Fact-checked by Grok 2 weeks ago

Convex polytope

A convex polytope is a bounded of \mathbb{R}^d that can be defined either as the of a of points or as the of a finite number of half-spaces, with these two representations being equivalent in finite dimensions. This structure ensures that the polytope is , meaning the between any two points in the set lies entirely within it, and it possesses a finite number of extreme points (vertices), edges, faces, and higher-dimensional facets. Convex polytopes generalize familiar shapes such as polygons in two dimensions and in three dimensions, where the latter include regular forms like the Platonic solids. The combinatorial structure of convex polytopes is described by their face lattice, which enumerates vertices, edges, faces, and facets, satisfying relations like V - E + F = 2 for three-dimensional polyhedra, where V, E, and F denote the numbers of vertices, edges, and faces, respectively. Notable examples include the d- (with d+1 vertices), the (with $2^d vertices), and the (dual to the hypercube). These objects exhibit rich geometric properties, such as being simple (each vertex incident to exactly d edges) or simplicial (each facet a ), which influence their enumeration and decomposition. The study of convex polytopes traces back to antiquity, with early investigations into regular polyhedra by mathematicians, and saw significant modern development in the mid-20th century through works on combinatorial and geometric aspects. Branko Grünbaum's monograph Convex Polytopes provided a foundational treatment of their combinatorial theory, inspiring advances in face enumeration and upper bound theorems. Today, convex polytopes are pivotal in fields like , where optimization problems are solved over polyhedral feasible regions, and in , where they underlie toric varieties and Ehrhart theory for counting lattice points. Their applications extend to , including algorithms for polytope representation conversion between vertex and half-space descriptions.

History

Early developments

The study of convex polytopes originated in classical geometry with investigations into three-dimensional polyhedra, which served as precursors to the more general concept of polytopes. Around 250 BCE, Archimedes enumerated the thirteen semi-regular polyhedra, now known as Archimedean solids, in a now-lost treatise referenced by Pappus of Alexandria; these uniform polyhedra, with regular polygonal faces and identical vertex configurations, highlighted early combinatorial and geometric properties of convex bounded polyhedra in Euclidean space. A significant combinatorial advancement came in the 18th century with Leonhard Euler's 1752 publication of his polyhedron formula, which relates the vertices V, edges E, and faces F of a convex polyhedron by V - E + F = 2; this relation provided an early topological invariant applicable to convex polyhedra and foreshadowed broader insights into polytope structure. In 1794, contributed to the understanding of polyhedral volumes in his Éléments de géométrie, where he presented an independent proof of and developed methods for computing volumes of polyhedra through into pyramids and tetrahedra, emphasizing the role of convexity in ensuring decomposition without overlaps or gaps. During the 19th century, and advanced the theory of bodies, including polytopes, by focusing on their geometric properties and decompositions. In , Cauchy demonstrated that the dihedral angles of a are uniquely determined by its facial structure, establishing a rigidity theorem that underscored as essential for structural integrity. Steiner, in works from the , introduced symmetrization techniques and the parallel body formula, which decomposes the volume of a body's into intrinsic measures, further solidifying as a foundational property for analyzing shapes and their transformations. These efforts laid the groundwork for treating bodies as decomposable into simpler components, influencing subsequent geometric studies.

20th-century advancements

In the late 19th and early 20th centuries, Hermann Minkowski laid foundational work for abstract convex polytopes in higher dimensions through his development of the geometry of numbers. In his 1896 book Geometrie der Zahlen, Minkowski formalized the convex hull as the smallest convex set containing a given set of points, extending this concept rigorously to n-dimensional Euclidean space and emphasizing its role in bounding lattice points within convex bodies. He further established Minkowski's theorem on lattice polytopes, stating that any convex body in \mathbb{R}^d that is symmetric about the origin and has volume greater than $2^d must contain a non-zero lattice point in its interior or boundary. This theorem provided essential tools for understanding the distribution of lattice points in n-dimensional convex sets, bridging number theory and convex geometry. Building on these ideas, advancements in the theory of volumes for convex bodies emerged in the mid-20th century. Mixed volumes generalize the notion of volume for sums of convex bodies, and their role in the Brunn-Minkowski inequality, originally proved by Minkowski in 1896 for convex bodies, which asserts that for non-empty compact convex subsets K, L \subset \mathbb{R}^d, the volume satisfies |K + L|^{1/d} \geq |K|^{1/d} + |L|^{1/d}, with equality if and only if K and L are homothetic. This inequality underscored the subadditivity of volumes under Minkowski summation and became a cornerstone for in n dimensions. In 1967, Branko Grünbaum published his seminal monograph Convex Polytopes, which provided a comprehensive foundational treatment of their combinatorial theory, including detailed studies on face lattices, realizations of combinatorial types, and inequalities for face numbers. This work inspired significant advances, such as the upper bound theorem and efforts toward classifying polytopes. In the 1970s, Peter McMullen advanced the combinatorial structure of convex polytopes by formalizing the face lattice and its associated invariants. The face lattice of a convex polytope encodes the inclusion relations among its faces, from vertices to the full polytope itself, providing a poset structure that captures the topology and combinatorics in higher dimensions. McMullen's 1970 upper bound theorem determined the maximum number of k-faces for a d-dimensional polytope with v vertices, achieved by cyclic polytopes, and relied on the f-vector (f_0, f_1, \dots, f_{d-1}), where f_i counts the i-dimensional faces. For simplicial polytopes, he incorporated the Dehn-Sommerville relations, which impose linear equations on the f-vector such as \sum_{i=0}^d (-1)^i f_i = 0 (Euler's formula generalized), linking face counts across dimensions and enabling the g-theorem's characterization of possible f-vectors. The Schläfli symbol, introduced by Ludwig Schläfli in 1852, saw significant generalization and application to higher-dimensional regular polytopes during the , particularly through enumerative and symmetry studies. The symbol \{p_1, p_2, \dots, p_{n-1}\} recursively describes regular polytopes where each entry specifies the number of facets meeting at lower-dimensional elements, naturally extending to n dimensions beyond the classical Platonic solids. In four dimensions, Schläfli identified six regular 4-polytopes (Schläfli polytopes), including the \{3,3,3\}, 8-cell \{4,3,3\}, \{3,3,4\}, \{3,4,3\}, \{5,3,3\}, and \{3,3,5\}, whose constructions and symmetry groups were further explored in works like H.S.M. Coxeter's 1948 analysis, confirming no additional regular 4-polytopes exist. These developments highlighted that in each dimension greater than 4, there are exactly three regular convex polytopes: the regular simplex, the , and the .

Definitions and Basic Concepts

Core definition

A is a fundamental object in , defined as the of a of points in \mathbb{R}^d. This construction captures bounded regions with flat faces, such as polygons in two dimensions or polyhedra in three dimensions. Formally, given a V = \{v_1, \dots, v_n\} \subset \mathbb{R}^d, the P is P = \operatorname{conv}(V), where the operator is defined as \operatorname{conv}(V) = \left\{ \sum_{i=1}^n \lambda_i v_i \;\middle|\; \lambda_i \geq 0 \ \forall i, \ \sum_{i=1}^n \lambda_i = 1, \ v_i \in V \right\}. This set consists of all possible convex combinations of the points in V, ensuring that P is the smallest containing V. A key property of a convex polytope is its convexity: for any two points x, y \in P, the entire connecting them, parameterized as (1-t)x + t y for t \in [0,1], lies within P. Equivalently, a convex polytope can be realized as the bounded intersection of finitely many half-spaces in \mathbb{R}^d. In distinction from general convex sets, which may be unbounded or possess infinitely many points, convex polytopes are compact (closed and bounded) and have only finitely many points (vertices), which are a subset of the generating set V.

Bounded versus unbounded polytopes

In , bounded polytopes are compact convex sets that have finite extent in every direction, meaning they are closed and bounded in . These polytopes are equivalently characterized as the of a finite number of points, a property that ensures they possess only finitely many vertices, edges, and faces. Unbounded polyhedra, on the other hand, arise as intersections of finitely many half-spaces that do not result in a , allowing infinite extension in certain directions. A key structural feature of such unbounded polyhedra is their recession cone, which consists of all directions along which the polyhedron remains invariant under , effectively describing the unbounded rays emanating from any point within it. For instance, a single half-space in \mathbb{R}^d, defined by a such as \{x \in \mathbb{R}^d \mid a^\top x \leq b\}, serves as a classic example of an unbounded , where the recession is the set \{d \mid a^\top d \leq 0\}. Regarding terminology, the word "polytope" is frequently reserved for bounded cases in modern literature, distinguishing them from the broader class of convex polyhedra that may be unbounded; this convention is particularly prevalent in optimization contexts like linear programming, where unbounded feasible regions are common.

Representations

Vertex-based representation

A convex polytope P in \mathbb{R}^d admits a vertex-based representation, or V-representation, as the convex hull of a finite set of points known as its vertices or extreme points: P = \conv\{ v_1, \dots, v_n \} \subseteq \mathbb{R}^d, where each v_i is a point in \mathbb{R}^d and n is finite. This formulation defines P directly from its boundary-generating elements, with the vertices forming the minimal set whose convex combinations yield all points in P. The extreme points theorem characterizes this representation precisely for polytopes. Every point in P can be expressed as a convex combination of the vertices, and no vertex itself is a nontrivial of the other points in P. This follows from the and of P, ensuring that the vertices are exactly the extreme points, and distinguishes polytopes from more general convex sets by their finite extremal structure. Consequently, the V-representation provides a complete and non-redundant description of P via its n vertices. This approach offers advantages in scenarios requiring generation from discrete points, such as approximating shapes from sampled data. It is intuitive for computational purposes, facilitating direct manipulation of boundary points without implicit constraints. In , vertex representations enable efficient modeling and rendering of polyhedral objects by specifying coordinates for , edges, and faces. In optimization, it supports tasks like for solving linear programs over or computing minimal-norm points within simplices derived from . A fundamental invariant in this context is f_0(P), the number of , which quantifies the combinatorial complexity of P and influences algorithms for manipulation. This V-representation is equivalent to the half-space form in fully describing bounded sets but emphasizes generative points over bounding inequalities.

Half-space representation

A convex polytope P in \mathbb{R}^d admits a half-space representation as the of a finite number of closed half-spaces, expressed as P = \{ x \in \mathbb{R}^d \mid A x \leq b \}, where A is an m \times d real matrix and b is a real vector in \mathbb{R}^m, with m finite. This formulation captures the polytope as the solution set to a of linear inequalities, each defining a bounding half-space. Each row a_i^\top of A, paired with the corresponding entry b_i of b, specifies a \{ x \mid a_i^\top x = b_i \}, with the half-space \{ x \mid a_i^\top x \leq b_i \} containing P. The facets of P, which are the (d-1)-dimensional faces, arise from the inequalities that are tight (i.e., active) for some points in P, meaning equality holds on those facets while the polytope lies strictly on one side of the . Redundant inequalities, which do not contribute new facets, may be present but can be eliminated to obtain a minimal representation. This half-space representation is advantageous in optimization contexts, where polytopes often model feasible regions defined by linear constraints, as in problems. It provides a perspective to the vertex-based description, facilitating conversions between representations for computational purposes. For P to be bounded (a rather than an unbounded ), the representation must ensure no nonzero recession direction exists; specifically, the recession cone \{ x \mid A x \leq 0 \} must be trivial, containing only the origin. A common normalization assumes the origin lies in the interior of P, which implies strict feasibility of the inequalities and guarantees boundedness when the recession cone condition holds.

Equivalence between representations

A fundamental result in polyhedral establishes the between the and half-space representations for polytopes. Specifically, every bounded polytope in \mathbb{R}^d can be expressed both as the of a of vertices V = \{v_1, \dots, v_k\}, denoted \operatorname{conv}(V), and as the intersection of finitely many half-spaces \{x \in \mathbb{R}^d \mid A x \leq b\}, where A is an m \times d and b \in \mathbb{R}^m. This is known as the Minkowski–Weyl theorem. The vertices in the half-space representation \{x \mid A x \leq b\} correspond precisely to the basic feasible solutions of the linear system, obtained by solving subsystems of d linearly independent tight inequalities A_I x = b_I, where I is a subset of d rows of (A, b) with full rank. A proof sketch of the theorem utilizes Farkas' lemma, which characterizes the feasibility of linear inequalities via alternatives. To show that a vertex representation yields a half-space one, consider the boundedness implying compactness; the supporting hyperplanes at extreme points generate the finite inequalities defining the polytope. Conversely, the finite intersection ensures the feasible region has finitely many extreme points, recoverable as solutions to the tight systems above, with Farkas' lemma guaranteeing no unbounded rays in the bounded case. This equivalence has key implications: it guarantees the existence of conversion algorithms between representations, such as double description methods that enumerate vertices from facets or vice versa, enabling computational verification and manipulation of polytopes. For unbounded polyhedra, the equivalence extends if the recession cone—capturing the directions of unboundedness—is polyhedral, i.e., finitely generated as a conical hull of extreme rays.

Support function representation

The support function of a convex polytope P \subset \mathbb{R}^d provides a functional representation defined by h_P(u) = \sup \{ \langle x, u \rangle \mid x \in P \} for all directions u \in \mathbb{R}^d, where \langle \cdot, \cdot \rangle denotes the standard inner product. Since P is compact, the supremum is attained and equals the maximum inner product. This function fully determines the polytope, as P = \{ x \in \mathbb{R}^d \mid \langle x, u \rangle \leq h_P(u) \ \forall \, u \in \mathbb{R}^d \}. The h_P is , as it arises as the pointwise supremum of linear functions, and positively homogeneous of degree one, satisfying h_P(\lambda u) = \lambda h_P(u) for all \lambda \geq 0 and u \in \mathbb{R}^d. For a with finitely many vertices v_1, \dots, v_m, it takes the explicit form h_P(u) = \max_{i=1}^m \langle v_i, u \rangle, which renders h_P piecewise linear, with linear pieces corresponding to the normal cones at the vertices. This representation offers computational advantages in geometric operations. Notably, the of the Minkowski sum of two polytopes P and Q satisfies h_{P+Q}(u) = h_P(u) + h_Q(u) for all u \in \mathbb{R}^d, simplifying the analysis of sums without enumerating vertices or facets. Similarly, it facilitates the study of , where the polar dual P^\circ = \{ y \in \mathbb{R}^d \mid \langle y, x \rangle \leq 1 \ \forall \, x \in P \} admits a description in terms of the reciprocal behavior of h_P, aiding duality transformations in .

Examples

Low-dimensional cases

In one dimension, a convex polytope is simply the convex hull of two distinct points, forming a closed line segment. In two dimensions, convex polytopes are known as convex polygons, which are planar figures bounded by a finite chain of line segments such that no internal angle exceeds 180 degrees and all line segments lie on the boundary. A basic example is the triangle, a convex polygon with 3 vertices and 3 edges. Another simple case is the quadrilateral, featuring 4 vertices and 4 edges. Regular convex polygons, such as the pentagon with 5 equal sides and angles, illustrate symmetric cases where all vertices lie on a common circle. In three dimensions, convex polytopes take the form of convex polyhedra, bounded solids with flat polygonal faces, straight edges, and vertices. Prominent examples are the five Platonic solids, which are regular convex polyhedra with identical regular polygonal faces meeting the same number of times at each vertex: the (4 triangular faces), cube (6 square faces, 8 vertices, 12 edges), octahedron (8 triangular faces), dodecahedron (12 pentagonal faces), and icosahedron (20 triangular faces). Beyond these, the 13 Archimedean solids represent convex polyhedra composed of regular polygons of two or more types, arranged with identical vertex configurations, such as the or . A fundamental relation for convex polyhedra that are topologically equivalent to a is Euler's polyhedral formula:
V - E + F = 2,
where V is the number of vertices, E the number of edges, and F the number of faces; for instance, the satisfies $8 - 12 + 6 = 2.

Higher-dimensional cases

In dimensions four and higher, polytopes exhibit rapidly increasing combinatorial complexity, with the number of faces growing exponentially with the dimension, making visualization impossible but their abstract structures rich in and extremal properties. A fundamental example is the d-simplex in \mathbb{R}^d, defined as the of d+1 affinely independent points, serving as the d-dimensional analog of a or with the minimal number of vertices for a full-dimensional polytope. Among infinite families generalizing low-dimensional regulars, the n-dimensional (or orthoplex) is the of the $2npoints obtained by permuting coordinates of(\pm 1, 0, \dots, 0)in\mathbb{R}^n, featuring tetrahedral cells in 4D and embodying the [dual](/page/Dual) of the [hypercube](/page/Hypercube) with maximal [symmetry](/page/Symmetry) under the l_1-norm unit ball.[](https://www.heldermann-verlag.de/jgg/jgg01_05/jgg0414.pdf) Its vertex count of &#36;2n highlights the linear growth in vertices contrasting with face explosion in higher dimensions. The n-hypercube (or n-cube), with $2^n vertices corresponding to all binary coordinate combinations in \mathbb{R}^n, generalizes the cube and tessellates \mathbb{R}^n, possessing n \cdot 2^{n-1} edges and demonstrating exponential combinatorial growth, as each dimension doubles the facets. In four dimensions, notable examples include the hypersimplex \Delta_{k, n}, a (n-1)-dimensional polytope as the convex hull of all (0,1)-vectors in \mathbb{R}^n summing to k (with $1 \leq k \leq n-1), generalizing the simplex and appearing in combinatorial optimization with \binom{n}{k} vertices. The regular 24-cell, or octaplex, is a self-dual convex 4-polytope with 24 vertices, 96 edges, and 24 octahedral cells, unique among regulars for lacking a 3D analog. The 600-cell, with Schläfli symbol \{3,3,5\}, comprises 600 tetrahedral cells, 1200 triangular faces, and 120 vertices, while its dual the 120-cell has 120 dodecahedral cells, 720 vertices, 1200 pentagonal faces, and 600 vertices—representing the 4D icosahedral symmetry extreme. For extremal combinatorial properties, cyclic polytopes in dimension d with v vertices (v > d) achieve the maximum number of k-faces for $0 < k < d, constructed as the convex hull of points on the moment curve (t, t^2, \dots, t^d) for distinct t_i, as proven in the Upper Bound Theorem. This family underscores the boundary of polytope complexity, far exceeding simpler polytopes like simplices.

Properties

Faces and the face lattice

A face of a convex polytope P \subset \mathbb{R}^d is defined as the intersection of P with a supporting hyperplane H of P, where P lies entirely on one side of H. More precisely, such an intersection F = P \cap H is a k-face if it has dimension k, with the affine hull of F having dimension k. The empty set and P itself are improper faces, while the proper faces include vertices as 0-faces (extreme points of P), edges as 1-faces (line segments connecting vertices), and facets as (d-1)-faces (the bounding hyperplane sections of maximum dimension). Every point of P lies in the relative interior of a unique face, ensuring a partition of P into the relative interiors of its faces. The relative interior of a face F, denoted \operatorname{relint}(F), consists of all points x \in F such that there exists \epsilon > 0 with B_\epsilon(x) \cap \operatorname{aff}(F) \subseteq F, where B_\epsilon(x) is the open ball centered at x and \operatorname{aff}(F) is the affine hull of F. This excludes lower-dimensional faces contained in the boundary of F relative to its affine span, providing a topological interior within the subspace spanned by F. The collection of all faces of P, including improper ones, forms the face \Phi(P), a (poset) ordered by face inclusion: F \leq G if and only if F \subseteq G. This poset is a , with the meet of two faces F and G given by their F \cap G (itself a face) and the join F \vee G defined as the smallest face containing both (the of all faces of P that contain F \cup G). The minimal element is the , and the maximal element is P itself, making \Phi(P) finite, graded by face dimension, (every element covers atoms, the vertices), and coatomic. The f-vector of a d-dimensional convex polytope P encodes the combinatorial structure of its faces as the tuple f(P) = (f_0, f_1, \dots, f_{d-1}), where f_k is the number of k-faces of P. For example, in a 3-dimensional polytope like , f_0 = 8 (vertices), f_1 = 12 (edges), and f_2 = 6 (facets). This captures essential counting invariants of the face , with f_0 equaling the number of vertices and f_{d-1} the number of facets.

Topological properties

A convex polytope in \mathbb{R}^d is a compact with nonempty interior. Its interior is to the open d-dimensional ball \mathbb{B}^d, while its boundary is to the (d-1)-dimensional S^{d-1}. This homeomorphism follows from the fact that any nonempty open in is diffeomorphic to \mathbb{R}^d, which is in turn to the open ball. Each k-face of a polytope, for $0 \leq k \leq d, is itself a convex k- and thus homeomorphic to the closed k- \overline{\mathbb{B}}^k. As a consequence, the entire polytope is contractible, meaning it is equivalent to a point; this holds because convex sets admit a straight-line homotopy to any fixed interior point. Moreover, since it is homeomorphic to a ball, the polytope is simply connected: every loop within it can be continuously contracted to a point. The Euler-Poincaré formula captures a key topological of the . For a d-dimensional convex P with f_k denoting the number of k-faces (where f_d = 1 accounts for P itself), the alternating satisfies \sum_{k=0}^d (-1)^k f_k = 1. This the topological \chi(P) = 1, reflecting the contractibility of P. For the boundary complex, excluding the d-face, the \sum_{k=0}^{d-1} (-1)^k f_k the Euler characteristic of S^{d-1}, which is $1 + (-1)^{d-1}. Convex polytopes are star-shaped with respect to any point in their interior: for any interior point x_0 and any point y in the polytope, the line segment [x_0, y] lies entirely within the polytope. This property underscores their radial convexity from interior vantage points and reinforces their simply connected nature.

Combinatorial properties

The f-vector of a provides a basic combinatorial invariant that encodes the number of faces of each . For a d-dimensional P, the f-vector is f(P) = (f_0(P), f_1(P), \dots, f_{d-1}(P)), where f_k(P) denotes the number of k-dimensional faces of P. A central result in the of faces is the Upper Bound Theorem, established by Peter McMullen in 1970. This theorem asserts that, for fixed d \geq 3 and fixed number of vertices v = f_0 \geq d+1, the maximum possible value of f_k over all d-dimensional convex polytopes with v vertices is attained uniquely by the cyclic polytope C_d(v). The theorem provides explicit bounds on these maxima, highlighting the extremal role of cyclic polytopes in combinatorial polytope theory. McMullen's proof relies on shellability properties of polytope boundaries and has profound implications for bounding the complexity of polytopes. For simplicial polytopes—those in which every face is a —the entries of the f-vector are constrained by the Dehn-Sommerville relations, which consist of d linear equations on the f-vector, or equivalently the symmetry conditions h_i = h_{d-i} for i = 0, \dots, d on the h-vector. These relations, originally derived by Max Dehn in 1905 for polytopes of dimensions 4 and 5 and generalized by Duncan M. Y. Sommerville in 1927 to arbitrary dimensions, express dependencies among the f_k arising from the of the polytope's boundary. The relations hold for any simplicial d-polytope and provide essential tools for characterizing valid f-vectors. The h-vector offers a refined combinatorial closely tied to algebraic structures on . For a simplicial , the h-vector h(P) = (h_0(P), \dots, h_d(P)) is defined through the Hilbert series of the Stanley-Reisner ring of the 's boundary complex, where the numerator is \sum_{i=0}^d h_i(P) t^i / (1-t)^{d+1}. Introduced by in 1975, the h-vector transforms the f-vector via a linear and captures and other positivity properties central to the Upper Bound Theorem's proof for simplicial spheres. The toric ideal of the Stanley-Reisner ring further encodes the h-vector's combinatorial significance, linking it to the generators of the ideal in the over the facets. Stanley's framework revolutionized the study of polytope face enumeration by bridging and . Convex polytopes are classified combinatorially as simplicial or simple based on their face structures. A polytope is simplicial if every proper face is a simplex, meaning each facet has exactly d vertices and induces a (d-1)-simplex. In contrast, a polytope is simple if every vertex is incident to precisely d edges (or facets in the dual), so that the link of each vertex is a (d-1)-simplex. These properties are dual: the polar of a simplicial polytope is simple, and vice versa, preserving the face lattice under reversal. Examples include the simplex (both simplicial and simple) and the cube (simple but not simplicial). This dichotomy underlies many duality results in polytope theory.

Decompositions and Equivalence

Simplicial decompositions

A simplicial decomposition, or triangulation, of a convex polytope P in d-dimensional space is a partition of P into d-dimensional simplices such that the vertices of each simplex are vertices of P, the union of the simplices is exactly P, and any two simplices intersect at most in a common face (including the empty set). This ensures no additional (Steiner) points are introduced, preserving the combinatorial structure of the original polytope. Every convex polytope admits such a triangulation. The existence follows from inductive constructions on the dimension and the number of vertices, leveraging the convexity to ensure simplices fill the interior without overlaps or gaps. In a triangulation of a d-polytope with n vertices, the number of d-simplices is bounded above in terms of the face numbers; for instance, it is at most O(n^{\lceil d/2 \rceil}), with tighter bounds depending on the specific polytope structure, such as the number of facets. One standard method to construct a triangulation is the pulling triangulation, which proceeds by induction on dimension. For a base case in dimension 2, a convex polygon can be triangulated by choosing non-crossing diagonals from one vertex. In higher dimensions, select a vertex v, first obtain a triangulation of the boundary \partial P (which is a (d-1)-dimensional polytope), and then form new simplices by coning: for each (d-1)-simplex \sigma in the boundary triangulation, add the simplex \mathrm{conv}(v, \sigma). This "pulls" the triangulation from the boundary to the full polytope, ensuring the result is a valid triangulation using only original vertices. The process can be repeated if needed to refine coarser decompositions. Triangulations serve as a foundational tool for volume computation of convex polytopes, where the volume of P is the sum of the volumes of the constituent simplices, each of which has a closed-form expression via the determinant formula for simplex volume. This approach is particularly useful in numerical and symbolic computations, as it decomposes the problem into manageable linear algebra operations over the vertex coordinates.

Scissors congruence

In three-dimensional , two convex polyhedra are scissors congruent if one can be dissected into a finite number of smaller polyhedra that can be rigidly rearranged via isometries to form the other. This notion extends the classical idea of equidecomposability from two dimensions, where polygons of equal area are always scissors congruent by the Bolyai-Gerwien theorem, to higher dimensions where additional invariants are required. Hilbert's third problem, presented in 1900, inquired whether every pair of polyhedra with equal in are scissors congruent. Max Dehn resolved this negatively in 1901 by showing that a regular and a of the same cannot be dissected into congruent pieces of each other. Dehn achieved this by defining a new algebraic , the Dehn invariant, which sums contributions from each edge involving the edge length multiplied by a tensorial function of the at that edge; this remains unchanged under but takes different values for the and . The Dehn invariant provides a necessary condition for scissors congruence beyond mere volume equality, highlighting that dissections preserve not only measure but also angular and linear relations in a specific algebraic sense. In 1965, Jean-Pierre Sydler established the converse, proving that in three dimensions, two convex polyhedra are congruent if and only if they share the same volume and Dehn invariant. This result, known as the Dehn-Sydler theorem, fully characterizes congruence for three-dimensional polyhedra and has influenced subsequent work on equidecomposability in other geometries.

Computational Aspects

Constructing representations

Convex polytopes can be represented in two primary forms: the V-representation, consisting of vertices and possibly rays for unbounded polyhedra, and the H-representation, defined by a system of linear inequalities (half-spaces). Converting between these representations is fundamental for and optimization, enabling the analysis of polytope structure from either perspective. Vertex enumeration, which computes the V-representation from an H-representation, can be achieved using the double description method, originally developed by Motzkin, Raiffa, , and in 1953. This algorithm incrementally builds the set of extreme points and rays by processing inequalities one at a time, maintaining a pair of descriptions (inequalities and generators) at each step to ensure duality. An alternative approach is Fourier-Motzkin elimination, which successively eliminates variables from the inequality system to project onto lower dimensions, though it is less efficient due to in intermediate steps. Facet enumeration, the dual problem of computing the H-representation from a V-representation, involves calculating the of the given points (and rays). A widely used method is the beneath-beyond algorithm, which incrementally adds points to a current , first locating the "beneath" structure (visible facets from the new point) and then extending "beyond" to incorporate it, ensuring efficient updates in higher dimensions. This approach, refined in practical implementations, avoids redundant computations by leveraging the current facet structure. These conversion algorithms are output-sensitive, with running times typically bounded by O(f_{d-1} \log n) or similar, where n is the input size, d the , and f_{d-1} the number of facets (or elements), reflecting dependence on the polytope's combinatorial rather than worst-case . Practical software tools such as cddlib, which implements the double description method, and polymake, supporting both directions via beneath-beyond and other heuristics, facilitate these computations effectively in low to moderate dimensions. For unbounded polyhedra, the V-representation incorporates recession rays alongside vertices to capture the directions of unboundedness, ensuring the generators fully describe the polyhedron under the Minkowski-Weyl theorem; algorithms like double description naturally produce these rays during enumeration.

Volume computation

Computing the volume of a convex polytope is a fundamental problem in computational geometry, with methods varying by the polytope's representation, such as vertex-based (V-polytope) or half-space-based (H-polytope). For V-polytopes, a standard exact method involves triangulating the polytope into simplices and summing their individual volumes. A triangulation decomposes the polytope P into a collection of d-dimensional simplices \Delta_i such that P = \bigcup \Delta_i with disjoint interiors, where d is the dimension. The volume of each simplex \Delta with vertices v_0, v_1, \dots, v_d is given by \operatorname{vol}(\Delta) = \frac{1}{d!} \left| \det \begin{pmatrix} v_1 - v_0 & \cdots & v_d - v_0 \end{pmatrix} \right|, and the total volume is \operatorname{vol}(P) = \sum_i \operatorname{vol}(\Delta_i). This approach leverages simplicial decompositions and is exact when using precise arithmetic, though constructing the triangulation can be computationally intensive in high dimensions. For H-polytopes defined by inequalities, signed decomposition methods provide an alternative without explicit triangulation. The Lawrence-Varchenko method decomposes the polytope into signed simplices or uses generating functions over vertex cones to compute the volume exactly via integration formulas, often by triangulating the dual cone and applying inclusion-exclusion principles. For approximations, randomized sampling techniques, such as hit-and-run random walks, estimate the volume by , achieving polynomial-time approximation schemes independent of dimension. Exact volume computation poses significant challenges in high dimensions due to numerical instability and in the number of faces or vertices, particularly when vertices have coordinates. For polytopes with rational vertices, specialized software like integrale employs exact arithmetic over rational numbers to compute volumes as exact fractions, supporting dimensions up to around 10-12 in practice. These tools integrate or algorithms with libraries for multiprecision arithmetic to mitigate precision loss. In terms of , exact computation for convex polytopes is #P-hard in general, even for simple representations, due to the difficulty of enumerating combinatorial structures like . However, for fixed d, the problem is solvable in polynomial time relative to the input size, as the number of simplices in a triangulation is bounded polynomially.

Optimization problems

Convex polytopes play a central role in optimization, serving as the feasible regions for linear programming problems, where the objective is to minimize or maximize a linear function over a polyhedral set defined by linear inequalities. A standard linear programming problem is formulated as optimizing \mathbf{c}^\top \mathbf{x} subject to A \mathbf{x} \leq \mathbf{b} and \mathbf{x} \geq \mathbf{0}, where the feasible region is a convex polytope (or unbounded polyhedron) in the vertex representation, with vertices corresponding to basic feasible solutions. The optimal solution, if it exists, occurs at a vertex of the polytope, leveraging the convexity to ensure global optimality without local minima concerns. The simplex method, developed by in 1947, exploits this vertex structure by pivoting along edges of the polytope from one to an adjacent one, improving the objective value until optimality is reached. This algorithm is particularly effective for polytopes arising from practical constraints, though its worst-case exponential has motivated alternatives. For unbounded polyhedra, which extend polytopes, techniques like introducing slack variables to convert inequalities to equalities or the big-M method handle artificial variables to ensure boundedness during initialization. Interior-point methods, introduced by in 1984, provide a polynomial-time alternative by traversing the interior of the polytope using barrier functions and , avoiding vertex enumeration. These methods scale well for large-scale polytopes, achieving theoretical complexity of O(n^{3.5} L) iterations, where n is the and L the of the input. Applications in transportation networks and scheduling problems benefit from the polytope's convexity, enabling efficient modeling of as linear objectives over feasible sets of non-negative flows or assignments.

References

  1. [1]
    Polytope -- from Wolfram MathWorld
    A convex polytope may be defined as the convex hull of a finite set of points (which are always bounded), or as a bounded intersection of a finite set of half- ...
  2. [2]
    [PDF] CONVEX POLYTOPES
    A convex polytope P is defined to be the convex hull of any finite set of points in Ed. A ^-dimensional convex polytope P will be referred to, for brevity, as a.
  3. [3]
    [PDF] An Introduction to Convex Polytopes
    + Anxn is in fact a convex combination. The definition of a convex set expresses that any convex combination of two points from the set is again in the set ...
  4. [4]
    [PDF] 15 BASIC PROPERTIES OF CONVEX POLYTOPES - CSUN
    INTRODUCTION. Convex polytopes are fundamental geometric objects that have been investigated since antiquity.
  5. [5]
    [PDF] Combinatorial Aspects of Convex Polytopes - Margaret Bayer
    Aug 1, 1991 · 1.1 Introduction. A convex polyhedron is a subset of d that is the intersection of a finite number of closed halfspaces.
  6. [6]
    [PDF] Convex Polytopes - UC Davis Math
    A bounded convex polyhedron is called a convex polytope. Since most polyhedra under consideration will be convex, this adjective will usually be omitted.
  7. [7]
    Archimedes - Biography - MacTutor - University of St Andrews
    There are references to other works of Archimedes which are now lost. Pappus refers to a work by Archimedes on semi-regular polyhedra, Archimedes himself ...
  8. [8]
    Euler's polyhedron formula | plus.maths.org
    Jun 1, 2007 · Euler's formula tells us that V - E + F = 2; or, in words: the number of vertices, minus the number of edges, plus the number of faces, is equal to two.
  9. [9]
    The Flaw in Euler's Proof of His Polyhedral Formula - jstor
    The first new proof was given in. 1794 by Legendre in his popular textbook ?l?ments de g?om?trie [15]. In this paper we present Euler's proof along with ...
  10. [10]
    Augustin-Louis Cauchy (1789 - 1857) - Biography - MacTutor
    In addition to his heavy workload Cauchy undertook mathematical researches and he proved in 1811 that the angles of a convex polyhedron are determined by its ...
  11. [11]
    [PDF] The Steiner Formula for Minkowski Valuations
    The famous Steiner formula, dating back to the 19th century, expresses the volume of the parallel set of a convex body K at distance r ≥ 0 as a polynomial ...Missing: Jakob | Show results with:Jakob
  12. [12]
    The maximum numbers of faces of a convex polytope | Mathematika
    Feb 26, 2010 · In this paper we give a proof of the long-standing Upper-bound Conjecture for convex polytopes, which states that, for 1 ≤ j < d < v, the ...
  13. [13]
    Schläfli Symbol -- from Wolfram MathWorld
    A symbol of the form {p,q,r,...} used to describe regular polygons, polyhedra, and their higher-dimensional counterparts.Missing: 20th century
  14. [14]
    [PDF] Lecture 4 1 Introduction 2 Polyhedra and Polytopes
    Sep 4, 2014 · Theorem 2 (Representation of Bounded Polyhedra) A bounded polyhedron P is the set of all convex combinations of its vertices, and is therefore ...
  15. [15]
    [PDF] An introduction to convex and discrete geometry Lecture Notes
    Recall that we showed that a polytope is the convex hull of its vertices (Theorem 4.6) and the vertices are exactly the extreme points (Theorem 4.7). 6.1 ...
  16. [16]
    [PDF] Chapter 4 Polyhedra and Polytopes - UPenn CIS
    4.1 Polyhedra, H-Polytopes and V-Polytopes. There are two natural ways to define a convex polyhedron, A: (1) As the convex hull of a finite set of points. (2) ...
  17. [17]
    [PDF] Vertex Representations and their Applications in Computer Graphics
    Mar 30, 1998 · Abstract. The vertex representation, a new data structure for representing and manipulating orthogonal objects, is presented.Missing: polytope optimization
  18. [18]
    [PDF] THE MINIMUM EUCLIDEAN-NORM POINT IN A CONVEX POLYTOPE
    The main result in this section reduces linear programming to finding the mini- mum norm point in a (vertex-representation) simplex. Theorem 2.2. LP reduces ...
  19. [19]
    Lectures on Polytopes - SpringerLink
    This is an excellent book on convex polytopes written by a young and extremely active researcher.
  20. [20]
    None
    ### Summary of Farkas-Minkowski-Weyl Theorem for Convex Polyhedra
  21. [21]
    Convex Polygon -- from Wolfram MathWorld
    A planar polygon is convex if it contains all the line segments connecting any pair of its points. Thus, for example, a regular pentagon is convex.
  22. [22]
    Platonic Solid -- from Wolfram MathWorld
    The Platonic solids, also called the regular solids or regular polyhedra, are convex polyhedra with equivalent faces composed of congruent convex regular ...
  23. [23]
    Archimedean Solid -- from Wolfram MathWorld
    The 13 Archimedean solids are the convex polyhedra that have a similar arrangement of nonintersecting regular convex polygons of two or more different types.
  24. [24]
    Polyhedral Formula -- from Wolfram MathWorld
    A formula relating the number of polyhedron vertices V, faces F, and polyhedron edges E of a simply connected (i.e., genus 0) polyhedron (or polygon).
  25. [25]
    The maximum numbers of faces of a convex polytope - McMullen
    Dec 1, 1970 · In this paper we give a proof of the long-standing Upper-bound Conjecture for convex polytopes, which states that, for 1 ≤ j < d < v, ...
  26. [26]
    [PDF] Flexible Cross-Polytopes in the Euclidean 4-Space
    Abstract. It is shown that the examples presented 1998 by A. Walz are special cases of a more general class of flexible cross-polytopes in E4.
  27. [27]
    [PDF] COVERING THE CROSSPOLYTOPE BY EQUAL BALLS
    We determine the minimal radius of n = 2, d or 2d congruent balls, which cover the d-dimensional crosspolytope. 1. Introduction. Packings of n equal balls into ...
  28. [28]
    [PDF] Topological Properties of Hypercubes - Computer Science
    An n-cube graph is an undirected graph consisting of k = 2" vertices labeled from 0 to 2" - 1 and such that there is an edge between any two vertices if and ...
  29. [29]
    A combinatorial formula for the Ehrhart h⁎-vector of the hypersimplex
    In particular it is an integral polytope. The hypersimplex can be found in several algebraic and geometric contexts, for example, as a moment polytope for the ...
  30. [30]
    [PDF] A construction of magic 24-cells - arXiv
    The 24-cell is the convex regular 4-polytope which is composed of 24 octahedral cells. The 120-cell is the convex regular 4-polytope which is composed of 120.<|separator|>
  31. [31]
    Regular Polytopes - Harold Scott Macdonald Coxeter - Google Books
    H. S. M. Coxeter's book is the foremost book available on regular polyhedra, incorporating not only the ancient Greek work on the subject, but also the vast ...
  32. [32]
    The maximum numbers of faces of a convex polytope
    In this paper we give a proof of the long-standing Upper-bound. Conjecture for convex polytopes, which states that, for 1 < j < d < v, the maximum possible ...
  33. [33]
    [PDF] Chapter 9 Convex Polytopes
    The faces can be defined for a polyhedron P in the same way as for polytopes: F is a. face if there exists a hyperplane h such that F = P ∩ h and P ⊂ h+. For ...
  34. [34]
    [PDF] face numbers and subdivisions of convex - polytopes
    A d-dimensional polytope has faces of dimension 0 (vertices), 1 (edges), and so on, up to d~ 1. (facets). Write f; for the number of i-dimensional faces. The f- ...
  35. [35]
    [PDF] Counting Faces of Polytopes - University of Kentucky
    The face-vector of a d-dimensional polytope is f = (f0,f1,...,fd−1) ... Theorem (Dehn-Sommerville Relations). For every simple d-polytope, hi = hd−i ...Missing: lattice | Show results with:lattice
  36. [36]
    [PDF] from polytopes to enumeration - CIS UPenn
    ... meet or join for x and y. We can see that the face poset of a polytope contains the meet of any two faces, namely their intersection. What about the join?
  37. [37]
    [PDF] Basic convex analysis for optimization Andersen Ang - angms.science
    Jun 6, 2017 · • S ⊂ Rn is (open) convex set =⇒ S is homeomorphic to open ball B.<|control11|><|separator|>
  38. [38]
    [PDF] A Short Proof of Euler–Poincaré Formula - arXiv
    Sep 9, 2021 · to convex polytopes in every finite dimension, also known as the Euler–Poincaré Formula. We provide another short inductive proof of the ...
  39. [39]
    [PDF] 15 FACE NUMBERS OF POLYTOPES AND COMPLEXES
    (i) f is the f-vector of a simplicial complex;. (ii) f is a K-sequence. A simplicial complex is connected if its 1-skeleton is connected in the sense of graph ...
  40. [40]
    [PDF] Computational Geometry: Polytopes
    Apr 29, 2010 · Theorem 2.2. The polar of a simple polytope is simplicial and the polar of a simplicial polytope is simple.
  41. [41]
    [PDF] The Complexity of Finding Small Triangulations of Convex 3 ... - arXiv
    A triangulation of a d-dimensional convex polytope P is a set of d-simplices whose union is the polytope, their vertices are extreme points of P, and any two ...
  42. [42]
    [PDF] 16 SUBDIVISIONS AND TRIANGULATIONS OF POLYTOPES - CSUN
    Aug 7, 2017 · The flatness theorem for nonsymmetric convex bodies via the local theory of Banach spaces. ... Regular triangulations of convex polytopes.
  43. [43]
    [PDF] Regular Triangulations of Convex Polytopes - UC Davis Mathematics
    THEOREM 1. Let P be any simple d-polytope with n facets. Then there exists a simple (d+1)-polytope Q with n+1 facets, one of which is congruent to P, such that ...
  44. [44]
    [PDF] Exact Volume Computation for Polytopes: A Practical Study - Hal-Inria
    When P = {x|A x ≤ b} for some m × d real matrix A and m-vector b, the pair (A, b) is called a halfspace representation or simply H–representation of P. In this ...<|control11|><|separator|>
  45. [45]
    [PDF] Scissors Congruence - UChicago Math
    Sep 24, 2016 · This expository paper discusses Dehn's answer to Hilbert's third problem, and the stronger Dehn - Sydler Theorem. We start with the Wal ...
  46. [46]
    [PDF] Mathematical Problems
    A reprint of appears in Mathematical Developments Arising from Hilbert Problems edited by Felix ... 1900, pp. 253-297, and in Archiv der Mathematik und ...
  47. [47]
    Ueber raumgleiche Polyeder - EuDML
    Dehn, M.. "Ueber raumgleiche Polyeder." Nachrichten von der Gesellschaft der Wissenschaften zu Göttingen, Mathematisch-Physikalische Klasse 1900 (1900): ...Missing: Max PDF
  48. [48]
    Conditions nécessaires et suffisantes pour l'équivalence des ...
    Conditions nécessaires et suffisantes pour l'équivalence des polyèdres de l'espace euclidien à trois dimensions. Published: December 1965. Volume 40, pages 43– ...
  49. [49]
    [PDF] Fourier-Motzkin Elimination - UC Davis Mathematics
    Jun 23, 2011 · Weyl-Minkowski. Theorem: [Weyl-Minkowski] Every polytope is a polyhedron. Every bounded polyhedron is a polytope. This allows us to represent ...
  50. [50]
    [PDF] The Quickhull Algorithm for Convex Hulls - UF CISE
    This article presents a practical convex hull algorithm that combines the two-dimensional Quick- hull Algorithm with the general-dimension Beneath-Beyond ...
  51. [51]
    cddlib: Double description method for polyhedral representation ...
    This means that one can move back and forth between an inequality representation and a generator (i.e. vertex and ray) representation of a polyhedron with cdd.
  52. [52]
    [PDF] arXiv:2010.13147v1 [math.CO] 25 Oct 2020
    Oct 25, 2020 · The double description method (DDM) is one of the most popular algorithms ... If P is bounded then P is a convex polytope, i.e. it can be ...
  53. [53]
    [PDF] We present a method for computing exactly the volume of a convex ...
    This settles a question posed by Dyer and Frieze. 1. INTRODUCTION. We present a method for computing exactly the volume of a convex polytope given as the set ...
  54. [54]
    [PDF] Polytope volume in Normaliz
    Jul 6, 2022 · (3) volume by signed decomposition, the Lawrence algorithm: Normaliz computes a triangulation of the dual cone and converts it into a signed ...
  55. [55]
    Algorithm for finding the volume of a convex polytope - MathOverflow
    Oct 18, 2009 · There are polynomial time approximation schemes for volume of convex bodies independent of dimension, based on random walks within the body.Formula for volume of a convex polytope - MathOverflowSecondary polytope - convex geometry - MathOverflowMore results from mathoverflow.net
  56. [56]
    LattE - Computations with Polyhedra - UC Davis Mathematics
    LattE (Lattice point Enumeration) is a computer software dedicated to the problems of counting lattice points and integration inside convex polytopes.Missing: high dimensions
  57. [57]
    [PDF] Effective Lattice Point Counting in Rational Convex Polytopes
    Jul 3, 2003 · This paper discusses algorithms and software for the enumeration of all lattice points inside a rational convex polytope: we describe LattE, a ...
  58. [58]
    [PDF] On the complexity of computing the volume of a polyhedron
    Abstract. We show that computing the volume of a polyhedron given either as a list of facets or as a list of vertices is as hard as computing the permanent ...
  59. [59]
    [PDF] A Fast and Practical Method to Estimate Volumes of Convex Polytopes
    Dec 31, 2013 · Dyer et.al. [1] and Khachiyan [2, 3] proved respectively that exact volume computation is #P-hard, even for explicitly described poly-.