Pick's theorem

Last updated

Farey sunburst of order 6, with 1 interior (red) and 96 boundary (green) points giving an area of 1 +
.mw-parser-output .sfrac{white-space:nowrap}.mw-parser-output .sfrac.tion,.mw-parser-output .sfrac .tion{display:inline-block;vertical-align:-0.5em;font-size:85%;text-align:center}.mw-parser-output .sfrac .num,.mw-parser-output .sfrac .den{display:block;line-height:1em;margin:0 0.1em}.mw-parser-output .sfrac .den{border-top:1px solid}.mw-parser-output .sr-only{border:0;clip:rect(0,0,0,0);height:1px;margin:-1px;overflow:hidden;padding:0;position:absolute;width:1px}
96/2 - 1 = 48 Farey sunburst 6.svg
Farey sunburst of order 6, with 1 interior (red) and 96 boundary (green) points giving an area of 1 + 96/2 − 1 = 48

In geometry, Pick's theorem provides a formula for the area of a simple polygon with integer vertex coordinates, in terms of the number of integer points within it and on its boundary. The result was first described by Georg Alexander Pick in 1899. [2] It was popularized in English by Hugo Steinhaus in the 1950 edition of his book Mathematical Snapshots. [3] [4] It has multiple proofs, and can be generalized to formulas for certain kinds of non-simple polygons.

Contents

Formula

i = 7, b = 8, A = i +
b/2 - 1 = 10 Pick theorem simple.svg
i = 7, b = 8, A = i + b/2 − 1 = 10

Suppose that a polygon has integer coordinates for all of its vertices. Let be the number of integer points interior to the polygon, and let be the number of integer points on its boundary (including both vertices and points along the sides). Then the area of this polygon is: [5] [6] [7] [8]

The example shown has interior points and boundary points, so its area is square units.

Proofs

Via Euler's formula

One proof of this theorem involves subdividing the polygon into triangles with three integer vertices and no other integer points. One can then prove that each subdivided triangle has area exactly . Therefore, the area of the whole polygon equals half the number of triangles in the subdivision. After relating area to the number of triangles in this way, the proof concludes by using Euler's polyhedral formula to relate the number of triangles to the number of grid points in the polygon. [5]

Tiling of the plane by copies of a triangle with three integer vertices and no other integer points, as used in the proof of Pick's theorem Pick triangle tessellation.svg
Tiling of the plane by copies of a triangle with three integer vertices and no other integer points, as used in the proof of Pick's theorem

The first part of this proof shows that a triangle with three integer vertices and no other integer points has area exactly , as Pick's formula states. The proof uses the fact that all triangles tile the plane, with adjacent triangles rotated by 180° from each other around their shared edge. [9] For tilings by a triangle with three integer vertices and no other integer points, each point of the integer grid is a vertex of six tiles. Because the number of triangles per grid point (six) is twice the number of grid points per triangle (three), the triangles are twice as dense in the plane as the grid points. Any scaled region of the plane contains twice as many triangles (in the limit as the scale factor goes to infinity) as the number of grid points it contains. Therefore, each triangle has area , as needed for the proof. [5] A different proof that these triangles have area is based on the use of Minkowski's theorem on lattice points in symmetric convex sets. [10]

Subdivision of a grid polygon into special triangles Grid polygon triangulation.svg
Subdivision of a grid polygon into special triangles

This already proves Pick's formula for a polygon that is one of these special triangles. Any other polygon can be subdivided into special triangles: add non-crossing line segments within the polygon between pairs of grid points until no more line segments can be added. The only polygons that cannot be subdivided in this way are the special triangles considered above; therefore, only special triangles can appear in the resulting subdivision. Because each special triangle has area , a polygon of area will be subdivided into special triangles. [5]

The subdivision of the polygon into triangles forms a planar graph, and Euler's formula gives an equation that applies to the number of vertices, edges, and faces of any planar graph. The vertices are just the grid points of the polygon; there are of them. The faces are the triangles of the subdivision, and the single region of the plane outside of the polygon. The number of triangles is , so altogether there are faces. To count the edges, observe that there are sides of triangles in the subdivision. Each edge interior to the polygon is the side of two triangles. However, there are edges of triangles that lie along the polygon's boundary and form part of only one triangle. Therefore, the number of sides of triangles obeys the equation , from which one can solve for the number of edges, . Plugging these values for , , and into Euler's formula gives

Pick's formula is obtained by solving this linear equation for . [5] An alternative but similar calculation involves proving that the number of edges of the same subdivision is , leading to the same result. [11]

It is also possible to go the other direction, using Pick's theorem (proved in a different way) as the basis for a proof of Euler's formula. [6] [12]

Other proofs

Alternative proofs of Pick's theorem that do not use Euler's formula include the following.

Pick's theorem was included in a 1999 web listing of the "top 100 mathematical theorems", which later became used by Freek Wiedijk as a benchmark set to test the power of different proof assistants. As of 2021, Pick's theorem had been formalized and proven in only one of the ten proof assistants recorded by Wiedijk. [17]

Generalizations

i = 2, b = 12, h = 1, A = i +
b/2 + h - 1 = 8 Pick theorem hole.svg
i = 2, b = 12, h = 1, A = i + b/2 + h − 1 = 8

Generalizations to Pick's theorem to non-simple polygons are more complicated and require more information than just the number of interior and boundary vertices. [3] [18] For instance, a polygon with holes bounded by simple integer polygons, disjoint from each other and from the boundary, has area [19]

It is also possible to generalize Pick's theorem to regions bounded by more complex planar straight-line graphs with integer vertex coordinates, using additional terms defined using the Euler characteristic of the region and its boundary, [18] or to polygons with a single boundary polygon that can cross itself, using a formula involving the winding number of the polygon around each integer point as well as its total winding number. [3]

Reeve tetrahedra showing that Pick's theorem does not apply in higher dimensions Reeve tetrahedrons.svg
Reeve tetrahedra showing that Pick's theorem does not apply in higher dimensions

The Reeve tetrahedra in three dimensions have four integer points as vertices and contain no other integer points, but do not all have the same volume. Therefore, there does not exist an analogue of Pick's theorem in three dimensions that expresses the volume of a polyhedron as a function only of its numbers of interior and boundary points. [20] However, these volumes can instead be expressed using Ehrhart polynomials. [21] [22]

Several other mathematical topics relate the areas of regions to the numbers of grid points. Blichfeldt's theorem states that every shape can be translated to contain at least its area in grid points. [23] The Gauss circle problem concerns bounding the error between the areas and numbers of grid points in circles. [24] The problem of counting integer points in convex polyhedra arises in several areas of mathematics and computer science. [25] In application areas, the dot planimeter is a transparency-based device for estimating the area of a shape by counting the grid points that it contains. [26] The Farey sequence is an ordered sequence of rational numbers with bounded denominators whose analysis involves Pick's theorem. [27]

Another simple method for calculating the area of a polygon is the shoelace formula. It gives the area of any simple polygon as a sum of terms computed from the coordinates of consecutive pairs of its vertices. Unlike Pick's theorem, the shoelace formula does not require the vertices to have integer coordinates. [28]

Related Research Articles

<span class="mw-page-title-main">Area</span> Size of a two-dimensional surface

Area is the measure of a region's size on a surface. The area of a plane region or plane area refers to the area of a shape or planar lamina, while surface area refers to the area of an open surface or the boundary of a three-dimensional object. Area can be understood as the amount of material with a given thickness that would be necessary to fashion a model of the shape, or the amount of paint necessary to cover the surface with a single coat. It is the two-dimensional analogue of the length of a curve or the volume of a solid . Two different regions may have the same area ; by synecdoche, "area" sometimes is used to refer to the region, as in a "polygonal area".

In geometry, a polygon is a plane figure made up of line segments connected to form a closed polygonal chain.

<span class="mw-page-title-main">Prime number</span> Number divisible only by 1 or itself

A prime number is a natural number greater than 1 that is not a product of two smaller natural numbers. A natural number greater than 1 that is not prime is called a composite number. For example, 5 is prime because the only ways of writing it as a product, 1 × 5 or 5 × 1, involve 5 itself. However, 4 is composite because it is a product (2 × 2) in which both numbers are smaller than 4. Primes are central in number theory because of the fundamental theorem of arithmetic: every natural number greater than 1 is either a prime itself or can be factorized as a product of primes that is unique up to their order.

<span class="mw-page-title-main">Pythagorean triple</span> Integer side lengths of a right triangle

A Pythagorean triple consists of three positive integers a, b, and c, such that a2 + b2 = c2. Such a triple is commonly written (a, b, c), and a well-known example is (3, 4, 5). If (a, b, c) is a Pythagorean triple, then so is (ka, kb, kc) for any positive integer k. A primitive Pythagorean triple is one in which a, b and c are coprime (that is, they have no common divisor larger than 1). For example, (3, 4, 5) is a primitive Pythagorean triple whereas (6, 8, 10) is not. A triangle whose sides form a Pythagorean triple is called a Pythagorean triangle and is a right triangle.

<span class="mw-page-title-main">Quadrilateral</span> Polygon with four sides and four corners

In geometry a quadrilateral is a four-sided polygon, having four edges (sides) and four corners (vertices). The word is derived from the Latin words quadri, a variant of four, and latus, meaning "side". It is also called a tetragon, derived from Greek "tetra" meaning "four" and "gon" meaning "corner" or "angle", in analogy to other polygons. Since "gon" means "angle", it is analogously called a quadrangle, or 4-angle. A quadrilateral with vertices , , and is sometimes denoted as .

<span class="mw-page-title-main">Gauss–Bonnet theorem</span> Differential geometry theorem

In the mathematical field of differential geometry, the Gauss–Bonnet theorem is a fundamental formula which links the curvature of a surface to its underlying topology.

In mathematics, and more specifically in algebraic topology and polyhedral combinatorics, the Euler characteristic is a topological invariant, a number that describes a topological space's shape or structure regardless of the way it is bent. It is commonly denoted by .

<span class="mw-page-title-main">Equilateral triangle</span> Shape with three equal sides

In geometry, an equilateral triangle is a triangle in which all three sides have the same length. In the familiar Euclidean geometry, an equilateral triangle is also equiangular; that is, all three internal angles are also congruent to each other and are each 60°. It is also a regular polygon, so it is also referred to as a regular triangle.

<span class="mw-page-title-main">Cyclic quadrilateral</span> Quadrilateral whose vertices can all fall on a single circle

In Euclidean geometry, a cyclic quadrilateral or inscribed quadrilateral is a quadrilateral whose vertices all lie on a single circle. This circle is called the circumcircle or circumscribed circle, and the vertices are said to be concyclic. The center of the circle and its radius are called the circumcenter and the circumradius respectively. Other names for these quadrilaterals are concyclic quadrilateral and chordal quadrilateral, the latter since the sides of the quadrilateral are chords of the circumcircle. Usually the quadrilateral is assumed to be convex, but there are also crossed cyclic quadrilaterals. The formulas and properties given below are valid in the convex case.

<span class="mw-page-title-main">Isosceles triangle</span> Triangle with at least two sides congruent

In geometry, an isosceles triangle is a triangle that has two sides of equal length. Sometimes it is specified as having exactly two sides of equal length, and sometimes as having at least two sides of equal length, the latter version thus including the equilateral triangle as a special case. Examples of isosceles triangles include the isosceles right triangle, the golden triangle, and the faces of bipyramids and certain Catalan solids.

<span class="mw-page-title-main">Incenter</span> Center of the inscribed circle of a triangle

In geometry, the incenter of a triangle is a triangle center, a point defined for any triangle in a way that is independent of the triangle's placement or scale. The incenter may be equivalently defined as the point where the internal angle bisectors of the triangle cross, as the point equidistant from the triangle's sides, as the junction point of the medial axis and innermost point of the grassfire transform of the triangle, and as the center point of the inscribed circle of the triangle.

<span class="mw-page-title-main">Reuleaux triangle</span> Curved triangle with constant width

A Reuleaux triangle is a curved triangle with constant width, the simplest and best known curve of constant width other than the circle. It is formed from the intersection of three circular disks, each having its center on the boundary of the other two. Constant width means that the separation of every two parallel supporting lines is the same, independent of their orientation. Because its width is constant, the Reuleaux triangle is one answer to the question "Other than a circle, what shape can a manhole cover be made so that it cannot fall down through the hole?"

<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.

<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.

In the mathematical field of graph theory, Fáry's theorem states that any simple, planar graph can be drawn without crossings so that its edges are straight line segments. That is, the ability to draw graph edges as curves instead of as straight line segments does not allow a larger class of graphs to be drawn. The theorem is named after István Fáry, although it was proved independently by Klaus Wagner (1936), Fáry (1948), and Sherman K. Stein (1951).

<span class="mw-page-title-main">Barrow's inequality</span>

In geometry, Barrow's inequality is an inequality relating the distances between an arbitrary point within a triangle, the vertices of the triangle, and certain points on the sides of the triangle. It is named after David Francis Barrow.

<span class="mw-page-title-main">Handshaking lemma</span> Every graph has evenly many odd vertices

In graph theory, a branch of mathematics, the handshaking lemma is the statement that, in every finite undirected graph, the number of vertices that touch an odd number of edges is even. For example, if there is a party of people who shake hands, the number of people who shake an odd number of other people's hands is even. The handshaking lemma is a consequence of the degree sum formula, also sometimes called the handshaking lemma, according to which the sum of the degrees equals twice the number of edges in the graph. Both results were proven by Leonhard Euler (1736) in his famous paper on the Seven Bridges of Königsberg that began the study of graph theory.

<span class="mw-page-title-main">Reeve tetrahedra</span>

In geometry, the Reeve tetrahedra are a family of polyhedra in three-dimensional space with vertices at (0, 0, 0), (1, 0, 0), (0, 1, 0) and (1, 1, r) where r is a positive integer. They are named after John Reeve, who in 1957 used them to show that higher-dimensional generalizations of Pick's theorem do not exist.

<span class="mw-page-title-main">Blichfeldt's theorem</span> High-area shapes can shift to hold many grid points

Blichfeldt's theorem is a mathematical theorem in the geometry of numbers, stating that whenever a bounded set in the Euclidean plane has area , it can be translated so that it includes at least points of the integer lattice. Equivalently, every bounded set of area contains a set of points whose coordinates all differ by integers.

<span class="mw-page-title-main">Lexell's theorem</span> Characterizes spherical triangles with fixed base and area

In spherical geometry, Lexell's theorem holds that every spherical triangle with the same surface area on a fixed base has its apex on a small circle, called Lexell's circle or Lexell's locus, passing through each of the two points antipodal to the two base vertices.

References

  1. Kiradjiev, Kristian (October 2018). "Connecting the dots with Pick's theorem" (PDF). Mathematics Today. pp. 212–214.
  2. Pick, Georg (1899). "Geometrisches zur Zahlenlehre". Sitzungsberichte des deutschen naturwissenschaftlich-medicinischen Vereines für Böhmen "Lotos" in Prag. (Neue Folge). 19: 311–319. JFM   33.0216.01. CiteBank:47270
  3. 1 2 3 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.
  4. Steinhaus, H. (1950). Mathematical Snapshots. Oxford University Press. p. 76. MR   0036005.
  5. 1 2 3 4 5 Aigner, Martin; Ziegler, Günter M. (2018). "Three applications of Euler's formula: Pick's theorem". Proofs from THE BOOK (6th ed.). Springer. pp. 93–94. doi:10.1007/978-3-662-57265-8. ISBN   978-3-662-57265-8.
  6. 1 2 Wells, David (1991). "Pick's theorem". The Penguin Dictionary of Curious and Interesting Geometry. Penguin Books. pp. 183–184.
  7. 1 2 Beck, Matthias; Robins, Sinai (2015). "2.6 Pick's theorem". Computing the Continuous Discretely: Integer-Point Enumeration in Polyhedra. Undergraduate Texts in Mathematics (2nd ed.). Springer. pp. 40–43. doi:10.1007/978-1-4939-2969-6. ISBN   978-1-4939-2968-9. MR   3410115.
  8. 1 2 Ball, Keith (2003). "Chapter 2: Counting Dots". Strange Curves, Counting Rabbits, and Other Mathematical Explorations. Princeton University Press, Princeton, NJ. pp. 25–40. ISBN   0-691-11321-1. MR   2015451.
  9. Martin, George Edward (1982). Transformation geometry. Undergraduate Texts in Mathematics. Springer-Verlag. Theorem 12.1, page 120. doi:10.1007/978-1-4612-5680-9. ISBN   0-387-90636-3. MR   0718119.
  10. Ram Murty, M.; Thain, Nithum (2007). "Pick's theorem via Minkowski's theorem". The American Mathematical Monthly . 114 (8): 732–736. doi:10.1080/00029890.2007.11920465. JSTOR   27642309. MR   2354443. S2CID   38855683.
  11. Funkenbusch, W. W. (June–July 1974). "From Euler's formula to Pick's formula using an edge theorem". Classroom Notes. The American Mathematical Monthly . 81 (6): 647–648. doi:10.2307/2319224. JSTOR   2319224. MR   1537447.
  12. DeTemple, Duane; Robertson, Jack M. (March 1974). "The equivalence of Euler's and Pick's theorems". The Mathematics Teacher . 67 (3): 222–226. doi:10.5951/mt.67.3.0222. JSTOR   27959631. MR   0444503.
  13. Varberg, Dale E. (1985). "Pick's theorem revisited". The American Mathematical Monthly . 92 (8): 584–587. doi:10.2307/2323172. JSTOR   2323172. MR   0812105.
  14. Trainin, J. (November 2007). "An elementary proof of Pick's theorem". The Mathematical Gazette . 91 (522): 536–540. doi:10.1017/S0025557200182270. JSTOR   40378436. S2CID   124831432.
  15. Diaz, Ricardo; Robins, Sinai (1995). "Pick's formula via the Weierstrass -function". The American Mathematical Monthly . 102 (5): 431–437. doi:10.2307/2975035. JSTOR   2975035. MR   1327788.
  16. Brandolini, L.; Colzani, L.; Robins, S.; Travaglini, G. (2021). "Pick's theorem and convergence of multiple Fourier series". The American Mathematical Monthly . 128 (1): 41–49. doi:10.1080/00029890.2021.1839241. MR   4200451. S2CID   231624428.
  17. Wiedijk, Freek. "Formalizing 100 Theorems". Radboud University Institute for Computing and Information Sciences. Retrieved 2021-07-10.
  18. 1 2 Rosenholtz, Ira (1979). "Calculating surface areas from a blueprint". Mathematics Magazine . 52 (4): 252–256. doi:10.1080/0025570X.1979.11976797. JSTOR   2689425. MR   1572312.
  19. Sankar, P. V.; Krishnamurthy, E. V. (August 1978). "On the compactness of subsets of digital pictures". Computer Graphics and Image Processing. 8 (1): 136–143. doi:10.1016/s0146-664x(78)80021-5.
  20. Reeve, J. E. (1957). "On the volume of lattice polyhedra". Proceedings of the London Mathematical Society . Third Series. 7: 378–395. doi:10.1112/plms/s3-7.1.378. MR   0095452.
  21. Beck & Robins (2015), 3.6 "From the discrete to the continuous volume of a polytope", pp. 76–77
  22. Diaz, Ricardo; Robins, Sinai (1997). "The Ehrhart polynomial of a lattice polytope". Annals of Mathematics . Second Series. 145 (3): 503–518. doi:10.2307/2951842. JSTOR   2951842. MR   1454701.
  23. Olds, C. D.; Lax, Anneli; Davidoff, Giuliana P. (2000). "Chapter 9: A new principle in the geometry of numbers". The Geometry of Numbers. Anneli Lax New Mathematical Library. Vol. 41. Mathematical Association of America, Washington, DC. pp. 119–127. ISBN   0-88385-643-3. MR   1817689.
  24. Guy, Richard K. (2004). "F1: Gauß's lattice point problem". Unsolved Problems in Number Theory. Problem Books in Mathematics. Vol. 1 (3rd ed.). New York: Springer-Verlag. pp. 365–367. doi:10.1007/978-0-387-26677-0. ISBN   0-387-20860-7. MR   2076335.
  25. Barvinok, Alexander (2008). Integer Points In Polyhedra. Zurich Lectures in Advanced Mathematics. Zürich: European Mathematical Society. doi:10.4171/052. ISBN   978-3-03719-052-4. MR   2455889.
  26. Bellhouse, D. R. (1981). "Area estimation by point-counting techniques". Biometrics. 37 (2): 303–312. doi:10.2307/2530419. JSTOR   2530419. MR   0673040.
  27. Bruckheimer, Maxim; Arcavi, Abraham (1995). "Farey series and Pick's area theorem". The Mathematical Intelligencer . 17 (4): 64–67. doi:10.1007/BF03024792. MR   1365013. S2CID   55051527.
  28. Braden, Bart (1986). "The surveyor's area formula" (PDF). The College Mathematics Journal. 17 (4): 326–337. doi:10.2307/2686282. JSTOR   2686282.