Alexandrov's uniqueness theorem

Last updated

The Alexandrov uniqueness theorem is a rigidity theorem in mathematics, describing three-dimensional convex polyhedra in terms of the distances between points on their surfaces. It implies that convex polyhedra with distinct shapes from each other also have distinct metric spaces of surface distances, and it characterizes the metric spaces that come from the surface distances on polyhedra. It is named after Soviet mathematician Aleksandr Danilovich Aleksandrov, who published it in the 1940s. [1] [2] [3]

Contents

Statement of the theorem

The surface of any convex polyhedron in Euclidean space forms a metric space, in which the distance between two points is measured by the length of the shortest path from one point to the other along the surface. Within a single shortest path, distances between pairs of points equal the distances between corresponding points of a line segment of the same length; a path with this property is known as a geodesic. This property of polyhedral surfaces, that every pair of points is connected by a geodesic, is not true of many other metric spaces, and when it is true the space is called a geodesic space. The geodesic space formed from the surface of a polyhedron is called its development. [3]

Four regular hexagons can be folded and glued to form the surface of a regular octahedron. In this example the edges of the hexagons do not fall along edges of the octahedron. The same gluing pattern can also produce a non-convex polyhedron with 24 triangular faces. 4-hex octahedron.svg
Four regular hexagons can be folded and glued to form the surface of a regular octahedron. In this example the edges of the hexagons do not fall along edges of the octahedron. The same gluing pattern can also produce a non-convex polyhedron with 24 triangular faces.

The polyhedron can be thought of as being folded from a sheet of paper (a net for the polyhedron) and it inherits the same geometry as the paper: for every point p within a face of the polyhedron, a sufficiently small open neighborhood of p will have the same distances as a subset of the Euclidean plane. The same thing is true even for points on the edges of the polyhedron: they can be modeled locally as a Euclidean plane folded along a line and embedded into three-dimensional space, but the fold does not change the structure of shortest paths along the surface. However, the vertices of the polyhedron have a different distance structure: the local geometry of a polyhedron vertex is the same as the local geometry at the apex of a cone. Any cone can be formed from a flat sheet of paper with a wedge removed from it by gluing together the cut edges where the wedge was removed. The angle of the wedge that was removed is called the angular defect of the vertex; it is a positive number less than  2π. The defect of a polyhedron vertex can be measured by subtracting the face angles at that vertex from 2π. For instance, in a regular tetrahedron, each face angle is π/3, and there are three of them at each vertex, so subtracting them from 2π leaves a defect of π at each of the four vertices. Similarly, a cube has a defect of π/2 at each of its eight vertices. Descartes' theorem on total angular defect (a form of the Gauss–Bonnet theorem) states that the sum of the angular defects of all the vertices is always exactly 4π. In summary, the development of a convex polyhedron is geodesic, homeomorphic (topologically equivalent) to a sphere, and locally Euclidean except for a finite number of cone points whose angular defect sums to 4π. [3]

Alexandrov's theorem gives a converse to this description. It states that if a metric space is geodesic, homeomorphic to a sphere, and locally Euclidean except for a finite number of cone points of positive angular defect (necessarily summing to 4π), then there exists a convex polyhedron whose development is the given space. Moreover, this polyhedron is uniquely defined from the metric: any two convex polyhedra with the same surface metric must be congruent to each other as three-dimensional sets. [3]

Limitations

Two square sheets joined along their edges form a degenerate flat polyhedron, with four points of angular deficiency p at the four corners of the square. It can be inflated without stretching to form this non-convex shape, making the conical nature of the corners more apparent. Teabag.jpg
Two square sheets joined along their edges form a degenerate flat polyhedron, with four points of angular deficiency π at the four corners of the square. It can be inflated without stretching to form this non-convex shape, making the conical nature of the corners more apparent.

The polyhedron representing the given metric space may be degenerate: it may form a doubly-covered two-dimensional convex polygon (a dihedron) rather than a fully three-dimensional polyhedron. In this case, its surface metric consists of two copies of the polygon (its two sides) glued together along corresponding edges. [3] [6]

The regular icosahedron has the same surface metric as a non-convex deltahedron in which one of its five-triangle pyramids is pushed in rather than protruding Icosahedron.svg
The regular icosahedron has the same surface metric as a non-convex deltahedron in which one of its five-triangle pyramids is pushed in rather than protruding
Non-convex 24-sided tetrakis hexahedron with the same surface geometry as a regular octahedron Pyramid augmented cube.png
Non-convex 24-sided tetrakis hexahedron with the same surface geometry as a regular octahedron

Although Alexandrov's theorem states that there is a unique convex polyhedron whose surface has a given metric, it may also be possible for there to exist non-convex polyhedra with the same metric. An example is given by the regular icosahedron: if five of its triangles are removed, and are replaced by five congruent triangles forming an indentation into the polyhedron, the resulting surface metric stays unchanged. [7] This example uses the same creases for the convex and non-convex polyhedron, but that is not always the case. For instance, the surface of a regular octahedron can be re-folded along different creases into a non-convex polyhedron with 24 equilateral triangle faces, the Kleetope obtained by gluing square pyramids onto the squares of a cube. Six triangles meet at each additional vertex introduced by this refolding, so they have zero angular defect and remain locally Euclidean. In the illustration of an octahedron folded from four hexagons, these 24 triangles are obtained by subdividing each hexagon into six triangles. [5]

The development of any polyhedron can be described concretely by a collection of two-dimensional polygons together with instructions for gluing them together along their edges to form a metric space, and the conditions of Alexandrov's theorem for spaces described in this way are easily checked. However, the edges where two polygons are glued together could become flat and lie in the interior of faces of the resulting polyhedron, rather than becoming polyhedron edges. (For an example of this phenomenon, see the illustration of four hexagons glued to form an octahedron.) Therefore, even when the development is described in this way, it may not be clear what shape the resulting polyhedron has, what shapes its faces have, or even how many faces it has. Alexandrov's original proof does not lead to an algorithm for constructing the polyhedron (for instance by giving coordinates for its vertices) realizing the given metric space. In 2008, Bobenko and Izmestiev provided such an algorithm. [8] Their algorithm can approximate the coordinates arbitrarily accurately, in pseudo-polynomial time. [9]

One of the first existence and uniqueness theorems for convex polyhedra is Cauchy's theorem, which states that a convex polyhedron is uniquely determined by the shape and connectivity of its faces. Alexandrov's theorem strengthens this, showing that even if the faces are allowed to bend or fold, without stretching or shrinking, then their connectivity still determines the shape of the polyhedron. In turn, Alexandrov's proof of the existence part of his theorem uses a strengthening of Cauchy's theorem by Max Dehn to infinitesimal rigidity. [3]

An analogous result to Alexandrov's holds for smooth convex surfaces: a two-dimensional Riemannian manifold whose Gaussian curvature is everywhere positive and totals 4π can be represented uniquely as the surface of a smooth convex body in three dimensions. The uniqueness of this representation is a result of Stephan Cohn-Vossen from 1927, with some regularity conditions on the surface that were removed in later research. Its existence was proven by Alexandrov, using an argument involving limits of polyhedral metrics. [10] Aleksei Pogorelov generalized both these results, characterizing the developments of arbitrary convex bodies in three dimensions. [3]

Another result of Pogorelov on the geodesic metric spaces derived from convex polyhedra is a version of the theorem of the three geodesics: every convex polyhedron has at least three simple closed quasigeodesics. These are curves that are locally straight lines except when they pass through a vertex, where they are required to have angles of less than π on both sides of them. [11]

The developments of ideal hyperbolic polyhedra can be characterized in a similar way to Euclidean convex polyhedra: every two-dimensional manifold with uniform hyperbolic geometry and finite area, combinatorially equivalent to a finitely-punctured sphere, can be realized as the surface of an ideal polyhedron. [12]

Related Research Articles

<span class="mw-page-title-main">Dual polyhedron</span> Polyhedron associated with another by swapping vertices for faces

In geometry, every polyhedron is associated with a second dual structure, where the vertices of one correspond to the faces of the other, and the edges between pairs of vertices of one correspond to the edges between pairs of faces of the other. Such dual figures remain combinatorial or abstract polyhedra, but not all can also be constructed as geometric polyhedra. Starting with any given polyhedron, the dual of its dual is the original polyhedron.

<span class="mw-page-title-main">Polyhedron</span> 3D shape with flat faces, straight edges and sharp corners

In geometry, a polyhedron is a three-dimensional shape with flat polygonal faces, straight edges and sharp corners or vertices.

In geometry, a Platonic solid is a convex, regular polyhedron in three-dimensional Euclidean space. Being a regular polyhedron means that the faces are congruent regular polygons, and the same number of faces meet at each vertex. There are only five such polyhedra:

<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">Schläfli symbol</span> Notation that defines regular polytopes and tessellations

In geometry, the Schläfli symbol is a notation of the form that defines regular polytopes and tessellations.

In geometry, the (angular) defect means the failure of some angles to add up to the expected amount of 360° or 180°, when such angles in the Euclidean plane would. The opposite notion is the excess.

<span class="mw-page-title-main">Triaugmented triangular prism</span> Convex polyhedron with 14 triangle faces

The triaugmented triangular prism, in geometry, is a convex polyhedron with 14 equilateral triangles as its faces. It can be constructed from a triangular prism by attaching equilateral square pyramids to each of its three square faces. The same shape is also called the tetrakis triangular prism, tricapped trigonal prism, tetracaidecadeltahedron, or tetrakaidecadeltahedron; these last names mean a polyhedron with 14 triangular faces. It is an example of a deltahedron and of a Johnson solid.

<span class="mw-page-title-main">Net (polyhedron)</span> Edge-joined polygons which fold into a polyhedron

In geometry, a net of a polyhedron is an arrangement of non-overlapping edge-joined polygons in the plane which can be folded to become the faces of the polyhedron. Polyhedral nets are a useful aid to the study of polyhedra and solid geometry in general, as they allow for physical models of polyhedra to be constructed from material such as thin cardboard.

In mathematics, Hilbert's fourth problem in the 1900 list of Hilbert's problems is a foundational question in geometry. In one statement derived from the original, it was to find — up to an isomorphism — all geometries that have an axiomatic system of the classical geometry, with those axioms of congruence that involve the concept of the angle dropped, and `triangle inequality', regarded as an axiom, added.

<span class="mw-page-title-main">Dihedron</span> Polyhedron with 2 faces

A dihedron is a type of polyhedron, made of two polygon faces which share the same set of n edges. In three-dimensional Euclidean space, it is degenerate if its faces are flat, while in three-dimensional spherical space, a dihedron with flat faces can be thought of as a lens, an example of which is the fundamental domain of a lens space L(p,q). Dihedra have also been called bihedra, flat polyhedra, or doubly covered polygons.

Cauchy's theorem is a theorem in geometry, named after Augustin Cauchy. It states that convex polytopes in three dimensions with congruent corresponding faces must be congruent to each other. That is, any polyhedral net formed by unfolding the faces of the polyhedron onto a flat surface, together with gluing instructions describing which faces should be connected to each other, uniquely determines the shape of the original polyhedron. For instance, if six squares are connected in the pattern of a cube, then they must form a cube: there is no convex polyhedron with six square faces connected in the same way that does not have the same shape.

In geometry, a skew apeirohedron is an infinite skew polyhedron consisting of nonplanar faces or nonplanar vertex figures, allowing the figure to extend indefinitely without folding round to form a closed surface.

<span class="mw-page-title-main">Edge (geometry)</span> Line segment joining two adjacent vertices in a polygon or polytope

In geometry, an edge is a particular type of line segment joining two vertices in a polygon, polyhedron, or higher-dimensional polytope. In a polygon, an edge is a line segment on the boundary, and is often called a polygon side. In a polyhedron or more generally a polytope, an edge is a line segment where two faces meet. A segment joining two vertices while passing through the interior or exterior is not an edge but instead is called a diagonal.

In polyhedral combinatorics, a branch of mathematics, Steinitz's theorem is a characterization of the undirected graphs formed by the edges and vertices of three-dimensional convex polyhedra: they are exactly the 3-vertex-connected planar graphs. That is, every convex polyhedron forms a 3-connected planar graph, and every 3-connected planar graph can be represented as the graph of a convex polyhedron. For this reason, the 3-connected planar graphs are also known as polyhedral graphs.

<span class="mw-page-title-main">Density (polytope)</span> Number of windings of a polytope around its center of symmetry

In geometry, the density of a star polyhedron is a generalization of the concept of winding number from two dimensions to higher dimensions, representing the number of windings of the polyhedron around the center of symmetry of the polyhedron. It can be determined by passing a ray from the center to infinity, passing only through the facets of the polytope and not through any lower dimensional features, and counting how many facets it passes through. For polyhedra for which this count does not depend on the choice of the ray, and for which the central point is not itself on any facet, the density is given by this count of crossed facets.

In differential geometry the theorem of the three geodesics, also known as Lyusternik–Schnirelmann theorem, states that every Riemannian manifold with the topology of a sphere has at least three simple closed geodesics. The result can also be extended to quasigeodesics on a convex polyhedron, and to closed geodesics of reversible Finsler 2-spheres. The theorem is sharp: although every Riemannian 2-sphere contains infinitely many distinct closed geodesics, only three of them are guaranteed to have no self-intersections. For example, by a result of Morse if the lengths of three principal axes of an ellipsoid are distinct, but sufficiently close to each other, then the ellipsoid has only three simple closed geodesics.

Geometric Folding Algorithms: Linkages, Origami, Polyhedra is a monograph on the mathematics and computational geometry of mechanical linkages, paper folding, and polyhedral nets, by Erik Demaine and Joseph O'Rourke. It was published in 2007 by Cambridge University Press (ISBN 978-0-521-85757-4). A Japanese-language translation by Ryuhei Uehara was published in 2009 by the Modern Science Company (ISBN 978-4-7649-0377-7).

<span class="mw-page-title-main">Ideal polyhedron</span> Shape in hyperbolic geometry

In three-dimensional hyperbolic geometry, an ideal polyhedron is a convex polyhedron all of whose vertices are ideal points, points "at infinity" rather than interior to three-dimensional hyperbolic space. It can be defined as the convex hull of a finite set of ideal points. An ideal polyhedron has ideal polygons as its faces, meeting along lines of the hyperbolic space.

Convex Polyhedra is a book on the mathematics of convex polyhedra, written by Soviet mathematician Aleksandr Danilovich Aleksandrov, and originally published in Russian in 1950, under the title Выпуклые многогранники. It was translated into German by Wilhelm Süss as Konvexe Polyeder in 1958. An updated edition, translated into English by Nurlan S. Dairbekov, Semën Samsonovich Kutateladze and Alexei B. Sossinsky, with added material by Victor Zalgaller, L. A. Shor, and Yu. A. Volkov, was published as Convex Polyhedra by Springer-Verlag in 2005.

References

  1. Senechal gives a date of 1941, while O'Rourke lists 1948. See: Senechal, Marjorie (2013), Shaping Space: Exploring Polyhedra in Nature, Art, and the Geometrical Imagination, Springer, p. 62, ISBN   9780387927145 . O’Rourke, Joseph (2011), How to Fold It: The Mathematics of Linkages, Origami and Polyhedra, Cambridge University Press, p. 134, ISBN   9781139498548 .
  2. Alexandrov, A. D. (2006), Convex Polyhedra, Springer Monographs in Mathematics, Springer, ISBN   9783540263401 . Translated into English by N. S. Dairbekov, S. S. Kutateladze, and A. B. Sossinsky. The uniqueness part of the theorem is covered in Chapter 3, and the existence part is covered in Chapter 4.
  3. 1 2 3 4 5 6 7 Connelly, Robert (March 2006), "Convex Polyhedra by A. D. Alexandrov" (PDF), SIAM Review, 48 (1): 157–160, doi:10.1137/SIREAD000048000001000149000001, JSTOR   204537, archived from the original (PDF) on 2017-08-30
  4. Khramtcova, Elena; Langerman, Stefan (2017), "Which convex polyhedra can be made by gluing regular hexagons?", Abstracts of the 20th Japan Conference on Discrete and Computational Geometry, Graphs, and Games (PDF), pp. 63–64, archived from the original (PDF) on 2017-09-12, retrieved 2018-02-27
  5. 1 2 Rus, Jacob (2017), "Flowsnake Earth", in Swart, David; Séquin, Carlo H.; Fenyvesi, Kristóf (eds.), Proceedings of Bridges 2017: Mathematics, Art, Music, Architecture, Education, Culture, Phoenix, Arizona: Tessellations Publishing, pp. 237–244, ISBN   978-1-938664-22-9
  6. O'Rourke, Joseph (2010), On flat polyhedra deriving from Alexandrov's theorem, arXiv: 1007.2016 , Bibcode:2010arXiv1007.2016O
  7. Hartshorne, Robin (2000), "Example 44.2.3, the "punched-in icosahedron"", Geometry: Euclid and beyond, Undergraduate Texts in Mathematics, Springer-Verlag, New York, p. 442, doi:10.1007/978-0-387-22676-7, ISBN   0-387-98650-2, MR   1761093 .
  8. Bobenko, Alexander I.; Izmestiev, Ivan (2008), "Alexandrov's theorem, weighted Delaunay triangulations, and mixed volumes", Annales de l'Institut Fourier, 58 (2): 447–505, arXiv: math/0609447 , doi:10.5802/aif.2358, MR   2410380, S2CID   14879349
  9. Kane, Daniel; Price, Gregory N.; Demaine, Erik D. (2009), "A pseudopolynomial algorithm for Alexandrov's theorem", in Dehne, Frank; Gavrilova, Marina; Sack, Jörg-Rüdiger; Tóth, Csaba D. (eds.), Algorithms and data structures. 11th International Symposium, WADS 2009, Banff, Canada, August 21–23, 2009, Proceedings, Lecture Notes in Computer Science, vol. 5664, Berlin: Springer, pp. 435–446, arXiv: 0812.5030 , doi:10.1007/978-3-642-03367-4_38, ISBN   978-3-642-03366-7, MR   2550627, S2CID   453313
  10. Guan, Pengfei; Li, Yan Yan (1994), "The Weyl problem with nonnegative Gauss curvature", Journal of Differential Geometry, 39 (2): 331–342, doi: 10.4310/jdg/1214454874 , MR   1267893, S2CID   117698037
  11. Pogorelov, Aleksei V. (1949), "Quasi-geodesic lines on a convex surface", Matematicheskii Sbornik (in Russian), 25 (62): 275–306, MR   0031767
  12. Springborn, Boris (2020), "Ideal hyperbolic polyhedra and discrete uniformization", Discrete & Computational Geometry , 64 (1): 63–108, doi:10.1007/s00454-019-00132-8, MR   4110530, S2CID   203035718