered offset curves based on straight skeletons". Computer-Aided Design and Applications. 12 (4): 414–424. doi:10.1080/16864360.2014.997637.
  • ^ Oks, Eduard; Sharir, Micha (2006). "Minkowski sums of monotone and general simple polygons". Discrete & Computational Geometry. 35 (2): 223–240. doi:10.1007/s00454-005-1206-y. MR 2195052.
  • ^ Trefethen, Lloyd N.; Driscoll, Tobin A. (1998). "Schwarz–Christoffel mapping in the computer era". Proceedings of the International Congress of Mathematicians, Vol. III (Berlin, 1998). Documenta Mathematica. pp. 533–542. MR 1648186.
  • ^ Quintas, L. V.; Supnick, Fred (1965). "On some properties of shortest Hamiltonian circuits". The American Mathematical Monthly. 72 (9): 977–980. doi:10.2307/2313333. JSTOR 2313333. MR 0188872.
  • ^ Demaine, Erik D.; Fekete, Sándor P.; Keldenich, Phillip; Krupke, Dominik; Mitchell, Joseph S. B. (2022). "Area-optimal simple polygonalizations: the CG challenge 2019". ACM Journal of Experimental Algorithmics. 27: A2.4:1–12. doi:10.1145/3504000. hdl:1721.1/146480. MR 4390039.
  • ^ Dobkin, David; Guibas, Leonidas; Hershberger, John; Snoeyink, Jack (1993). "An efficient algorithm for finding the CSG representation of a simple polygon". Algorithmica. 10 (1): 1–23. doi:10.1007/BF01908629. MR 1230699.
  • ^ Ghosh, Subir Kumar; Goswami, Partha P. (2013). "Unsolved problems in visibility graphs of points, segments, and polygons". ACM Computing Surveys. 46 (2): 22:1–22:29. arXiv:1012.5187. doi:10.1145/2543581.2543589.
  • External links[edit]


    Retrieved from "https://en.wikipedia.org/w/index.php?title=Simple_polygon&oldid=1199869976"

    Category: 
    Types of polygons
    Hidden categories: 
    Good articles
    Articles with short description
    Short description is different from Wikidata
    Use mdy dates from July 2023
    Use list-defined references from July 2023
     



    This page was last edited on 28 January 2024, at 04:56 (UTC).

    Text is available under the Creative Commons Attribution-ShareAlike License 4.0; additional terms may apply. By using this site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc., a non-profit organization.



    Privacy policy

    About Wikipedia

    Disclaimers

    Contact Wikipedia

    Code of Conduct

    Developers

    Statistics

    Cookie statement

    Mobile view



    Wikimedia Foundation
    Powered by MediaWiki