Abstract

Let denote the class of square matrices containing in each row and in each column exactly 1’s. The minimal value of , for which the behavior of the permanent in is not quite studied, is . We give a simple algorithm for calculation of upper magnitudes of permanent in and consider some extremal problems in a generalized class , the matrices of which contain in each row and in each column nonzero elements , , and and zeros.

1. Introduction

The definition of permanent of a square matrix of order is , where sum is over all permutations of numbers . This definition is very combinatorial. For example, if is -matrix, then is the number of arrangements of nonattacking each other rooks on positions of 1’s of . Therefore, the most natural applications of the permanent are in combinatorics: enumerating the Latin rectangles and squares, permutations with restricted positions, and different problems in the theory of graphs. Many interesting examples of applications of the permanent one can find in [1], chapter 8 (and not only in Mathematics!). To applications of the permanent there promoted celebrate proofs of the best known van der Waerden’s and Minc’s conjectures by Egorychev [2], Falikman [3], and Brègman [4], respectively. For example, Alon and Friedland [5] excellently used the Minc-Bregman inequality for permanent of -matrices for finding the maximum number of perfect matchings in graphs with a given degree sequence.

Let denote the class of square matrices containing in each row and in each column exactly 1’s. If , then matrix is doubly stochastic, since all its row and column sums equal 1. Therefore, -matrices are also called doubly stochastic (0,1)-matrices. These matrices have especially simple and attractive structure and have many applications. It is important that especially class in permanent’s history was a nontrivial test area for different researches of the permanent. For example, one can mention a remarkable Merriell research [6] of the maximum of permanent on class . An author’s research [7] of the permanent on -matrices (see Section 9) has led to the disproof of an important Balasubramanian conjecture [8] which would yield a full proof of 1967 Ryser hypothesis [9] on transversals of Latin squares.

In the present paper we consider many other problems for . We study also the following natural generalization of . For given real or complex nonzero numbers , , and , denote by the class of square matrices containing every number from exactly one time in each row and in each column, such that the other elements are 0’s. It is clear that, if and , then the matrix is doubly stochastic.

Definition 1. One calls -spectrum in (denoting it the set of all the values which are taken by the permanent in .

Note that -spectrum in trivially is . It is known (cf. Tarakanov [10]) that

But, for ,  -spectrum of , generally speaking, is unknown. Greenstein (cf. [1], point 8.4, Problem 3) put the problem of describing the -spectrum in . In this paper we give an algorithm for calculation of upper values of -spectrum in . We also obtain several results for a generalized class with real nonzero numbers , , and . Some results of the present paper were announced by the author in [11].

2. What Is Known about ?

(1) Explicit formula for (cf. Stanley [12], chapter 1) is

(2) Asymptotic formula for (cf. O’Neil [13]) is where is arbitrarily small for sufficiently large .

In addition, note that with different , , and is, evidently, the number of 3-rowed Latin rectangles of length such that where is the number of reduced 3-rowed Latin rectangles with the first row . It is known (Riordan [14], pp. 204–210) that where is subfactorial: and is sequence of Lucas numbers of the Ménage problem which is defined by Cayley recursion (cf. [14], p. 201) (see [15], sequences A102761, A000186).

Denote, furthermore, by the set of matrices in with 1’s on the main diagonal. Note that

Indeed, it is well known that every -matrix has a diagonal of ones (i.e., a set of 1’s not two in the same row or column). Let be such a diagonal. There exists a permutation of rows and columns such that will be the main diagonal of . Nevertheless, and (8) follows.

(3) A known explicit formula for (Shevelev [16]) has a close structure to (5): where sequence is defined by recursion

(4) Asymptotic formula for (Shevelev [16]) is where and is arbitrarily small for sufficiently large .

(5) Denote by the set of symmetric matrices in . -spectrum on is given by the following theorem (Shevelev [11]).

Theorem 2. Let denote the set of all partitions of with parts more than or equal to 3. To every partition , , put in a correspondence the number where sequence is defined by the recursion Then one has

(6) The maximal value of permanent in was found by Merriell [6].

Theorem 3. If , , then

Note that the case of (16) easily follows from the Minc-Bregman inequality for permanent of -matrices.

(7) Put . In case of , Bolshakov [17] showed that the second maximal of permanent in (such that interval is free from values of permanent in ) equals

Note that both and are attained in .

(8) Denote by the minimal value of permanent in . In 1979, Voorhoeve [18] obtained a beautiful lower estimate for :

This estimate remains the best even after proof by Egorychev [2] and Falikman [3], the famous van der Waerden conjectural lower estimate for every doubly stochastic matrix . Indeed, this estimate yields only , such that (18) is stronger for .

(9) Bolshakov [19] found -spectrum in in cases . Namely, he added to the evident -spectrums and also the following -spectrums:

3. A Generalization of Theorem 2 on Matrices of Class with Symmetric Positions of Elements

Denote by the set of matrices in with ’s on the main diagonal. It is clear that, together with (4),

Note that, as for sets , , we have

Denote, furthermore, by the set of matrices in with symmetric positions of elements: if and only if .

-spectrum on is given by the following theorem.

Theorem 4. If to every partition , , , corresponds the number where sequence is defined by the recursion then one has

Proof. Let be the symmetric permutation group of elements . Two positions , are called independent if , . We will say that in the matrix , a weight is appropriated to the position . Let have no cycle of length less than . Consider a map appropriating to the position the weight .
Now we need several lemmas.
Lemma A. (1) The map is bijective; (2) if is a set of pairwise independent positions, then is also a set of pairwise independent positions.
Proof. (a) Consider two distinct positions such that, at least, one of the two inequalities holds Let such that, say, . Show that . Indeed, if we suppose that ; then ;  that  is, has a cycle of length in spite of the condition. Conversely, if , then , since has not any cycle of length less than as well.
(b) Let positions (26) be independent. Both inequalities (27) hold and, as in (a), we have , ; that is, the positions , are independent as well.
Lemma B. Let have no cycle of length less than . Then -matrix having 1’s on only positionsis an incidence matrix of .
Proof. Since has no cycles of length less than , then is a permutation of numbers . Thus the set of positions of 1’s of matrix coincides with the set of 1’s of the incidence matrix of .
Let be -matrix with 1’s on positions only.
Lemma C. Let have not got any cycle of length less than . If and are the incidence matrices of and , then one hasProof. Both formulas follow from Lemma B.
Noting that , where is the identity matrix, we conclude that Moreover, since, by the bijection , to every diagonal (i.e., to every set of pairwise independent positions) of the matrix corresponds one and only one diagonal of the matrix with the same products of weights, then we have Note that from the definition it follows that, for every matrix , we have a representation where is the incidence matrix of a substitution . In case when has not got any cycle of length less than , the matrix is a completely indecomposable matrix in . Thus, by (31), all completely indecomposable matrices of have the same permanent, equal to .
In general, a substitution with the incidence matrix in (32) cannot have cycles of length less than 3. Indeed, if, for some , we have either or , then in both cases which means coincidence of positions 1’s of the matrices and in the th row.
Let be an arbitrary substitution with cycles of length more than 2. Let where , , and , be the decomposition of in a product of cycles. Then the matrix is a direct sum of the matrices such that, by (31), and we have It is left to notice that Minc [20] found a recursion (23) for and, as well known, the multiplication of an matrix by does not change its permanent. Therefore, .

Example 5. Let us find .

We have the following partitions of 11 with the parts not less than 3: According to (23), for , we have , and for

Using induction, we find

Therefore,

In the following examples we calculate -spectrum for arbitrary .

Example 6. Let us find .

By induction, for , we have

Further, again using induction, one can find that, if is even, then and, if is odd, then

Example 7. Analogously, in case of , for , we have

It is interesting that, in case of multiple of 3, the permanent omits the value .

4. Merriell-Type Theorems in a Subclass of

Note that in class the Minc-Bregman inequality and the Merriell theorem, generally speaking, do not hold even for positive , ,  and  . Nevertheless, some restrictions on , ,  and   allow proving some analogs of the Merriell theorem. Recall that (16) is attained in . Denote by the maximal value of permanent in .

Theorem 8. Consider a class with the numbers , , and satisfying “triangle inequalities” and the following additional conditions: where sequence is defined by recursion (23). Then, for multiple of 3, one has

Proof. Note that conditions (44)-(45) are satisfied, for example, in case . Using induction, let us prove that
Indeed, for , this inequality is trivial, while, for , it follows from the first condition (45). Let it hold for . Then, according to (23), we have
Note that, according to Theorem 4, the equality in (47) holds in a direct sum of -matrices of which corresponds to the partition . Let now . By Theorem 4, there exists a partition of with the parts not less than 3, , such that and, in view of (47), we have This proves (46).

Example 9. Consider case .

Let us find the values of , depending on the magnitude of , for which the conditions of Theorem 8 are satisfied. According to (23), we have

Thus the condition means that or And it is easy to verify that the second condition in (45) is satisfied as well. As a corollary, we obtain the following result.

Theorem 10. If (52) holds, then, for multiple of 3, one has

Simple forms of sequence in Examples 5 and 6 allow supposing that in case (or symmetrical case ) sequence keeps a sufficiently simple form. We find this form in the following lemma.

Lemma 11. If , then sequence which is defined by recursion (23) has the form

Proof. Using induction with the base (51), suppose that (54) holds for . Then, by (23), for even , we have while, if is odd, then we have

Let now

Theorem 12. If (57) holds, then, for multiple of 4, one has

Proof. From (51) and (57), we conclude that
Let us show that, for ,
For , inequality (60) is trivial. For , we have and thus, using Lemma 11, we have
Let now . By Theorem 4, there exists a partition of with the parts not less than 3, , such that and, in view of (60), we have with the equality in a direct sum of -matrices of which corresponds to the partition .

Note that, if , then in Theorem 10 we have only maximizing matrix (up to a permutation of the rows and columns) which corresponds to the partition ; in Theorem 12 we also have only maximizing matrix (up to a permutation of the rows and columns) which corresponds to the partition . It is interesting that, only in case of the equality , where , Theorems 1012 are true for every multiple of 12 with the equality of the maximums . Thus, up to a positive factor ,  the class possesses an interesting extremal property: it contains  maximizing matrices (up to a permutation of the rows and columns), instead of only maximizing matrix, if .

Indeed, the number of the maximizing matrices (up to a permutation of the rows and columns) is defined by the number of the following partitions of (mod 12):

5. Estimate of Cardinality of -Spectrum on Circulants in and

Denote by the set of the circulants in . Note that a circulant has a form , , where is -matrix with 1’s on positions only. Multiplying by we obtain circulant of the form with . Since is defined by a choice of two different values , then trivially

Now we prove essentially more exact and practically unimprovable estimate.

Theorem 13. One has

Proof. Let us return back to the general form
Note that is defined by a choice of a vector , but its rotation, that is, a passage to a vector of the form , does not change the magnitude of . Indeed it corresponds to the multiplication by , and our statement follows from the equality . Besides, its reflection relative to some diameter of the imaginary circumference of the rotation, by the symmetry, keeps magnitude of the permanent. Since geometrically three points on the imaginary circumference define a triangle, then our problem reduces to a triangle case of the following general problem, posed by Professor Richard H. Reis (South-East University of Massachusetts, USA) in a private communication to Hansraj Gupta in 1978:

Let a circumference is split by the same parts. It is required to find the number of the incongruent convex -gons, which could be obtained by connection of some from dividing points. Two -gons are considered congruent if they are coincided at the rotation of one relatively other along the circumference and (or) by reflection of one of the -gons relatively some diameter.


In 1979, Gupta [21] gave a solution of the Reis problem in the form
If we denote by the set of the circulants in , then from our arguments it follows that In case , from (71)-(72) we find and (69) follows.

Example 14. In case we have only two incongruent triangles corresponding to circulants and .

Nevertheless, the calculations give . Thus , and .

Example 15. In case we have three incongruent triangles corresponding to circulants , , and .

The calculations give and ,  while . Thus , and .

Note that a respectively large magnitude of is explained by its decomposability in a direct product of circulants , such that .

It is clear that, in case of circulants in , the upper estimate (69) yields either the same estimate, if , or , if (and in the symmetric cases), or , if , , and are distinct numbers.

Add that, using a bijection, one can apply formula (71) to enumerate the two-color bracelets of beads, of which are black and are white (see, e.g., the author’s explicit formulas for sequences A032279-A032282, A005513-A005516 in [15]).

6. Algorithm of Calculations of Upper Magnitudes of the Permanent in

Theorem 2 allows, using some additional arguments, giving an algorithm of explicit expressions of upper magnitudes of the permanent in via numbers (14). For that we need the following lemma.

Lemma 16. For , one has

Proof. By the usual way, from (14) we find where is the golden ratio.
Denote . Since , then , and, consequently, if , then . Therefore, we have

Note that, actually, the difference between the hand sides of (76) is more than .

Let now , , 1, and 2, and . Let denote the set of all partitions of with parts more than or equal to . For us an important role play cases , 4. To , put in a correspondence the sets

In case , when , let us agree that is a singleton .

Consider now the set :

Theorem 17 (algorithm of calculation of upper magnitudes of the permanent in ). If , then the set taken in decreasing order of its elements gives upper magnitudes of the permanent in .

Proof. Note that the proof is the same for every value of . Therefore, let us consider, say, . From (74) it follows that, if contains parts 3 and , then, for , we have This means that, for the formation of the list of all upper magnitudes of the permanent in in the condition , which are bounded from below by , it is sufficient to consider only a part of the spectrum containing numbers , where with the opposite condition . From the equality with the condition , we have Since does not depend on , there is only a finite assembly of such partition for arbitrary . This ensures a possibility of the realization of the algorithm.
For the considered , for , we have , where , and has the form , . Thus we should choose only , and this yields

In order to use Theorem 17 for calculating the upper magnitudes of the permanent in , in case, say, ,(1)we write a list of partition of numbers , , with the parts not less than 4;(2)we compare the corresponding values of with and keep only ;(3)after that, we take in decreasing order the numbers .

Below we give the first 10 upper magnitudes of the permanent in , for , via numbers (14). Consider

Formula (82) shows that is attained in . Moreover

7. Main Conjectural Inequality for Maximum of Permanent in Completely Indecomposable -Matrices

Denote by the set of completely indecomposable matrices in , that is, the set of -matrices containing no -submatrices. Let denote the maximum of permanent in . Our very plausible conjecture which we call “main conjectural inequality (MCI)” is the following.

Conjecture 18 (cf. Shevelev [11], pp. 165-166). For , one has

In Lemma 16 we essentially proved that in subclass the MCI is valid.

Besides, in all known cases MCI holds. Moreover, as we shall see, our algorithm of calculation of the consecutive upper magnitudes of permanent in , which is based on MCI, reproduces all Merriell’s and Bolshakov’s results for and . Note also that, for sufficiently large , the number of consecutive upper magnitudes of permanent in grows very quickly with every step of extension of the list of known -spectrums for small . For example, using the found by Bolshakov , , we obtain, for sufficiently large , 4, 7, and 11 upper values of in cases , , and respectively. Further, after calculating , the number of upper values of, for example, increases by more than thrice.

8. Algorithm of Calculations of Upper Magnitudes of the Permanent in Based on MCI

Let , , 1, 2, and . Let denote the set of all partitions of with parts more than or equal to . For us an important role play cases . To , put in a correspondence sets where runs through all values of permanent in set of completely indecomposable matrices in (in case , when , let us agree that is a singleton ).

Consider now the set :

Theorem 19 (algorithm of calculation of upper magnitudes of the permanent in ). If , then the set taken in decreasing order of its elements gives upper magnitudes of the permanent in .

Proof. We need three lemmas.

Lemma 20. For , one has

Proof. Let, firstly, . Note that . Using (92), we find
Let, furthermore, , , 2, 3. Note that, by (19), . Therefore, using (92), we have

Lemma 21. Let be a partition of with the parts not less than 3. If and has the form , then, for completely indecomposable matrices , , one has

Proof. Using Lemma 20, we have

Lemma 22. Let and where is a nonnegative integer and is the residue of modulo 3,  , 1, 2; then, for completely indecomposable matrices , , one has

Proof. Put , . Now the lemma follows from Lemma 21.

It is left to note that, after these lemmas, the proof of Theorem 19 is the same as proof of Theorem 17.

Note that the use of this algorithm is based on the upper values of -spectrum for smaller values of .

Consider, for example, case , . According to (94), we have

Note that the second set is a simpleton, since, by MCI, . Since, by (19), , then the first set in (102) is empty. Thus is simpleton: and we have which corresponds to Merriell’s result in case .

Further research of the set (94), using (19), gives the following results.

(1) Consider , :

The continuation of this list requires knowing . Note that a more detailed analysis shows that, after calculation of , in this case one can obtain the first upper magnitudes of the permanent in , where .

(2) Consider , :

It is interesting that in this case is not used up to , but the continuation of this list requires knowing .

(3) Consider , :

Note that the method not only gives a possibility to calculate the upper magnitudes of the permanent in , but also indicates those direct products on which they are attained. For example, in (107) is attained on direct products of some matrices

Note also that the comparison of (105)–(107) with (82)–(91) shows that the following calculated are attained in ,

in case , in case , (but is not attained);

in case , (but , , , and are not attained).

9. Algorithm for Testing the Parity of Values of the Permanent in

It seems that, among all known methods of calculation of the permanent, only Ryser’s method (cf. [1], chapter 7) could be used for creating an algorithm for testing the parity of values of the permanent. Let be -matrix. Let be a matrix which is obtained by changing some columns of by zero columns. Denote by the product of row sums of . Then, by Ryser’s formula, we have

Let now have integer elements. Introduce the following matrix function:

From (112) we have

Using (114), let us create an algorithm of a search of the odd values of the permanent in . Since, evidently, , then should have pairwise distinct columns. Note that cases (mod 3), , are considered by the same way. Suppose, say, . According to (114), we are interested in only cases when, after removing columns of , all row sums will be odd. Suppose that, after removing columns of , we have that sums remain equal to 3 and sums equal to 1. This means that the total number of the removed 1’s equals . Since, removing a column, we remove three 1’s, then the number of the removed columns equals . Thus and , . However, if , then . By the condition, these two columns are distinct; therefore, we conclude that at least one row sum equals 2. The contradiction shows that the testing sequence is . In cases we obtain the same testing sequence.

Example 23. Let us check the parities of values of the permanent of circulants in .

In this case and, therefore, the testing sequence contains only term . Note that matrix has all odd rows if and only if one row sum equals 3 and each of the 6 other row sums equals 1. Indeed, let, after removing 4 columns of , sums remain equal to 3 and sums equal to 1. This means that the total number of the removed 1’s equals and the number of the removed columns equals ; that is, . Moreover, since in a circulant all rows are congruent shifts of the first one, it is sufficient to consider the case when precisely the first row sum equals 3 and others equal 1 (the multiplication on 7 does not change the parity of the result). This opens a possibility of a momentary handy test on the parity every circulant of class . This test consists of removing all four columns beginning with 0. If now every row has one 1, then the permanent is even; otherwise, it is odd. We check now directly that from circulants exactly 21 ones have odd permanent.

Remark 24. In 1967, Ryser [9] did a conjecture that the number of the transversals of a Latin square from elements (i.e., the number of subsets of its pairwise distinct elements, with none in the same row or column) has the same parity as . If is even, then the conjecture has been proved by Balasubramanian [8]. Besides, in [8] Balasubramanian posed a conjecture for the parity of a sum of permanents, such that the truth of this conjecture yields Ryser’s hypothesis for odd . In the same year (1990), using the result of Example 23, the author disproved Balasubramanian’s conjecture. It is interesting that soon Parker (see [22], p. 258) indeed found several Latin squares of order 7 with even number of transversals. Add that later ([7]) we found even an infinite set of counterexamples to the Balasubramanian conjecture.

10. Open Problems

(1)To prove the MCI (Section 7).(2)Consider class , where (cf. [23], pp. 171-172). Since , then Voorhoeve’s lower estimate for permanent of -matrices trivially holds for matrices in . It is clear that, for , an essentially stronger lower estimate should exist. However, using van der Waerden-Egorychev-Falikman theorem to class of doubly stochastic matrices, for the permanent of -matrices, we obtain even weaker lower estimate of the order in the case . The problem is to find a stronger lower estimate for the permanent in in this case.(3)Let be a circulant of order with integer elements (cf. [23], pp. 115-116). We conjecture that, for every integer , we have , where -matrix consists of 1’s only. In [7] the conjecture was proved for every and prime . Thus the question is open in case of arbitrary composite number .(4)Let us call two Latin rectangles equivalent if the sets of their elements in the corresponding columns are the same. Note that one can treat numbers as the numbers of equivalence classes of Latin triangles. Let . It could be proved that the cardinality of the corresponding equivalent class is . To find the cardinality of the equivalent class which is defined by matrix .

Acknowledgment

The author is grateful to both of the anonymous referees for several suggestions which promoted improvement in the text.