Small Latin squares and quasigroups

Last updated

Latin squares and quasigroups are equivalent mathematical objects, although the former has a combinatorial nature while the latter is more algebraic. The listing below will consider the examples of some very small orders, which is the side length of the square, or the number of elements in the equivalent quasigroup.

Contents

The equivalence

Given a quasigroup Q with n elements, its Cayley table (almost universally called its multiplication table) is an (n + 1) × (n + 1) table that includes borders; a top row of column headers and a left column of row headers. Removing the borders leaves an n × n array that is a Latin square. This process can be reversed, starting with a Latin square, introduce a bordering row and column to obtain the multiplication table of a quasigroup. While there is complete arbitrariness in how this bordering is done, the quasigroups obtained by different choices are sometimes equivalent in the sense given below.

Isotopy and isomorphism

Two Latin squares, L1 and L2 of size n are isotopic if there are three bijections from the rows, columns and symbols of L1 onto the rows, columns and symbols of L2, respectively, that map L1 to L2. [1] Isotopy is an equivalence relation and the equivalence classes are called isotopy classes.

A stronger form of equivalence exists. Two Latin squares, L1 and L2 of side n with common symbol set S that is also the index set for the rows and columns of each square, are isomorphic if there is a bijection g:S → S such that g(L1(i, j)) = L2(g(i), g(j)) for all i, j in S. [1] An alternate way to define isomorphic Latin squares is to say that a pair of isotopic Latin squares are isomorphic if the three bijections used to show that they are isotopic are, in fact, equal. [2] Isomorphism is also an equivalence relation and its equivalence classes are called isomorphism classes.

An alternate representation of a Latin square is given by an orthogonal array. For a Latin square of order n this is an n2 × 3 matrix with columns labeled r, c and s and whose rows correspond to a single position of the Latin square, namely, the row of the position, the column of the position and the symbol in the position. Thus for the order three Latin square,

123
231
312

the orthogonal array is given by:

rcs
111
122
133
212
223
231
313
321
332

The condition for an appropriately sized matrix to represent a Latin square is that for any two columns the n2 ordered pairs determined by the rows in those columns are all the pairs (i, j) with 1 ≤ i, jn, once each.

This property is not lost by permuting the three columns (but not the labels), so another orthogonal array (and thus, another Latin square) is obtained. For example, by permuting the first two columns, which corresponds to transposing the square (reflecting about its main diagonal) gives another Latin square, which may or may not be isotopic to the original. In this case, if the quasigroup corresponding to this Latin square satisfies the commutative law, the new Latin square is the same as the original one. Altogether there are six possibilities including "do nothing", giving at most six Latin squares called the conjugates (also parastrophes) of the original square. [3]

Two Latin squares are said to be paratopic, also main class isotopic, if one of them is isotopic to a conjugate of the other. This is also an equivalence relation, with the equivalence classes called main classes, species, or paratopy classes. [3] Each main class contains up to six isotopy classes.

A main class is a disjoint union of isotopy classes and an isotopy class is a disjoint union of isomorphism classes. [4]

Isotopic quasigroups

Let (Q,∘) and (R,∗) be two quasigroups. An ordered triple (f,g,h) of bijections from Q onto R is called an isotopism of (Q,∘) onto (R,∗) if f(x) ∗ g(y) = h(xy) for all x, y in G. Such quasigroups are said to be isotopic. [5]

If in the above definition f = g = h then the quasigroups are said to be isomorphic. [2]

Unlike the situation with Latin squares, when two isotopic quasigroups are represented by Cayley tables (bordered Latin squares), the permutations f and g operate only on the border headings and do not move columns and rows, while h operates on the body of the table. [5]

Permuting the rows and columns of a Cayley table (including the headings) does not change the quasigroup it defines, however, the Latin square associated with this table will be permuted to an isotopic Latin square. Thus, normalizing a Cayley table (putting the border headings in some fixed predetermined order by permuting rows and columns including the headings) preserves the isotopy class of the associated Latin square. Furthermore, if two normalized Cayley tables represent isomorphic quasigroups then their associated Latin squares are also isomorphic. Hence, the number of distinct quasigroups of a given order is the number of isomorphism classes of Latin squares of that order. [6]

Notation

The set of symbols used in a Latin square (or quasigroup) is arbitrary and individual symbols carry no meaning, even if they may have a meaning in other contexts. Thus, since it is most common to see the symbol sets {1,2, ... , n} or {0,1, ... , n − 1} used, one must remember that these symbols carry no numerical meaning. To stress this point, small Latin squares sometimes use letters of the alphabet as a symbol set.

Counting Latin squares

As a Latin square is a combinatorial object, the symbol set used to write the square is immaterial. Thus, as Latin squares, these should be considered the same:

  and  

Similarly, and for the same reason,

  and  

should be thought of as the same. Thus,

  and  

can not be thought of as different Latin squares.

This intuitive argument can be made more precise. The Latin squares

  and  

are isotopic, in several ways. Let (a,b) be the involutorial permutation on the set S = {a,b}, sending a to b and b to a. Then the isotopy {(a,b), id, id} will interchange the two rows of the first square to give the second square (id is the identity permutation). But, {id, (a,b), id} which interchanges the two columns is also an isotopy, as is {id, id, (a,b)} which interchanges the two symbols. However, {(a,b), (a,b), (a,b)} is also an isotopy between the two squares, and so, this pair of squares are isomorphic.

Reduced squares

Finding a given Latin square's isomorphism class can be a difficult computational problem for squares of large order. To reduce the problem somewhat, a Latin square can always be put into a standard form known as a reduced square. A reduced square has its top row elements written in some natural order for the symbol set (for example, integers in increasing order or letters in alphabetical order). The left column entries are put in the same order. As this can be done by row and column permutations, every Latin square is isotopic to a reduced square. Thus, every isotopy class must contain a reduced Latin square, however, a Latin square may have more than one reduced square that is isotopic to it. In fact, there may be more than one reduced square in a given isomorphism class. [7]

For example, the reduced Latin squares of order four,

  and  

are both in the isomorphism class that also contains the reduced square

This can be shown by the isomorphisms {(3,4), (3,4), (3,4)} and {(2,3), (2,3), (2,3)} respectively. [8]

Since isotopy classes are disjoint, the number of reduced Latin squares gives an upper bound on the number of isotopy classes. Also, the total number of Latin squares is n!(n − 1)! times the number of reduced squares. [9]

One can normalize a Cayley table of a quasigroup in the same manner as a reduced Latin square. Then the quasigroup associated to a reduced Latin square has a (two sided) identity element (namely, the first element among the row headers). A quasigroup with a two sided identity is called a loop . Some, but not all, loops are groups. To be a group, the associative law must also hold.

Isotopy invariants

The counts of various substructures in a Latin square can be useful in distinguishing them from one another. Some of these counts are the same for every isotope of a Latin square and are referred to as isotopy invariants. One such invariant is the number of 2 × 2 subsquares, called intercalates. Another is the total number of transversals, a set of n positions in a Latin square of order n, one in each row and one in each column, that contain no element twice. Latin squares with different values for these counts must lie in different isotopy classes. The number of intercalates is also a main class invariant.

Order 1

For order one there is only one Latin square with symbol 1 and one quasigroup with underlying set {1}; it is a group, the trivial group.

Order 2

There is only one reduced Latin square of order two (and only two in total), namely

Since there is only one reduced square of this order, there is only one isotopy class. In fact, this isotopy class is also an isomorphism class (shown above). [8] [1]

As there is only one isomorphism class of Latin squares, there is only one quasigroup of order two (up to isomorphism) and it is the group usually denoted by the cyclic group of order two.

Order 3

There is also only one reduced Latin square of order three (out of 12),

Thus, there is only one isotopy class. [8] However, this isotopy class is the union of five isomorphism classes. [1]

Three of the five isomorphism classes contain three Latin squares each, one contains two and one contains just one. The reduced square is in an isomorphism class with three elements and so the corresponding quasigroup is a loop, in fact it is a group, the cyclic group of order three. A Latin square representative of each of these isomorphism classes is given by (the number below each is the number of squares in the corresponding isomorphism class):

Order 4

There are four reduced Latin squares of order four (out of 576 squares). These are:

The last three of these are isomorphic (see above). There are two main classes, two isotopy classes and 35 isomorphism classes. Among the 35 quasigroups, only two are loops, and they are in fact groups. Corresponding to the first square above is the Klein four group, while corresponding to any of other three squares is the cyclic group The first square has eight transversals and 12 intercalates, while each of the others have no transversals and four intercalates.

The isomorphism class of the Klein four group has four members, while the isomorphism class of the cyclic group has 12 members. Of the 576 Latin squares, 288 are solutions of the 4×4 version of Sudoku, sometimes called Shi Doku .

Order 5

Of the 161,280 Latin squares of order five, there are 56 reduced squares. There are two main classes and only two isotopy classes, but 1,411 isomorphism classes. There are six isomorphism classes that contain reduced squares, that is, there are six loops, only one of which is a group, the cyclic group of order five. [1]

Below are two reduced Latin squares of order five, one from each isotopy class. [10]

The first square has 15 transversals, no intercalates and is the unbordered Cayley table of the cyclic group The second square has three transversals and four intercalates. It represents a loop that is not a group, since, for instance, 2 + (3 + 4) = 2 + 0 = 2, while (2 + 3) + 4 = 0 + 4 = 4, so the associative law does not hold.

Orders 6 to 10

The number of Latin squares, as the order increases, exhibits the phenomenon known as combinatorial explosion; for even small increases in size there correspond huge increases in varieties. The basic counts are given in the next two tables, [1] and not much beyond what is presented here is known with exactness.

Number of isotopy classes of Latin squares and Quasigroups (isomorphism classes)
nmain classesisotopy classesisomorphism classes
612221,130,531
714756412,198,455,835
8283,6571,676,2672,697,818,331,680,661
919,270,853,541115,618,721,53315,224,734,061,438,247,321,497
1034,817,397,894,749,939208,904,371,354,363,0062,750,892,211,809,150,446,995,735,533,513
The numbers of reduced Latin squares and loops of various sizes
nreduced Latin squares of size nloops of size n
69,408109
716,942,08023,746
8535,281,401,856106,228,849
9377,597,570,964,258,8169,365,022,303,540
107,580,721,483,160,132,811,489,28020,890,436,195,945,769,617

History

This account follows McKay, Meynert & Myrvold (2007 , p. 100).

The counting of Latin squares has a long history, but the published accounts contain many errors. Euler in 1782, [11] and Cayley in 1890, [12] both knew the number of reduced Latin squares up to order five. In 1915, MacMahon [13] approached the problem in a different way, but initially obtained the wrong value for order five. M.Frolov in 1890, [14] and Tarry in 1901, [15] [16] found the number of reduced squares of order six. M. Frolov gave an incorrect count of reduced squares of order seven. R.A. Fisher and F. Yates, [17] unaware of earlier work of E. Schönhardt, [18] gave the number of isotopy classes of orders up to six. In 1939, H. W. Norton found 562 isotopy classes of order seven, [19] but acknowledged that his method was incomplete. A. Sade, in 1951, [20] but privately published earlier in 1948, and P. N. Saxena [21] found more classes and, in 1966, D. A. Preece noted that this corrected Norton's result to 564 isotopy classes. [22] However, in 1968, J. W. Brown announced an incorrect value of 563, [23] which has often been repeated. He also gave the wrong number of isotopy classes of order eight. The correct number of reduced squares of order eight had already been found by M. B. Wells in 1967, [24] and the numbers of isotopy classes, in 1990, by G. Kolesova, C.W.H. Lam and L. Thiel. [25] The number of reduced squares for order nine was obtained by S. E. Bammel and J. Rothstein, [26] for order 10 by B. D. McKay and E. Rogoyski, [27] and for order 11 by B. D. McKay and I. M. Wanless. [28]

See also

Notes

  1. 1 2 3 4 5 6 Colbourn & Dinitz 2007 , p. 136
  2. 1 2 Dénes & Keedwell 1974 , p. 24
  3. 1 2 Dénes & Keedwell 1974 , p. 126
  4. Dénes & Keedwell 1974 , pp. 127-8
  5. 1 2 Dénes & Keedwell 1974 , p. 23
  6. McKay, Meynert & Myrvold 2007 , p. 105
  7. Dénes & Keedwell 1974 , p. 128
  8. 1 2 3 Dénes & Keedwell 1974 , p. 129
  9. McKay, Meynert & Myrvold 2007 , p. 100
  10. Colbourn & Dinitz 2007 , p. 137
  11. Euler, L. (1782), "Recherches sur une nouvelle espèce de quarrés magiques", Verhandelingen/Uitgegeven Door Het Zeeuwsch Genootschap der Wetenschappen te Vlissingen (9): 85–239
  12. Cayley, A. (1890), "On Latin Squares", Oxford Camb. Dublin Messenger of Math., 19: 85–239
  13. MacMahon, P.A. (1915), Combinatory Analysis, Cambridge University Press, p. 300
  14. Frolov, M. (1890), "Sur les permutations carrées", J. De Math. Spéc., IV: 8–11, 25–30
  15. Tarry, Gaston (1900). "Le Probléme de 36 Officiers". Compte Rendu de l'Association Française pour l'Avancement des Sciences. 1. Secrétariat de l'Association: 122–123.
  16. Tarry, Gaston (1901). "Le Probléme de 36 Officiers". Compte Rendu de l'Association Française pour l'Avancement des Sciences. 2. Secrétariat de l'Association: 170–203.
  17. Fisher, R.A.; Yates, F. (1934), "The 6 × 6 Latin squares", Proc. Cambridge Philos. Soc., 30 (4): 492–507, Bibcode:1934PCPS...30..492F, doi:10.1017/S0305004100012731, S2CID   120585553
  18. Schönhardt, E. (1930), "Über lateinische Quadrate und Unionen", J. Reine Angew. Math., 1930 (163): 183–230, doi:10.1515/crll.1930.163.183, S2CID   115237080
  19. Norton, H.W. (1939), "The 7 × 7 squares", Annals of Eugenics, 9 (3): 269–307, doi: 10.1111/j.1469-1809.1939.tb02214.x
  20. Sade, A. (1951), "An omission in Norton's list of 7 × 7 squares", Ann. Math. Stat., 22 (2): 306–307, doi: 10.1214/aoms/1177729654
  21. Saxena, P.N. (1951), "A simplified method of enumerating Latin squares by MacMahon's differential operators; II. The 7 × 7 Latin squares", J. Indian Soc. Agric. Statistics, 3: 24–79
  22. Preece, D.A. (1966), "Classifying Youden rectangles", J. Roy. Statist. Soc. Ser. B, 28: 118–130, doi:10.1111/j.2517-6161.1966.tb00625.x
  23. Brown, J.W. (1968), "Enumeration of Latin squares with application to order 8", Journal of Combinatorial Theory, 5 (2): 177–184, doi: 10.1016/S0021-9800(68)80053-5
  24. Wells, M.B. (1967), "The number of Latin squares of order eight", Journal of Combinatorial Theory, 3: 98–99, doi: 10.1016/S0021-9800(67)80021-8
  25. Kolesova, G.; Lam, C.W.H.; Thiel, L. (1990), "On the number of 8 × 8 Latin squares", Journal of Combinatorial Theory, Series A, 54: 143–148, doi: 10.1016/0097-3165(90)90015-O
  26. Bammel, S.E.; Rothstein, J. (1975), "The number of 9 × 9 Latin squares", Discrete Mathematics, 11: 93–95, doi: 10.1016/0012-365X(75)90108-9
  27. McKay, B.D.; Rogoyski, E. (1995), "Latin squares of order ten", Electronic Journal of Combinatorics, 2: 4, doi: 10.37236/1222
  28. McKay, B.D.; Wanless, I.M. (2005), "On the number of Latin squares", Ann. Comb., 9 (3): 335–344, doi:10.1007/s00026-005-0261-7, S2CID   7289396

Related Research Articles

In mathematics, the determinant is a scalar value that is a certain function of the entries of a square matrix. The determinant of a matrix A is commonly denoted det(A), det A, or |A|. Its value characterizes some properties of the matrix and the linear map represented, on a given basis, by the matrix. In particular, the determinant is nonzero if and only if the matrix is invertible and the corresponding linear map is an isomorphism.

<span class="mw-page-title-main">Gaussian elimination</span> Algorithm for solving systems of linear equations

In mathematics, Gaussian elimination, also known as row reduction, is an algorithm for solving systems of linear equations. It consists of a sequence of row-wise operations performed on the corresponding matrix of coefficients. This method can also be used to compute the rank of a matrix, the determinant of a square matrix, and the inverse of an invertible matrix. The method is named after Carl Friedrich Gauss (1777–1855). To perform row reduction on a matrix, one uses a sequence of elementary row operations to modify the matrix until the lower left-hand corner of the matrix is filled with zeros, as much as possible. There are three types of elementary row operations:

<span class="mw-page-title-main">Linear algebra</span> Branch of mathematics

Linear algebra is the branch of mathematics concerning linear equations such as:

<span class="mw-page-title-main">Quasigroup</span> Magma obeying the Latin square property

In mathematics, especially in abstract algebra, a quasigroup is an algebraic structure resembling a group in the sense that "division" is always possible. Quasigroups differ from groups mainly in that the associative and identity element properties are optional. In fact, nonempty associative quasigroup equals group.

<span class="mw-page-title-main">Latin square</span> Square array with symbols that each occur once per row and column

In combinatorics and in experimental design, a Latin square is an n × n array filled with n different symbols, each occurring exactly once in each row and exactly once in each column. An example of a 3×3 Latin square is

In mathematics, the concept of an inverse element generalises the concepts of opposite and reciprocal of numbers.

In order theory, a field of mathematics, an incidence algebra is an associative algebra, defined for every locally finite partially ordered set and commutative ring with unity. Subalgebras called reduced incidence algebras give a natural construction of various types of generating functions used in combinatorics and number theory.

<span class="mw-page-title-main">Transpose</span> Matrix operation which flips a matrix over its diagonal

In linear algebra, the transpose of a matrix is an operator which flips a matrix over its diagonal; that is, it switches the row and column indices of the matrix A by producing another matrix, often denoted by AT.

<span class="mw-page-title-main">Orthogonal group</span> Type of group in mathematics

In mathematics, the orthogonal group in dimension n, denoted O(n), is the group of distance-preserving transformations of a Euclidean space of dimension n that preserve a fixed point, where the group operation is given by composing transformations. The orthogonal group is sometimes called the general orthogonal group, by analogy with the general linear group. Equivalently, it is the group of n × n orthogonal matrices, where the group operation is given by matrix multiplication (an orthogonal matrix is a real matrix whose inverse equals its transpose). The orthogonal group is an algebraic group and a Lie group. It is compact.

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.

In abstract algebra, a matrix ring is a set of matrices with entries in a ring R that form a ring under matrix addition and matrix multiplication. The set of all n × n matrices with entries in R is a matrix ring denoted Mn(R) (alternative notations: Matn(R) and Rn×n). Some sets of infinite matrices form infinite matrix rings. A subring of a matrix ring is again a matrix ring. Over a rng, one can form matrix rngs.

In abstract algebra, the split-quaternions or coquaternions form an algebraic structure introduced by James Cockle in 1849 under the latter name. They form an associative algebra of dimension four over the real numbers.

Named after the 19th century British mathematician Arthur Cayley, a Cayley table describes the structure of a finite group by arranging all the possible products of all the group's elements in a square table reminiscent of an addition or multiplication table. Many properties of a group – such as whether or not it is abelian, which elements are inverses of which elements, and the size and contents of the group's center – can be discovered from its Cayley table.

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.

<span class="mw-page-title-main">Circulant graph</span> Undirected graph acted on by a vertex-transitive cyclic group of symmetries

In graph theory, a circulant graph is an undirected graph acted on by a cyclic group of symmetries which takes any vertex to any other vertex. It is sometimes called a cyclic graph, but this term has other meanings.

In abstract algebra, a bicomplex number is a pair (w, z) of complex numbers constructed by the Cayley–Dickson process that defines the bicomplex conjugate , and the product of two bicomplex numbers as

In linear algebra, the Hermite normal form is an analogue of reduced echelon form for matrices over the integers Z. Just as reduced echelon form can be used to solve problems about the solution to the linear system Ax=b where x is in Rn, the Hermite normal form can solve problems about the solution to the linear system Ax=b where this time x is restricted to have integer coordinates only. Other applications of the Hermite normal form include integer programming, cryptography, and abstract algebra.

In the mathematical field of abstract algebra, isotopy is an equivalence relation used to classify the algebraic notion of loop.

<span class="mw-page-title-main">Matrix (mathematics)</span> Array of numbers

In mathematics, a matrix is a rectangular array or table of numbers, symbols, or expressions, arranged in rows and columns, which is used to represent a mathematical object or property of such an object.

References