Block design

Last updated

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[ clarification needed ] 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.

Contents

Without further specifications the term block design usually refers to a balanced incomplete block design (BIBD), specifically (and also synonymously) a 2-design, which has been the most intensely studied type historically due to its application in the design of experiments. [1] [2] Its generalization is known as a t-design.

Overview

A design is said to be balanced (up to t) if all t-subsets of the original set occur in equally many (i.e., λ) blocks[ clarification needed ]. When t is unspecified, it can usually be assumed to be 2, which means that each pair of elements is found in the same number of blocks and the design is pairwise balanced. For t=1, each element occurs in the same number of blocks (the replication number, denoted r) and the design is said to be regular. Any design balanced up to t is also balanced in all lower values of t (though with different λ-values), so for example a pairwise balanced (t=2) design is also regular (t=1). When the balancing requirement fails, a design may still be partially balanced if the t-subsets can be divided into n classes, each with its own (different) λ-value. For t=2 these are known as PBIBD(n) designs, whose classes form an association scheme.

Designs are usually said (or assumed) to be incomplete, meaning that the collection of blocks is not all possible k-subsets, thus ruling out a trivial design.

A block design in which all the blocks have the same size (usually denoted k) is called uniform or proper. The designs discussed in this article are all uniform. Block designs that are not necessarily uniform have also been studied; for t=2 they are known in the literature under the general name pairwise balanced designs (PBDs).

Block designs may or may not have repeated blocks. Designs without repeated blocks are called simple, [3] in which case the "family" of blocks is a set rather than a multiset.

In statistics, the concept of a block design may be extended to non-binary block designs, in which blocks may contain multiple copies of an element (see blocking (statistics)). There, a design in which each element occurs the same total number of times is called equireplicate, which implies a regular design only when the design is also binary. The incidence matrix of a non-binary design lists the number of times each element is repeated in each block.

Regular uniform designs (configurations)

The simplest type of "balanced" design (t=1) is known as a tactical configuration or 1-design. The corresponding incidence structure in geometry is known simply as a configuration, see Configuration (geometry). Such a design is uniform and regular: each block contains k elements and each element is contained in r blocks. The number of set elements v and the number of blocks b are related by , which is the total number of element occurrences.

Every binary matrix with constant row and column sums is the incidence matrix of a regular uniform block design. Also, each configuration has a corresponding biregular bipartite graph known as its incidence or Levi graph.

Pairwise balanced uniform designs (2-designs or BIBDs)

Given a finite set X (of elements called points) and integers k, r, λ ≥ 1, we define a 2-design (or BIBD, standing for balanced incomplete block design) B to be a family of k-element subsets of X, called blocks, such that any x in X is contained in r blocks, and any pair of distinct points x and y in X is contained in λ blocks. Here, the condition that any x in X is contained in r blocks is redundant, as shown below.

Here v (the number of elements of X, called points), b (the number of blocks), k, r, and λ are the parameters of the design. (To avoid degenerate examples, it is also assumed that v > k, so that no block contains all the elements of the set. This is the meaning of "incomplete" in the name of these designs.) In a table:

vpoints, number of elements of X
bnumber of blocks
rnumber of blocks containing a given point
knumber of points in a block
λnumber of blocks containing any 2 (or more generally t) distinct points

The design is called a (v, k, λ)-design or a (v, b, r, k, λ)-design. The parameters are not all independent; v, k, and λ determine b and r, and not all combinations of v, k, and λ are possible. The two basic equations connecting these parameters are

obtained by counting the number of pairs (B, p) where B is a block and p is a point in that block, and

obtained from counting for a fixed x the triples (x, y, B) where x and y are distinct points and B is a block that contains them both. This equation for every x also proves that r is constant (independent of x) even without assuming it explicitly, thus proving that the condition that any x in X is contained in r blocks is redundant and r can be computed from the other parameters.

These conditions are not sufficient as, for example, a (43,7,1)-design does not exist. [4]

The order of a 2-design is defined to be n = r  λ. The complement of a 2-design is obtained by replacing each block with its complement in the point set X. It is also a 2-design and has parameters v′ = v, b′ = b, r′ = b  r, k′ = v  k, λ′ = λ + b  2r. A 2-design and its complement have the same order.

A fundamental theorem, Fisher's inequality, named after the statistician Ronald Fisher, is that b  v in any 2-design.

A rather surprising and not very obvious but very general combinatorial result for these designs is that if points are denoted by any arbitrarily chosen set of equally or unequally spaced numerics, there is no choice of such a set which can make all block-sums (that is, sum of all points in a given block) constant. [5] [6] For other designs such as partially balanced incomplete block designs this may however be possible. Many such cases are discussed in. [7] However, it can also be observed trivially for the magic squares or magic rectangles which can be viewed as the partially balanced incomplete block designs.

Examples

The unique (6,3,2)-design (v = 6, k = 3, λ = 2) has 10 blocks (b = 10) and each element is repeated 5 times (r = 5). [8] Using the symbols 0 − 5, the blocks are the following triples:

012    013    024    035    045    125    134    145    234    235.

and the corresponding incidence matrix (a v×b binary matrix with constant row sum r and constant column sum k) is:

One of four nonisomorphic (8,4,3)-designs has 14 blocks with each element repeated 7 times. Using the symbols 0 − 7 the blocks are the following 4-tuples: [8]

0123    0124    0156    0257    0345    0367    0467    1267    1346    1357    1457    2347    2356    2456.

The unique (7,3,1)-design is symmetric and has 7 blocks with each element repeated 3 times. Using the symbols 0 − 6, the blocks are the following triples: [8]

013    026    045    124    156    235    346.

This design is associated with the Fano plane, with the elements and blocks of the design corresponding to the points and lines of the plane. Its corresponding incidence matrix can also be symmetric, if the labels or blocks are sorted the right way:

Symmetric 2-designs (SBIBDs)

The case of equality in Fisher's inequality, that is, a 2-design with an equal number of points and blocks, is called a symmetric design. [9] Symmetric designs have the smallest number of blocks among all the 2-designs with the same number of points.

In a symmetric design r = k holds as well as b = v, and, while it is generally not true in arbitrary 2-designs, in a symmetric design every two distinct blocks meet in λ points. [10] A theorem of Ryser provides the converse. If X is a v-element set, and B is a v-element set of k-element subsets (the "blocks"), such that any two distinct blocks have exactly λ points in common, then (X, B) is a symmetric block design. [11]

The parameters of a symmetric design satisfy

This imposes strong restrictions on v, so the number of points is far from arbitrary. The Bruck–Ryser–Chowla theorem gives necessary, but not sufficient, conditions for the existence of a symmetric design in terms of these parameters.

The following are important examples of symmetric 2-designs:

Projective planes

Finite projective planes are symmetric 2-designs with λ = 1 and order n > 1. For these designs the symmetric design equation becomes:

Since k = r we can write the order of a projective plane as n = k  1 and, from the displayed equation above, we obtain v = (n + 1)n + 1 = n2 + n + 1 points in a projective plane of order n.

As a projective plane is a symmetric design, we have b = v, meaning that b = n2 + n + 1 also. The number b is the number of lines of the projective plane. There can be no repeated lines since λ = 1, so a projective plane is a simple 2-design in which the number of lines and the number of points are always the same. For a projective plane, k is the number of points on each line and it is equal to n + 1. Similarly, r = n + 1 is the number of lines with which a given point is incident.

For n = 2 we get a projective plane of order 2, also called the Fano plane, with v = 4 + 2 + 1 = 7 points and 7 lines. In the Fano plane, each line has n + 1 = 3 points and each point belongs to n + 1 = 3 lines.

Projective planes are known to exist for all orders which are prime numbers or powers of primes. They form the only known infinite family (with respect to having a constant λ value) of symmetric block designs. [12]

Biplanes

A biplane or biplane geometry is a symmetric 2-design with λ = 2; that is, every set of two points is contained in two blocks ("lines"), while any two lines intersect in two points. [12] They are similar to finite projective planes, except that rather than two points determining one line (and two lines determining one point), two points determine two lines (respectively, points). A biplane of order n is one whose blocks have k = n + 2 points; it has v = 1 + (n + 2)(n + 1)/2 points (since r = k).

The 18 known examples [13] are listed below.

Algebraically this corresponds to the exceptional embedding of the projective special linear group PSL(2,5) in PSL(2,11) – see projective linear group: action on p points for details. [15]

Biplanes of orders 5, 6, 8 and 10 do not exist, as shown by the Bruck-Ryser-Chowla theorem.

Hadamard 2-designs

An Hadamard matrix of size m is an m × m matrix H whose entries are ±1 such that HH = mIm, where H is the transpose of H and Im is the m × m identity matrix. An Hadamard matrix can be put into standardized form (that is, converted to an equivalent Hadamard matrix) where the first row and first column entries are all +1. If the size m > 2 then m must be a multiple of 4.

Given an Hadamard matrix of size 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. [19] It contains blocks/points; each contains/is contained in points/blocks. Each pair of points is contained in exactly blocks.

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 size 4a.

Resolvable 2-designs

A resolvable 2-design is a BIBD whose blocks can be partitioned into sets (called parallel classes), each of which forms a partition of the point set of the BIBD. The set of parallel classes is called a resolution of the design.

If a 2-(v,k,λ) resolvable design has c parallel classes, then b  v + c  1. [20]

Consequently, a symmetric design can not have a non-trivial (more than one parallel class) resolution. [21]

Archetypical resolvable 2-designs are the finite affine planes. A solution of the famous 15 schoolgirl problem is a resolution of a 2-(15,3,1) design. [22]

General balanced designs (t-designs)

Given any positive integer t, a t-design B is a class of k-element subsets of X, called blocks, such that every point x in X appears in exactly r blocks, and every t-element subset T appears in exactly λ blocks. The numbers v (the number of elements of X), b (the number of blocks), k, r, λ, and t are the parameters of the design. The design may be called a t-(v,k,λ)-design. Again, these four numbers determine b and r and the four numbers themselves cannot be chosen arbitrarily. The equations are

where λi is the number of blocks that contain any i-element set of points and λt = λ.

Note that and .

Theorem: [23] Any t-(v,k,λ)-design is also an s-(v,ks)-design for any s with 1  s  t. (Note that the "lambda value" changes as above and depends on s.)

A consequence of this theorem is that every t-design with t ≥ 2 is also a 2-design.

A t-(v,k,1)-design is called a Steiner system.

The term block design by itself usually means a 2-design.

Derived and extendable t-designs

Let D = (X, B) be a t-(v,k,λ) design and p a point of X. The derived designDp has point set X  {p} and as block set all the blocks of D which contain p with p removed. It is a (t  1)-(v  1, k  1, λ) design. Note that derived designs with respect to different points may not be isomorphic. A design E is called an extension of D if E has a point p such that Ep is isomorphic to D; we call Dextendable if it has an extension.

Theorem: [24] If a t-(v,k,λ) design has an extension, then k + 1 divides b(v + 1).

The only extendable projective planes (symmetric 2-(n2 + n + 1, n + 1, 1) designs) are those of orders 2 and 4. [25]

Every Hadamard 2-design is extendable (to an Hadamard 3-design). [26]

Theorem:. [27] If D, a symmetric 2-(v,k,λ) design, is extendable, then one of the following holds:

  1. D is an Hadamard 2-design,
  2. v =  + 2)(λ2 +  + 2), k = λ2 +  + 1,
  3. v = 495, k = 39, λ = 3.

Note that the projective plane of order two is an Hadamard 2-design; the projective plane of order four has parameters which fall in case 2; the only other known symmetric 2-designs with parameters in case 2 are the order 9 biplanes, but none of them are extendable; and there is no known symmetric 2-design with the parameters of case 3. [28]

Inversive planes

A design with the parameters of the extension of an affine plane, i.e., a 3-(n2 + 1, n + 1, 1) design, is called a finite inversive plane, or Möbius plane, of order n.

It is possible to give a geometric description of some inversive planes, indeed, of all known inversive planes. An ovoid in PG(3,q) is a set of q2 + 1 points, no three collinear. It can be shown that every plane (which is a hyperplane since the geometric dimension is 3) of PG(3,q) meets an ovoid O in either 1 or q + 1 points. The plane sections of size q + 1 of O are the blocks of an inversive plane of order q. Any inversive plane arising this way is called egglike. All known inversive planes are egglike.

An example of an ovoid is the elliptic quadric, the set of zeros of the quadratic form

x1x2 + f(x3, x4),

where f is an irreducible quadratic form in two variables over GF(q). [f(x,y) = x2 + xy + y2 for example].

If q is an odd power of 2, another type of ovoid is known – the Suzuki–Tits ovoid.

Theorem. Let q be a positive integer, at least 2. (a) If q is odd, then any ovoid is projectively equivalent to the elliptic quadric in a projective geometry PG(3,q); so q is a prime power and there is a unique egglike inversive plane of order q. (But it is unknown if non-egglike ones exist.) (b) if q is even, then q is a power of 2 and any inversive plane of order q is egglike (but there may be some unknown ovoids).

Partially balanced designs (PBIBDs)

An n-class association scheme consists of a set X of size v together with a partition S of X×X into n + 1 binary relations, R0, R1, ..., Rn. A pair of elements in relation Ri are said to be ithassociates. Each element of X has ni ith associates. Furthermore:

An association scheme is commutative if for all i, j and k. Most authors assume this property.

A partially balanced incomplete block design with n associate classes (PBIBD(n)) is a block design based on a v-set X with b blocks each of size k and with each element appearing in r blocks, such that there is an association scheme with n classes defined on X where, if elements x and y are ith associates, 1 ≤ in, then they are together in precisely λi blocks.

A PBIBD(n) determines an association scheme but the converse is false. [29]

Example

Let A(3) be the following association scheme with three associate classes on the set X = {1,2,3,4,5,6}. The (i,j) entry is s if elements i and j are in relation Rs.

 123456
1 0  1  1  2  3  3 
2 1  0  1  3  2  3 
3 1  1  0  3  3  2 
4 2  3  3  0  1  1 
5 3  2  3  1  0  1 
6 3  3  2  1  1  0 

The blocks of a PBIBD(3) based on A(3) are:

 124  134  235  456 
 125  136  236  456 

The parameters of this PBIBD(3) are: v =  6, b =  8, k =  3, r =  4 and λ1 = λ2 = 2 and λ3 = 1. Also, for the association scheme we have n0 = n2 =  1 and n1 = n3 =  2. [30] The incidence matrix M is

and the concurrence matrix MMT is

from which we can recover the λ and r values.

Properties

The parameters of a PBIBD(m) satisfy: [31]

A PBIBD(1) is a BIBD and a PBIBD(2) in which λ1 =  λ2 is a BIBD. [32]

Two associate class PBIBDs

PBIBD(2)s have been studied the most since they are the simplest and most useful of the PBIBDs. [33] They fall into six types [34] based on a classification of the then known PBIBD(2)s by Bose & Shimamoto (1952): [35]

  1. group divisible;
  2. triangular;
  3. Latin square type;
  4. cyclic;
  5. partial geometry type;
  6. miscellaneous.

Applications

The mathematical subject of block designs originated in the statistical framework of design of experiments. These designs were especially useful in applications of the technique of analysis of variance (ANOVA). This remains a significant area for the use of block designs.

While the origins of the subject are grounded in biological applications (as is some of the existing terminology), the designs are used in many applications where systematic comparisons are being made, such as in software testing.

The incidence matrix of block designs provide a natural source of interesting block codes that are used as error correcting codes. The rows of their incidence matrices are also used as the symbols in a form of pulse-position modulation. [36]

Statistical application

Suppose that skin cancer researchers want to test three different sunscreens. They coat two different sunscreens on the upper sides of the hands of a test person. After a UV radiation they record the skin irritation in terms of sunburn. The number of treatments is 3 (sunscreens) and the block size is 2 (hands per person).

A corresponding BIBD can be generated by the R-function design.bib of the R-package agricolae and is specified in the following table:

PlotsBlockTreatment
10113
10212
20121
20223
30132
30231

The investigator chooses the parameters v = 3, k = 2 and λ = 1 for the block design which are then inserted into the R-function. Subsequently, the remaining parameters b and r are determined automatically.

Using the basic relations we calculate that we need b = 3 blocks, that is, 3 test people in order to obtain a balanced incomplete block design. Labeling the blocks A, B and C, to avoid confusion, we have the block design,

A = {2, 3},   B = {1, 3} and C = {1, 2}.

A corresponding incidence matrix is specified in the following table:

TreatmentBlock ABlock BBlock C
1011
2101
3110

Each treatment occurs in 2 blocks, so r = 2.

Just one block (C) contains the treatments 1 and 2 simultaneously and the same applies to the pairs of treatments (1,3) and (2,3). Therefore, λ = 1.

It is impossible to use a complete design (all treatments in each block) in this example because there are 3 sunscreens to test, but only 2 hands on each person.

See also

Notes

  1. Colbourn & Dinitz 2007 , pp.17−19
  2. Stinson 2003 , p.1
  3. P. Dobcsányi, D.A. Preece. L.H. Soicher (2007-10-01). "On balanced incomplete-block designs with repeated blocks". European Journal of Combinatorics . 28 (7): 1955–1970. doi: 10.1016/j.ejc.2006.08.007 . ISSN   0195-6698.
  4. Proved by Tarry in 1900 who showed that there was no pair of orthogonal Latin squares of order six. The 2-design with the indicated parameters is equivalent to the existence of five mutually orthogonal Latin squares of order six.
  5. Khattree 2019
  6. Khattree 2022
  7. Khattree 2022
  8. 1 2 3 Colbourn & Dinitz 2007 , p. 27
  9. They have also been referred to as projective designs or square designs. These alternatives have been used in an attempt to replace the term "symmetric", since there is nothing symmetric (in the usual meaning of the term) about these designs. The use of projective is due to P.Dembowski (Finite Geometries, Springer, 1968), in analogy with the most common example, projective planes, while square is due to P. Cameron (Designs, Graphs, Codes and their Links, Cambridge, 1991) and captures the implication of v = b on the incidence matrix. Neither term has caught on as a replacement and these designs are still universally referred to as symmetric.
  10. Stinson 2003 , pg.23, Theorem 2.2
  11. Ryser 1963 , pp. 102–104
  12. 1 2 Hughes & Piper 1985 , pg.109
  13. Hall 1986 , pp.320-335
  14. Assmus & Key 1992 , pg.55
  15. Martin, Pablo; Singerman, David (April 17, 2008), From Biplanes to the Klein quartic and the Buckyball (PDF), p. 4
  16. Salwach & Mezzaroba 1978
  17. Kaski & Östergård 2008
  18. Aschbacher 1971 , pp. 279–281
  19. Stinson 2003 , pg. 74, Theorem 4.5
  20. Hughes & Piper 1985 , pg. 156, Theorem 5.4
  21. Hughes & Piper 1985 , pg. 158, Corollary 5.5
  22. Beth, Jungnickel & Lenz 1986 , pg. 40 Example 5.8
  23. Stinson 2003 , pg.203, Corollary 9.6
  24. Hughes & Piper 1985 , pg.29
  25. Cameron & van Lint 1991 , pg. 11, Proposition 1.34
  26. Hughes & Piper 1985 , pg. 132, Theorem 4.5
  27. Cameron & van Lint 1991 , pg. 11, Theorem 1.35
  28. Colbourn & Dinitz 2007 , pg. 114, Remarks 6.35
  29. Street & Street 1987 , pg. 237
  30. Street & Street 1987 , pg. 238
  31. Street & Street 1987 , pg. 240, Lemma 4
  32. Colbourn & Dinitz 2007 , pg. 562, Remark 42.3 (4)
  33. Street & Street 1987 , pg. 242
  34. Not a mathematical classification since one of the types is a catch-all "and everything else".
  35. Raghavarao 1988 , pg. 127
  36. Noshad, Mohammad; Brandt-Pearce, Maïté (Jul 2012). "Expurgated PPM Using Symmetric Balanced Incomplete Block Designs". IEEE Communications Letters. 16 (7): 968–971. arXiv: 1203.5378 . Bibcode:2012arXiv1203.5378N. doi:10.1109/LCOMM.2012.042512.120457. S2CID   7586742.

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">Ellipsoid</span> Quadric surface that looks like a deformed sphere

An ellipsoid is a surface that can be obtained from a sphere by deforming it by means of directional scalings, or more generally, of an affine transformation.

In mathematics, particularly in linear algebra, a skew-symmetricmatrix is a square matrix whose transpose equals its negative. That is, it satisfies the condition

In linear algebra, a diagonal matrix is a matrix in which the entries outside the main diagonal are all zero; the term usually refers to square matrices. Elements of the main diagonal can either be zero or nonzero. An example of a 2×2 diagonal matrix is , while an example of a 3×3 diagonal matrix is. An identity matrix of any size, or any multiple of it is a diagonal matrix called scalar matrix, for example, . In geometry, a diagonal matrix may be used as a scaling matrix, since matrix multiplication with it results in changing scale (size) and possibly also shape; only a scalar matrix results in uniform change in scale.

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 the mathematical field of graph theory, the Laplacian matrix, also called the graph Laplacian, admittance matrix, Kirchhoff matrix or discrete Laplacian, is a matrix representation of a graph. Named after Pierre-Simon Laplace, the graph Laplacian matrix can be viewed as a matrix form of the negative discrete Laplace operator on a graph approximating the negative continuous Laplacian obtained by the finite difference method.

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.

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.

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.

A complex Hadamard matrix is any complex matrix satisfying two conditions:

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 must 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, 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:

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.

In geometry, a unital is a set of n3 + 1 points arranged into subsets of size n + 1 so that every pair of distinct points of the set are contained in exactly one subset. This is equivalent to saying that a unital is a 2-(n3 + 1, n + 1, 1) block design. Some unitals may be embedded in a projective plane of order n2 (the subsets of the design become sets of collinear points in the projective plane). In this case of embedded unitals, every line of the plane intersects the unital in either 1 or n + 1 points. In the Desarguesian planes, PG(2,q2), the classical examples of unitals are given by nondegenerate Hermitian curves. There are also many non-classical examples. The first and the only known unital with non prime power parameters, n=6, was constructed by Bhaskar Bagchi and Sunanda Bagchi. It is still unknown if this unital can be embedded in a projective plane of order 36, if such a plane exists.

References