Combinatorial design

Last updated

Combinatorial design theory is the part of combinatorial mathematics that deals with the existence, construction and properties of systems of finite sets whose arrangements satisfy generalized concepts of balance and/or symmetry. These concepts are not made precise so that a wide range of objects can be thought of as being under the same umbrella. At times this might involve the numerical sizes of set intersections as in block designs, while at other times it could involve the spatial arrangement of entries in an array as in sudoku grids.

Contents

Combinatorial design theory can be applied to the area of design of experiments. Some of the basic theory of combinatorial designs originated in the statistician Ronald Fisher's work on the design of biological experiments. Modern applications are also found in a wide gamut of areas including finite geometry, tournament scheduling, lotteries, mathematical chemistry, mathematical biology, algorithm design and analysis, networking, group testing and cryptography. [1]

Example

The Fano plane Fano plane.svg
The Fano plane

Given a certain number n of people, is it possible to assign them to sets so that each person is in at least one set, each pair of people is in exactly one set together, every two sets have exactly one person in common, and no set contains everyone, all but one person, or exactly one person? The answer depends on n.

This has a solution only if n has the form q2 + q + 1. It is less simple to prove that a solution exists if q is a prime power. It is conjectured that these are the only solutions. It has been further shown that if a solution exists for q congruent to 1 or 2 mod 4, then q is a sum of two square numbers. This last result, the Bruck–Ryser theorem, is proved by a combination of constructive methods based on finite fields and an application of quadratic forms.

When such a structure does exist, it is called a finite projective plane; thus showing how finite geometry and combinatorics intersect. When q = 2, the projective plane is called the Fano plane.

History

Combinatorial designs date to antiquity, with the Lo Shu Square being an early magic square. One of the earliest datable application of combinatorial design is found in India in the book Brhat Samhita by Varahamihira, written around 587 AD, for the purpose of making perfumes using 4 substances selected from 16 different substances using a magic square. [2]

Combinatorial designs developed along with the general growth of combinatorics from the 18th century, for example with Latin squares in the 18th century and Steiner systems in the 19th century. Designs have also been popular in recreational mathematics, such as Kirkman's schoolgirl problem (1850), and in practical problems, such as the scheduling of round-robin tournaments (solution published 1880s). In the 20th century designs were applied to the design of experiments, notably Latin squares, finite geometry, and association schemes, yielding the field of algebraic statistics.

Fundamental combinatorial designs

The classical core of the subject of combinatorial designs is built around balanced incomplete block designs (BIBDs), Hadamard matrices and Hadamard designs, symmetric BIBDs, Latin squares, resolvable BIBDs, difference sets, and pairwise balanced designs (PBDs). [3] Other combinatorial designs are related to or have been developed from the study of these fundamental ones.

Two Latin squares of order n are said to be orthogonal if the set of all ordered pairs consisting of the corresponding entries in the two squares has n2 distinct members (all possible ordered pairs occur). A set of Latin squares of the same order forms a set of mutually orthogonal Latin squares (MOLS) if every pair of Latin squares in the set are orthogonal. There can be at most n  1 squares in a set of MOLS of order n. A set of n  1 MOLS of order n can be used to construct a projective plane of order n (and conversely).
If D is a difference set, and g in G, then gD = {gd: d in D} is also a difference set, and is called a translate of D. The set of all translates of a difference set D forms a symmetric block design. In such a design there are v elements and v blocks. Each block of the design consists of k points, each point is contained in k blocks. Any two blocks have exactly λ elements in common and any two points appear together in λ blocks. This SBIBD is called the development of D. [7]
In particular, if λ = 1, then the difference set gives rise to a projective plane. An example of a (7,3,1) difference set in the group (an abelian group written additively) is the subset {1,2,4}. The development of this difference set gives the Fano plane.
Since every difference set gives an SBIBD, the parameter set must satisfy the Bruck–Ryser–Chowla theorem, but not every SBIBD gives a difference set.
Given an Hadamard matrix of order 4a in standardized form, remove the first row and first column and convert every −1 to a 0. The resulting 0–1 matrix M is the incidence matrix of a symmetric 2  (4a  1, 2a  1, a  1) design called an Hadamard 2-design. [8] This construction is reversible, and the incidence matrix of a symmetric 2-design with these parameters can be used to form an Hadamard matrix of order 4a. When a = 2 we obtain the, by now familiar, Fano plane as an Hadamard 2-design.
Fisher's inequality holds for PBDs: [9] For any non-trivial PBD, v  b.
This result also generalizes the famous Erdős–De Bruijn theorem: For a PBD with λ = 1 having no blocks of size 1 or size v, v  b, with equality if and only if the PBD is a projective plane or a near-pencil. [10]

Other combinatorial designs

The Handbook of Combinatorial Designs( Colbourn & Dinitz 2007 ) has, amongst others, 65 chapters, each devoted to a combinatorial design other than those given above. A partial listing is given below:

  1. Each element appears R = ρ1 + 2ρ2 times altogether, with multiplicity one in exactly ρ1 blocks and multiplicity two in exactly ρ2 blocks.
  2. Every pair of distinct elements appears Λ times (counted with multiplicity); that is, if mvb is the multiplicity of the element v in block b, then for every pair of distinct elements v and w, .
For example, one of the only two nonisomorphic BTD(4,8;2,3,8;4,6)s (blocks are columns) is: [11]
11122311
11122322
23434433
23434444
The incidence matrix of a BTD (where the entries are the multiplicities of the elements in the blocks) can be used to form a ternary error-correcting code analogous to the way binary codes are formed from the incidence matrices of BIBDs. [12]
  1. every element of V appears precisely once in each column, and
  2. every element of V appears at most twice in each row.
An example of a BTD(3) is given by
1 63 52 34 52 4
2 54 61 41 33 6
3 41 25 62 61 5
The columns of a BTD(n) provide a 1-factorization of the complete graph on 2n vertices, K2n. [13]
BTD(n)s can be used to schedule round-robin tournaments: the rows represent the locations, the columns the rounds of play and the entries are the competing players or teams.
A frequency square F1 of order n based on the set {s1,s2, ..., sm} with frequency vector (λ1, λ2, ...,λm) and a frequency square F2, also of order n, based on the set {t1,t2, ..., tk} with frequency vector (μ1, μ2, ...,μk) are orthogonal if every ordered pair (si, tj) appears precisely λiμj times when F1 and F2 are superimposed.
Any affine space AG(n,3) gives an example of an HTS. Such an HTS is an affine HTS. Nonaffine HTSs also exist.
The number of points of an HTS is 3m for some integer m  2. Nonaffine HTSs exist for any m  4 and do not exist for m = 2 or 3. [14]
Every Steiner triple system is equivalent to a Steiner quasigroup (idempotent, commutative and satisfying (xy)y = x for all x and y). A Hall triple system is equivalent to a Steiner quasigroup which is distributive, that is, satisfies a(xy) = (ax)(ay) for all a,x,y in the quasigroup. [15]
  1. Each cell of the array is either empty or contains an unordered pair from S,
  2. Each symbol occurs exactly once in each row and column of the array, and
  3. Every unordered pair of symbols occurs in at most one cell of the array.
An example of an H(4,6) is
0 4 1 32 5
2 31 40 5 
 3 52 40 1
1 50 2 3 4
An H(2n  1, 2n) is a Room square of side 2n  1, and thus the Howell designs generalize the concept of Room squares.
The pairs of symbols in the cells of a Howell design can be thought of as the edges of an s regular graph on 2n vertices, called the underlying graph of the Howell design.
Cyclic Howell designs are used as Howell movements in duplicate bridge tournaments. The rows of the design represent the rounds, the columns represent the boards, and the diagonals represent the tables. [16]
{1,2,3,4,7}     {1,2,5,6,7}     {3,4,5,6,7}.
Lotto designs model any lottery that is run in the following way: Individuals purchase tickets consisting of k numbers chosen from a set of n numbers. At a certain point the sale of tickets is stopped and a set of p numbers is randomly selected from the n numbers. These are the winning numbers. If any sold ticket contains t or more of the winning numbers, a prize is given to the ticket holder. Larger prizes go to tickets with more matches. The value of L(n,k,p,t) is of interest to both gamblers and researchers, as this is the smallest number of tickets that are needed to be purchased in order to guarantee a prize.
The Hungarian Lottery is a (90,5,5,t)-lotto design and it is known that L(90,5,5,2) = 100. Lotteries with parameters (49,6,6,t) are also popular worldwide and it is known that L(49,6,6,2) = 19. In general though, these numbers are hard to calculate and remain unknown. [18]
A geometric construction of one such design is given in Transylvanian lottery.
(0,1,2)    (1,0,3)    (2,1,3)    (0,2,3)
Any triple system can be made into a Mendelson triple system by replacing the unordered triple {a,b,c} with the pair of ordered triples (a,b,c) and (a,c,b), but as the example shows, the converse of this statement is not true.
If (Q,∗) is an idempotent semisymmetric quasigroup, that is, xx = x (idempotent) and x ∗ (yx) = y (semisymmetric) for all x, y in Q, let β = {(x,y,xy): x, y in Q}. Then (Q, β) is a Mendelsohn triple system MTS(|Q|,1). This construction is reversible. [19]
Every quasisymmetric block design gives rise to a strongly regular graph (as its block graph), but not all SRGs arise in this way. [22]
The incidence matrix of a quasisymmetric 2-(v,k,λ) design with kxy (mod 2) generates a binary self-orthogonal code (when bordered if k is odd). [23]
of total degree at most t is equal to the average value of f on the whole sphere, i.e., the integral of f divided by the area of the sphere.
  1. each row is a permutation of the n symbols and
  2. for any two distinct symbols a and b and for each m from 1 to k, there is at most one row in which b is m steps to the right of a.
If r = n and k = 1 these are referred to as Tuscan squares, while if r = n and k = n - 1 they are Florentine squares. A Roman square is a Tuscan square which is also a latin square (these are also known as row complete Latin squares). A Vatican square is a Florentine square which is also a Latin square.
The following example is a tuscan-1 square on 7 symbols which is not tuscan-2: [24]
6152437
2635471
5723146
4251673
3621745
1327564
7653412
A tuscan square on n symbols is equivalent to a decomposition of the complete graph with n vertices into n hamiltonian directed paths. [25]
In a sequence of visual impressions, one flash card may have some effect on the impression given by the next. This bias can be cancelled by using n sequences corresponding to the rows of an n × n tuscan-1 square. [26]
Notice that in the following example of a 3-{12,{4,6},1) design based on the set X = {1,2,...,12}, some pairs appear four times (such as 1,2) while others appear five times (6,12 for instance). [28]
1 2 3 4 5 6           1 2 7 8      1 2 9 11      1 2 10 12      3 5 7 8      3 5 9 11      3 5 10 12      4 6 7 8      4 6 9 11      4 6 10 12
7 8 9 10 11 12      2 3 8 9      2 3 10 7      2 3 11 12      4 1 8 9      4 1 10 7      4 1 11 12      5 6 8 9      5 6 10 7      5 6 11 12
                         3 4 9 10      3 4 11 8      3 4 7 12      5 2 9 10      5 2 11 8      5 2 7 12      1 6 9 10      1 6 11 8      1 6 7 12
                         4 5 10 11      4 5 7 9      4 5 8 12      1 3 10 11      1 3 7 9      1 3 8 12      2 6 10 11      2 6 7 9      2 6 8 12
                         5 1 11 7      5 1 8 10      5 1 9 12      2 4 11 7      2 4 8 10      2 4 9 12      3 6 11 7      3 6 8 10      3 6 9 12
1234567
2345671
3456712
5671234
The seven blocks (columns) form the order 2 biplane (a symmetric (7,4,2)-design).

See also

Notes

  1. Stinson 2003 , pg.1
  2. Hayashi, Takao (2008). "Magic Squares in Indian Mathematics". Encyclopaedia of the History of Science, Technology, and Medicine in Non-Western Cultures (2 ed.). Springer. pp. 1252–1259. doi:10.1007/978-1-4020-4425-0_9778. ISBN   978-1-4020-4559-2.
  3. Stinson 2003 , pg. IX
  4. Beth, Jungnickel & Lenz 1986 , pg. 40 Example 5.8
  5. Ryser 1963 , pg. 52, Theorem 3.1
  6. When the group G is an abelian group (or written additively) the defining property looks like d1 –d2 from which the term difference set comes from.
  7. Beth, Jungnickel & Lenz 1986 , pg. 262, Theorem 1.6
  8. Stinson 2003 , pg. 74, Theorem 4.5
  9. Stinson 2003 , pg. 193, Theorem 8.20
  10. Stinson 2003 , pg. 183, Theorem 8.5
  11. Colbourn & Dinitz 2007 , pg. 331, Example 2.2
  12. Colbourn & Dinitz 2007 , pg. 331, Remark 2.8
  13. Colbourn & Dinitz 2007 , pg. 333, Remark 3.3
  14. Colbourn & Dinitz 2007 , pg. 496, Theorem 28.5
  15. Colbourn & Dinitz 2007 , pg. 497, Theorem 28.15
  16. Colbourn & Dinitz 2007 , pg. 503, Remark 29.38
  17. Colbourn & Dinitz 2007 , pg. 512, Example 32.4
  18. Colbourn & Dinitz 2007 , pg. 512, Remark 32.3
  19. Colbourn & Dinitz 2007 , pg. 530, Theorem 35.15
  20. Colbourn & Dinitz 2007 , pg. 577, Theorem 47.15
  21. Colbourn & Dinitz 2007 , pp. 578-579
  22. Colbourn & Dinitz 2007 , pg. 579, Theorem 48.10
  23. Colbourn & Dinitz 2007 , pg. 580, Lemma 48.22
  24. Colbourn & Dinitz 2007 , pg. 652, Examples 62.4
  25. Colbourn & Dinitz 2007 , pg. 655, Theorem 62.24
  26. Colbourn & Dinitz 2007 , pg. 657, Remark 62.29
  27. Colbourn & Dinitz 2007 , pg. 657
  28. Colbourn & Dinitz 2007 , pg. 658, Example 63.5
  29. Raghavarao & Padgett 1988 , pg. 305-308
  30. Colbourn & Dinitz 2007 , pg. 669, Remark 65.3

Related Research Articles

<span class="mw-page-title-main">Steiner system</span> Block design in combinatorial mathematics

In combinatorial mathematics, a Steiner system is a type of block design, specifically a t-design with λ = 1 and t = 2 or (recently) t ≥ 2.

<span class="mw-page-title-main">Symmetric matrix</span> Matrix equal to its transpose

In linear algebra, a symmetric matrix is a square matrix that is equal to its transpose. Formally,

In mathematics, an incidence matrix is a logical matrix that shows the relationship between two classes of objects, usually called an incidence relation. If the first class is X and the second is Y, the matrix has one row for each element of X and one column for each element of Y. The entry in row x and column y is 1 if x and y are related and 0 if they are not. There are variations; see below.

The Bruck–Ryser–Chowla theorem is a result on the combinatorics of block designs that implies nonexistence of certain kinds of design. It states that if a (v, b, r, k, λ)-design exists with v = b (a symmetric block design), then:

In combinatorial mathematics, two Latin squares of the same size (order) are said to be orthogonal if when superimposed the ordered paired entries in the positions are all distinct. A set of Latin squares, all of the same order, all pairs of which are orthogonal is called a set of mutually orthogonal Latin squares. This concept of orthogonality in combinatorics is strongly related to the concept of blocking in statistics, which ensures that independent variables are truly independent with no hidden confounding correlations. "Orthogonal" is thus synonymous with "independent" in that knowing one variable's value gives no further information about another variable's likely value.

<span class="mw-page-title-main">Hadamard matrix</span> Mathematics concept

In mathematics, a Hadamard matrix, named after the French mathematician Jacques Hadamard, is a square matrix whose entries are either +1 or −1 and whose rows are mutually orthogonal. In geometric terms, this means that each pair of rows in a Hadamard matrix represents two perpendicular vectors, while in combinatorial terms, it means that each pair of rows has matching entries in exactly half of their columns and mismatched entries in the remaining columns. It is a consequence of this definition that the corresponding properties hold for columns as well as rows.

<span class="mw-page-title-main">Incidence structure</span> Abstract mathematical system of two types of objects and a relation between them

In mathematics, an incidence structure is an abstract system consisting of two types of objects and a single relationship between these types of objects. Consider the points and lines of the Euclidean plane as the two types of objects and ignore all the properties of this geometry except for the relation of which points are on which lines for all points and lines. What is left is the incidence structure of the Euclidean plane.

In combinatorial mathematics, a block design is an incidence structure consisting of a set together with a family of subsets known as blocks, chosen such that frequency of the elements satisfies certain conditions making the collection of blocks exhibit symmetry (balance). Block designs have applications in many areas, including experimental design, finite geometry, physical chemistry, software testing, cryptography, and algebraic geometry.

In mathematics, incidence geometry is the study of incidence structures. A geometric structure such as the Euclidean plane is a complicated object that involves concepts such as length, angles, continuity, betweenness, and incidence. An incidence structure is what is obtained when all other concepts are removed and all that remains is the data about which points lie on which lines. Even with this severe limitation, theorems can be proved and interesting facts emerge concerning this structure. Such fundamental results remain valid when additional concepts are added to form a richer geometry. It sometimes happens that authors blur the distinction between a study and the objects of that study, so it is not surprising to find that some authors refer to incidence structures as incidence geometries.

The theory of association schemes arose in statistics, in the theory of experimental design for the analysis of variance. In mathematics, association schemes belong to both algebra and combinatorics. In algebraic combinatorics, association schemes provide a unified approach to many topics, for example combinatorial designs and the theory of error-correcting codes. In algebra, association schemes generalize groups, and the theory of association schemes generalizes the character theory of linear representations of groups.

In mathematics, a two-graph is a set of (unordered) triples chosen from a finite vertex setX, such that every (unordered) quadruple from X contains an even number of triples of the two-graph. A regular two-graph has the property that every pair of vertices lies in the same number of triples of the two-graph. Two-graphs have been studied because of their connection with equiangular lines and, for regular two-graphs, strongly regular graphs, and also finite groups because many regular two-graphs have interesting automorphism groups.

In combinatorics, a difference set is a subset of size of a group of order such that every non-identity element of can be expressed as a product of elements of in exactly ways. A difference set is said to be cyclic, abelian, non-abelian, etc., if the group has the corresponding property. A difference set with is sometimes called planar or simple. If is an abelian group written in additive notation, the defining condition is that every non-zero element of can be written as a difference of elements of in exactly ways. The term "difference set" arises in this way.

In mathematics, a conference matrix (also called a C-matrix) is a square matrix C with 0 on the diagonal and +1 and −1 off the diagonal, such that CTC is a multiple of the identity matrix I. Thus, if the matrix has order n, CTC = (n−1)I. Some authors use a more general definition, which requires there to be a single 0 in each row and column but not necessarily on the diagonal.

In mathematics a regular Hadamard matrix is a Hadamard matrix whose row and column sums are all equal. While the order of a Hadamard matrix must be 1, 2, or a multiple of 4, regular Hadamard matrices carry the further restriction that the order be a square number. The excess, denoted E(H), of a Hadamard matrix H of order n is defined to be the sum of the entries of H. The excess satisfies the bound |E(H)| ≤ n3/2. A Hadamard matrix attains this bound if and only if it is regular.

Fisher's inequality is a necessary condition for the existence of a balanced incomplete block design, that is, a system of subsets that satisfy certain prescribed conditions in combinatorial mathematics. Outlined by Ronald Fisher, a population geneticist and statistician, who was concerned with the design of experiments such as studying the differences among several different varieties of plants, under each of a number of different growing conditions, called blocks.

In mathematics, the Turán number T(n,k,r) for r-uniform hypergraphs of order n is the smallest number of r-edges such that every induced subgraph on k vertices contains an edge. This number was determined for r = 2 by Turán (1941), and the problem for general r was introduced in Turán (1961). The paper (Sidorenko 1995) gives a survey of Turán numbers.

In mathematics, an orthogonal array is a "table" (array) whose entries come from a fixed finite set of symbols, arranged in such a way that there is an integer t so that for every selection of t columns of the table, all ordered t-tuples of the symbols, formed by taking the entries in each row restricted to these columns, appear the same number of times. The number t is called the strength of the orthogonal array. Here are two examples:

In combinatorial mathematics, a Latin rectangle is an r × n matrix, using n symbols, usually the numbers 1, 2, 3, ..., n or 0, 1, ..., n − 1 as its entries, with no number occurring more than once in any row or column.

Hadamard's maximal determinant problem, named after Jacques Hadamard, asks for the largest determinant of a matrix with elements equal to 1 or −1. The analogous question for matrices with elements equal to 0 or 1 is equivalent since, as will be shown below, the maximal determinant of a {1,−1} matrix of size n is 2n−1 times the maximal determinant of a {0,1} matrix of size n−1. The problem was posed by Hadamard in the 1893 paper in which he presented his famous determinant bound and remains unsolved for matrices of general size. Hadamard's bound implies that {1, −1}-matrices of size n have determinant at most nn/2. Hadamard observed that a construction of Sylvester produces examples of matrices that attain the bound when n is a power of 2, and produced examples of his own of sizes 12 and 20. He also showed that the bound is only attainable when n is equal to 1, 2, or a multiple of 4. Additional examples were later constructed by Scarpis and Paley and subsequently by many other authors. Such matrices are now known as Hadamard matrices. They have received intensive study.

References