Orthogonal realizations of random sign patterns and other applications of the SIPP
Abstract
A sign pattern is an array with entries in . A matrix is row orthogonal if . The Strong Inner Product Property (SIPP), introduced in [B.A. Curtis and B.L. Shader, Sign patterns of orthogonal matrices and the strong inner product property, Linear Algebra Appl. 592: 228–259, 2020], is an important tool when determining whether a sign pattern allows row orthogonality because it guarantees there is a nearby matrix with the same property, allowing zero entries to be perturbed to nonzero entries, while preserving the sign of every nonzero entry. This paper uses the SIPP to initiate the study of conditions under which random sign patterns allow row orthogonality with high probability. Building on prior work, nowhere zero sign patterns that minimally allow orthogonality are determined. Conditions on zero entries in a sign pattern are established that guarantee any row orthogonal matrix with such a sign pattern has the SIPP.
keywords:
Sign pattern, Orthogonality, Row orthogonal matrix, Strong Inner Product Property, SIPP, Random matrix, High probability.15B10, 15B35, 15B52, 60B20.
1 Introduction
A sign pattern is an array with entries coming from the set . The entries of sign patterns encode qualitative properties of real matrices. Sign patterns were introduced in applications where the entries of the matrix may be known only approximately (or not at all), but the signs of the entries are known. A matrix is row orthogonal provided . The problem of determining whether an sign pattern allows row orthogonality has been studied for many years [7, 10, 6, 5]. Recently the strong inner product property (SIPP) was introduced by Curtis and Shader in [6] to study sign patterns of row orthogonal matrices. This paper relies heavily on the SIPP to build on prior work (e.g., classifying small patterns that minimally allow orthogonality) and initiate the study of conditions under which random sign patterns allow row orthogonality with high probability.
Finding a certificate that a sign pattern allows row orthogonality is often difficult. By applying a variant of Gram-Schmidt orthogonalization to a nowhere zero nearly row orthogonal matrix we obtain conditions that guarantee the existence of a nearby row orthogonal matrix with the same sign pattern (see Section 2). In Section 3, we apply the SIPP to develop new tools to show that a sign pattern allows row orthogonality and use these tools (and the results from Section 2) to determine nowhere zero sign patterns that minimally allow orthogonality. We also establish conditions on zero entries in a sign pattern that guarantee any row orthogonal matrix with such a sign pattern has the SIPP. One of our main results, Theorem 4.12, utilizes the SIPP to obtain a lower bound such that the probability of a random sign pattern allowing row orthogonality goes to as tends toward for (here random means and are equally likely and the probability of is given). The remainder of this introduction defines terminology and notation (Section 1.1) and lists known results we will use (Section 1.2).
1.1 Definitions and notation
In the study of sign patterns, sometimes the distinction between zero and nonzero is more important than the sign. A zero-nonzero pattern or znz pattern is an array with entries coming from the set . We use the term pattern to mean a sign pattern or a zero-nonzero pattern. Given a real number ,
The sign pattern and zero-nonzero pattern of a matrix are and , respectively. The qualitative class of an sign pattern is the set
and the qualitative class of an znz pattern is the set
A matrix in the qualitative class is called a realization of the pattern . For a sign pattern , denotes the unique -matrix that is a realization of the sign pattern . Similarly, is the unique -matrix that is a realization of the zero-nonzero pattern . A superpattern of a sign pattern is a sign pattern of the same dimensions such that whenever ; if then .
A matrix with orthogonal rows is not necessarily row orthogonal; for us, the rows of a row orthogonal matrix have unit length. The set of row orthogonal matrices is denoted by and we write as shorthand for . Note that every matrix is orthogonal, i.e., . The set of real symmetric matrices is denoted by .
A zero matrix or zero vector has every entry equal to zero. An matrix or pattern is wide if . A wide matrix has full rank if its rank equals its number of rows, i.e., it has linearly independent rows. A row orthogonal matrix is necessarily wide. Let be a wide matrix. Then has the strong inner product property (SIPP) provided is the only symmetric matrix satisfying [5]. The strong inner product property is one of a number of strong properties of matrices that guarantee there is a nearby matrix with the same property, allowing zero entries to be perturbed to nonzero entries, while preserving the sign of every nonzero entry [8, Part 2].
An sign pattern allows row orthogonality if there is a row orthogonal matrix (equivalently, ). An sign pattern allows o-SIPP if there is a row orthogonal matrix that has the SIPP. Since scaling a matrix with a positive constant does not change its pattern, no pattern requires row orthogonality. A sign pattern requires o-SIPP if every has the SIPP and . Without the assumption that , the all zeros pattern would vacuously require o-SIPP.
An rectangular sign pattern is row potentially pairwise-orthogonal or row PPO if no row is a zero vector and for each pair with , there are realizations of row and row that are orthogonal. The term column PPO is defined analogously. A pair of rows and in a rectangular sign pattern has a negative -cycle if there are two columns and such that and , where multiplication on the set is defined in the obvious way that conforms to real arithmetic. A pair of rows and in an matrix or pattern is combinatorially orthogonal if implies for every .
A signed permutation matrix is a square -matrix with exactly one nonzero entry in each row and column. Matrices are sign equivalent if , where and are signed permutation matrices. Two sign patterns and are sign equivalent if and are sign equivalent.
For a vector , the support of , denoted by , is the set of indices of nonzero entries of . Let .
1.2 Known results
In the remainder of this introduction we provide some known results about the SIPP that we will use. The primary motivation for developing the SIPP is given by the next theorem of Curtis and Shader.111Theorem 4.5 in [6] actually says that every superpattern of allows row orthogonality, but the proof shows it allows o-SIPP. We provide a slightly stronger result in Theorem 3.13.
Theorem 1.1.
[6] If has the SIPP and , then every superpattern of allows o-SIPP.
Theorem 1.1 has many consequences. Here we list some that we use. A matrix with the SIPP or a sign pattern that allows the SIPP can be padded with additional zero columns and retain that property.
Lemma 1.2.
[6] Let and . Then has the SIPP if and only if the matrix has the SIPP.
Corollary 1.3.
[6] If has the SIPP and , then allows o-SIPP.
The next two results show that sign equivalence preserves having the SIPP, as does taking the transpose for (square) orthogonal matrices.
Proposition 1.4.
[6] Let be sign equivalent. Then has the SIPP if and only if has the SIPP.
Proposition 1.5.
[6] Let . Then has the SIPP if and only if has the SIPP.
The previous results provide some sufficient conditions for a sign pattern to allow row orthogonality. The next result provides a way to show a sign pattern does not allow row orthogonality.
Theorem 1.6.
[10] Let be a nowhere zero sign pattern and let be an submatrix of . If and , then does not allow row orthogonality.
2 From approximate orthogonality to exact orthogonality
In this section, we establish a result that gives conditions under which a collection of “nearly” orthogonal vectors necessarily implies the existence of a “nearby” collection of truly orthogonal vectors. Such a result is similar in spirit to the effective implicit function theorems used by, e.g., Cohn, Kumar and Minton [4] to derive the existence of an exact code from an approximate one. However, instead of using an implicit function theorem, we simply rely on the Gram–Schmidt process. Although the perturbations here are created by a different mechanism, we also point out that the idea of perturbing one solution to obtain another desired solution is a fundamental idea underlying strong properties. First, we define the function used to quantify the notion of “nearby.”
Definition 2.1.
For an integer and a real number ,
where is interpreted as , so for all .
For simplicity, we have defined the functions in closed form. In order to use these functions for the results of this section we need a recursive approach, which is given in the next lemma.
Lemma 2.2.
Given , can be computed recursively for all and all by
Proof 2.3.
It is easy to verify the result for . For ,
The next lemma provides the key step to go from approximately to exactly orthogonal.
Lemma 2.4.
Let be a positive integer, let and fix any inner product space . Additionally, let be any norm on (possibly unrelated to ). If satisfy
-
1.
for all , and
-
2.
for all ,
then there exists satisfying
-
1.
is orthonormal with respect to , and
-
2.
for all .
Proof 2.5.
We prove the result by induction on . The case is immediate by taking . Let be a positive integer and suppose the statement holds for .
Without loss of generality, for all . For , let and ; then . Since and , we know that . Therefore, the Pythagorean Theorem allows us to conclude that
(2.1) |
Since , we know that is non-zero. Let denote the unit vector in the direction of , i.e., .
Since , we see that . Then the triangle inequality applied to implies that . Together with (2.1), we have
(2.2) |
In particular,
(2.3) | |||||
where the second inequality follows by combining (2.1) and .
Now, for any other with , we have
Therefore, since are unit vectors by construction, we may apply the induction hypothesis to find an orthonormal set such that
for each . By invoking additionally (2.2) and (2.3), we bound
for all where the last equality follows from Lemma 2.2.
Finally, let . Then satisfy the claim, because by construction, and the former subspace is orthogonal to .
Observe that the process used to create the vectors is a reordering of the modified Gram-Schmit process. We stated Lemma 2.4 very generally in the hopes that other researchers will find it useful; for our uses, we specialize to the standard Euclidean inner product and the -norm to attain a result related to row orthogonal realizations.
We apply Lemma 2.4 to obtain Theorem 2.7, which will be used in Section 3.3 to characterize nowhere zero sign patterns that minimally allow orthogonality. Essentially this result says that for any matrix that is close to being row orthogonal, there exists a nearby matrix that is row orthogonal and has the same sign pattern.
Definition 2.6.
For a non-zero vector , define
Theorem 2.7.
Let be any non-zero vectors and let where is the standard Euclidean inner product. If
-
1.
, and
-
2.
,
then there exists an orthogonal set satisfying for all .
Proof 2.8.
We apply Lemma 2.4 to the vectors , specializing to the Euclidean inner product and the -norm, to locate an orthonormal set such that
for all . In particular, setting for each , we know that is an orthogonal set and that
Since for any , we conclude that for all .
One particularly useful feature of Theorem 2.7 is that it can be used to present reasonable certificates of the existence of row orthogonal realizations; in fact, it implies that integer-valued certificates can always be found. We illustrate this in the following example (which will be used in Section 3.3).
Example 2.9.
Consider the sign-pattern
Explicitly writing down a row orthogonal realization of would be difficult since this requires exact arithmetic. Despite this, it is not too difficult for a computer to find realizations of that are row orthogonal up to floating-point error. For example, the following matrix is such a realization for :
Of course, is not actually a row orthogonal matrix and so it does not directly demonstrate that has a row orthogonal realization; however, does satisfy the hypotheses of Theorem 2.7. In fact, by scaling and truncating appropriately, we find the following integer-valued matrix which satisfies the hypotheses of Theorem 2.7 as well:
Here and in similar examples, we use to denote where the are the rows of the matrix. To apply Theorem 2.7 to the matrix , observe that the value is obtained from row of and the value is obtained from rows and . Since is increasing on its domain, . We may therefore apply Theorem 2.7 to conclude that there exists a row orthogonal matrix with the same sign-pattern as .
We will use these same basic ideas in Section 3.3 to write down reasonable certificates for the existence of row orthogonal realizations for other sign patterns.
3 Results on the SIPP
In this section we present results related to the SIPP and orthogonality. Section 3.1 contains some useful tools for studying matrices that have the SIPP. Of particular interest is Theorem 3.13 which extends Theorem 1.1. In Section 3.2 we investigate sign patterns that require o-SIPP. Section 3.3 utilizes the SIPP to provide a complete characterization of nowhere zero sign patterns that minimally allow orthogonality for .
3.1 Tools
Recall that an matrix has the SIPP provided is the only symmetric matrix satisfying . It is often much easier to construct a matrix with orthogonal rows as opposed to a row orthogonal matrix. The next lemma allows us to study row orthogonal matrices with the SIPP without first normalizing the rows.
Lemma 3.1.
Suppose is an full rank matrix with orthogonal rows that has the SIPP and is any diagonal matrix with every diagonal entry nonzero. Then has the SIPP. Furthermore, can be chosen so that is row orthogonal.
Proof 3.2.
Let such that . By algebraic manipulation,
Since and has the SIPP, it follows that , which implies . Thus has the SIPP. Let denote the th row of . Define and , so .
The next three lemmas showcase additional hypotheses on that imply various entries in are .
Lemma 3.3.
Let be a wide matrix with full rank and let . Suppose that every entry of row of is nonzero. If , then every entry of row of is zero.
Proof 3.4.
Suppose . Let denote the rows of . Then implies that . Since has full rank there exists a matrix such that and so .
Lemma 3.5.
Suppose , and . Then .
Proof 3.6.
Let and write , , and . Since is row orthogonal, . The condition that implies that if . Therefore,
In other words, .
Lemma 3.7.
Let and satisfy the equation . Suppose that the only two nonzero entries in column of are and . Then .
Proof 3.8.
The next result extends one direction of [6, Proposition 3.9].
Lemma 3.9.
Suppose is a wide matrix partitioned as a block matrix with both nowhere zero (or vacuous). If has the SIPP and is full rank, then has the SIPP.
Proof 3.10.
Let be a symmetric matrix such that . By Lemma 3.3, . Therefore, . The equation implies that . Since is symmetric and has the SIPP, and has the SIPP.
Manifold theory, and in particular having manifolds interesect transversally, plays a fundamental role in strong properties, including the SIPP; see [8] for more information. Smooth manifolds and , both in , intersect transversally at a point if and the intersection of the normal spaces of at and of at contains only the zero vector. As the next result shows, a matrix having the SIPP is equivalent to and intersecting transversally at .
Theorem 3.11.
[6, Theorem 4.5] Let have sign pattern . The manifolds and intersect transversally at if and only if has the SIPP. If has the SIPP, then every superpattern of allows o-SIPP.
Theorem 3.13 improves the previous result by allowing us to control the relative magnitudes of the formerly zero entries in when applying the SIPP. This requires the following theorem of van der Holst, Lovász, and Schrijver.
Theorem 3.12.
[9] Let and be smooth families of manifolds in , and assume that and intersect transverally at . Then there is a neighborhood of the origin and a continuous function such that and for each , and intersect transversally at .
Note that the statement of Theorem 3.12 applies to a more general setting than we require. For our purposes, one of the smooth families of manifolds is replaced with a manifold. In such a setting we may think of as a continuous function of one variable from an interval about the origin to .
Theorem 3.13.
Let have sign pattern . If has the SIPP, then for all with , for every sufficiently small there is a matrix such that . Moreover, has the SIPP.
Proof 3.14.
Suppose that has the SIPP and let satisfy . Define the smooth family of manifolds by
for . Since has the SIPP, and intersect transversally at by Theorem 3.11. By Theorem 3.12, there exists a continuous function such that and the manifolds and intersect transversally at for each sufficiently small. Since is continuous we may choose small enough so that , where is the zero-nonzero pattern of and is the unique -matrix in . To complete the proof, observe that . Moreover, has the SIPP by Theorem 3.11 since and intersect transversally at .
We apply the previous theorem to prove the next result.
Proposition 3.15.
Let
be a sign pattern that allows row orthogonality, and let be a submatrix of with the same number of rows as . If allows o-SIPP, then
allows row orthogonality.
Proof 3.16.
Let be a row orthogonal realization of . Then
where the partition is consistent with that of . Assume that allows o-SIPP. Then there exists a row orthogonal realization of with the SIPP. Then is row orthogonal and, by Lemma 1.2, has the SIPP. By Theorem 3.13, there exists an and a matrix such that is row orthogonal and . Since is row orthogonal, . Therefore,
is row orthogonal and .
3.2 Sign patterns requiring o-SIPP
In this section we present results concerning sign patterns that require o-SIPP. As we shall see, both the number of zero entries and the location of the zero entries in a sign pattern play an important role in determining whether requires o-SIPP.
While sign patterns that require o-SIPP have not been previously studied, there are some known results that are closely related to requiring o-SIPP. For example, consider the lower Hessenberg matrix
which has orthogonal rows. The proof of Corollary 5.2 in [6] implies that any sign pattern that has the same zero-nonzero pattern as and allows orthogonality requires o-SIPP.
The next lemma provides an example of a structural property that guarantees a matrix has the SIPP and is used to establish Corollary 3.19, which is a slightly more general result than [6, Corollary 5.2]. For any integer , the -th diagonal of an matrix is the list of entries such that . The -th diagonal terminology is also applied to sign patterns.
Lemma 3.17.
Let be a wide matrix with full rank. Suppose that there is an integer such that , each entry of on the -th diagonal is nonzero for , and each entry of on the -th diagonal is zero for . Then has the SIPP.
Proof 3.18.
Note that if , then is nowhere zero and hence has the SIPP. Suppose that . Let and for . We begin by successively showing that each has the SIPP. Since contains a nonzero entry, Lemma 3.9 implies has the SIPP. Suppose that has the SIPP for some . Then Lemma 3.9 and Lemma 1.2 imply that has the SIPP. If , then has the SIPP by Lemma 1.2. Otherwise, , where is nowhere zero. By Lemma 3.9, has the SIPP.
Corollary 3.19.
Let be an wide sign pattern. Suppose that there is an integer such that , each entry of on the -th diagonal is nonzero for , and each entry of on the -th diagonal is zero for . If allows row orthogonality, then requires o-SIPP.
For sign equivalent matrices and , implies and has the SIPP implies has the SIPP. Thus the analogous statement with the upper part nonzero is also true.
Corollary 3.20.
Let be a wide sign pattern. Suppose that there is an integer such that , each entry of on the -th diagonal is nonzero for , and each entry of on the -th diagonal is zero for . If allows row orthogonality, then requires o-SIPP.
In this paper, a nonzero hollow matrix (respectively, sign pattern) is a square matrix (respectively, sign pattern) with zeros along the main diagonal and nonzero entries everywhere else. Recall that a signature matrix is a diagonal matrix with diagonal entries equal to . Matrices and are signature equivalent if there exist signature matrices and such that . Similarly, sign patterns and are signature equivalent if there exist signature matrices and such that . Theorem 5.7 in [6] states that a nonzero hollow matrix has the SIPP if and only if is not signature equivalent to a symmetric hollow matrix. The following corollary is an immediate consequence.
Corollary 3.21.
Let be a nonzero hollow sign pattern that allows orthogonality. If is not signature equivalent to a symmetric hollow sign pattern, then requires o-SIPP. If is signature equivalent to a symmetric hollow sign pattern, then does not allow o-SIPP.
As an example, consider
It is not difficult to verify that has orthogonal rows, and hence allows orthogonality. Further, is not signature equivalent to a symmetric sign pattern. Thus, requires o-SIPP.
Let be a graph with vertex set and edge set . The (vertex-edge) incidence matrix of is the matrix that has if and otherwise. An orientation of is the assignment of a direction to each edge. That is, the edge is replaced by exactly one of the two arcs or ; the arc associated with is denoted by . The incidence matrix of an orientation of is the matrix that has if , if , and otherwise.
Consider the complete graph and an orientation . For , define the matrix and its sign pattern . The sign pattern was shown to have a row orthogonal realization with the SIPP in [5]. We now show that requires o-SIPP. We note that this sign pattern will be instrumental in Section 4 for studying random sign patterns that allow row orthogonality.
Theorem 3.22.
For and any orientation of , the sign pattern requires o-SIPP.
Proof 3.23.
Let . It is an immediate consequence of Theorem 1.1 that if has a pair of combinatorially orthogonal rows, then does not have the SIPP. Similarly, if and has a pair of combinatorially orthogonal columns, then does not have the SIPP. Thus, when studying sign patterns that require o-SIPP, it is natural to assume that the rows (and in the square case the columns) are not cominbatorially orthogonal. Note that it is possible for a wide sign pattern to have combinatorially orthogonal columns and still require o-SIPP. For example, consider
Observe that has orthogonal rows. As we shall see, Corollary 3.27 implies requires o-SIPP.
The remainder of this section investigates how restricting the location and number of zero entries affects whether or not a sign pattern requires o-SIPP. Let be a wide matrix. It is not difficult to verify that if is nowhere zero, then has the SIPP. Thus every nowhere zero sign pattern that allows row orthogonality requires o-SIPP. For each additional entry in , the equation imposes one fewer linear equation on the entries of . This suggests that the more entries has, the larger the solution space to is going to be, reducing the likelihood that has the SIPP. In fact, this is the intuition behind the next theorem, which bounds the number of zero entries a matrix with the SIPP can have.
Theorem 3.24.
[6] Let have the SIPP. Then the number of zero entries in is at most .
The location of entries in a sign pattern also play an important role in determining if requires o-SIPP.
Theorem 3.25.
Let . Suppose that the zero entries of are contained in at most three rows of and that no pair of rows is combinatorially orthogonal. Then, has the SIPP.
Proof 3.26.
Begin by assuming that the zero entries of are contained in at most 1 row. Without loss of generality, the zero entries of are contained in the first row and the -entry of is nonzero. By Lemma 3.9, has the SIPP.
Now assume that exactly rows of contain a zero, where . Without loss of generality, the first rows each contain a zero. Let and suppose . By Lemma 3.3, , where . Observe that , where is the submatrix of formed from the first rows. Also, note that Lemma 3.5 implies .
First, consider the case . Since the rows of are not combinatorially orthogonal, has a nowhere zero column. Since , Lemma 3.7 implies that the off-diagonal entries of are zero. This, together with , implies .
Now consider the case . Suppose first that an off-diagonal entry of , without loss of generality the -entry, is zero, so Since the rows of are not combinatorially orthogonal, has a column with nonzero entries in the first and third rows. Then implies ; similarly . So suppose that is a nonzero hollow matrix. Then by Lemma 3.7, and the preceding argument, no column of has exactly one zero entry. This, along with the fact that no pair of rows of are combinatorially orthogonal, implies has a nowhere zero column . Observe that implies . This is impossible since . Thus, and has the SIPP.
Corollary 3.27.
Suppose is an sign pattern that allows row orthogonality such that the zero entries of are contained in at most three rows of and that no pair of rows are combinatorially orthogonal. Then, requires o-SIPP.
As the next example illustrates, Corollary 3.27 cannot be extended to 4 or more rows.
Example 3.28.
ex:n-zero_n-rowDefine the matrix
Observe that the rows of are orthogonal. It can be verified that
satisfies . Since and , it follows that does not have the SIPP.
Even if we also prohibit combinatorially orthogonal columns, there are examples of sign patterns with the zero entries restricted to the first four rows that do not require o-SIPP, as seen in the next example, which utilizes a construction method in [5].
Example 3.29.
Consider
It is readily verified that the rows of are orthogonal and . Thus, allows orthogonality but does not require o-SIPP.
By restricting the number of zero entries, we obtain the following result.
Proposition 3.30.
Let have at most four zero entries. Suppose that no pair of rows and no pair of columns columns of are combinatorially orthogonal. Then has the SIPP.
Proof 3.31.
Observe that if the zero entries of are contained in at most three rows, then Theorem 3.25 implies has the SIPP. So, suppose that each zero entry of is contained in a unique row.
Assume first that . If has a nowhere zero column, then Proposition 1.5 and Theorem 3.25 imply has the SIPP. Otherwise, without loss of generality, is a nonzero hollow matrix. Theorem 3.2 in [2] guarantees that is not symmetric. Thus, Corollary 3.21 implies has the SIPP.
Now assume that (and ). Without loss of generality, has the form
where all four zero entries of are contained in such that each column and each row of contains a zero entry. Note that has four rows, is nowhere zero, and and may be vacuous. We proceed via cases.
We first resolve the case where has exactly one column, i.e., . Then is nowhere zero. Thus has the SIPP and, by Lemma 3.9, has the SIPP.
Suppose that has at least two columns. Let and suppose . By Lemma 3.3, , where is symmetric and has four rows. Further, Lemma 3.5 implies , i.e., has the form
We now consider the case where has a column with exactly one zero entry. Without loss of generality, the zero appears in the last entry of . Then implies . It now follows from that . Thus, and has the SIPP.
For the final case, assume that has exactly two columns that contain exactly two zero entries each. Since the first two columns cannot be combinatorialy orthogonal, there must be at least five rows. Then, without loss of generality,
From it follows that . If for , then implies . Suppose this is not the case, i.e., for all . Then implies , and thus for some nonzero value . Since is row orthogonal
From the last two equations . Substituting this into the first equation implies , a contradiction. Thus, and has the SIPP.
It is possible to show that if has at most five zero entries, no pair of rows and no pair of columns columns of are combinatorially orthogonal, then has the SIPP. However, with the available tools, the argument is not illuminating and does not warrant the space that would be required. As the next example illustrates, it is possible for a sign pattern with six zero entries to allow row orthogonality, not have combinatorially orthogonal rows or columns, and not require o-SIPP.
Example 3.32.
Consider the sign pattern
Observe that is a conference matrix, i.e., is hollow, every off-diagonal entry is or , and . Hence allows orthogonality. By Corollary 3.21, does not require o-SIPP. In fact, it is not difficult to see that does not have the SIPP since the symmetric matrix satisfies .
3.3 Nowhere zero sign patterns that minimally allow orthogonality
In this section we determine the nowhere zero sign patterns with at most five rows that minimally allow orthogonality. These and previously known results are summarized in Table 3.1 at the end of this section, which lists a representative of each equivalence class of nowhere zero sign patterns that minimally allow orthogonality for . Recall that a sign pattern minimally allows orthogonality provided allows row orthogonality and every sign pattern obtained from by deleting one or more columns does not allow row orthogonality.
A complete characterization of nowhere zero sign patterns with at most 4 rows that minimally allow orthogonality was presented in [5]. We summarize these results in the following theorem.
Theorem 3.33.
[5, Section 5.2] Let be an nowhere zero sign pattern. If , then minimally allows orthogonality if and only if and is row and column PPO. If , then minimally allows orthogonality if and only if and is row and column PPO, or is sign equivalent to
We now determine all nowhere zero sign patterns that minimally allow orthogonality. The next theorem establishes the square case.
Theorem 3.34.
[5, Theorem 7.9] Let be a nowhere zero sign pattern. Then allows orthogonality if and only if is row and column PPO.
Lemma 3.35.
Let be a nowhere zero sign pattern. Then is sign equivalent to a sign pattern with at most 5 negative entries.
Proof 3.36.
By scaling the rows and columns of we can obtain the sign patterns
If contains at most 5 negative entries, then the proof is complete. So, suppose that has at least 6 negative entries.
First consider the case where has exactly 6 negative entries. If a row (or column) of has 3 negatives, then negating the corresponding row (or column) of reduces the total number of negative entries to at most 5. Otherwise has two rows and , each containing exactly 2 negative entries, and a third row that contains a negative entry; let denote the column index of this entry. Observe that negating the rows of corresponding to and does not change the total number of negative entries. Thus, we can scale the rows of so that column has 3 negative entries. Negating column now reduces the total number of negatives to at most 5.
Now suppose that has at least 7 negative entries. Then has at most 6 negative entries. As before, we can reduce the total number of negative entries to at most 5 if a row (or column) of contains at least 3 negative entries, or if has two rows that each contain exactly 2 negative entries. Thus, we may assume has 1 row with exactly 2 negative entries and 2 columns with exactly 2 negative entries. Without loss of generality is of the form
Negate columns 2 and 3, and rows 1 and 3 in the first case. Negate row 2 and column 4 in the second case.
The next example uses the ideas illustrated in Example 2.9.
Example 3.37.
We can apply Theorem 2.7 to the matrices
to obtain row orthogonal matrices with the same sign patterns: For , the value is obtained from row of and the value is obtained from rows and . Thus . Since is increasing on its domain, . For , the value is obtained from row of and the value is obtained from rows and . Thus and .
Theorem 3.38.
Let be a nowhere zero sign pattern. Then minimally allows orthogonality if and only if and is row and column PPO, or is sign equivalent to
Proof 3.39.
Observe that each of the three patterns , and allows row orthogonality by Examples 2.9 and 3.37. Removing a column from one of or results in a sign pattern with a duplicate column, and such a sign pattern is not column PPO. So by Theorem 3.34, removing a column from one of and results in a sign pattern that does not allow orthogonality. Thus, each of and minimally allows row orthogonality.
Assume that minimally allows orthogonality. Without loss of generality the first row and first column of have all positive entries. Suppose that has distinct columns . It is easy to see that : If had at most 3 distinct columns, then would have at most distinct rows, contradicting the fact that is row PPO.
First consider the case where has distinct columns. Since is row PPO,
is row PPO. Observe that is column PPO since are distinct. By Theorem 3.34, allows orthogonality. Since minimally allows orthogonality, .
Now suppose that and let
As before, if is row PPO, then allows orthogonality. Since minimally allows orthogonality, it follows from the preceding argument that is not row PPO. Without loss of generality
where the columns of are distinct and belong to the set
Either contains a column with exactly one negative entry or every column of has at least two negative entries. Observe that in the latter case, negating the last three rows of results in a column with all positive entries and a column with exactly one negative entry. Thus, we may assume
Since allows row orthogonality, for some . Observe that
is row and column PPO and hence allows orthogonality by Theorem 3.34. This is a contradiction since minimally allows orthogonality. Thus, .
Finally, we consider the case where has exactly distinct columns. Let . By Lemma 3.35 we may assume that has at most negative entries. Observe that at least rows of contain a negative entry since is row PPO.
Suppose has exactly negative entries. Then is sign equivalent to
and we assume has this form. Observe that can be obtained from by duplicating some of the columns. By Theorem 1.6 we must duplicate at least 2 distinct columns of to obtain . It follows that, up to sign equivalence, contains the submatrix
Suppose that has negatives. Observe that cannot have exactly 1 negative per row, as this would contradict being row PPO. Further, at most 1 row of has 2 negatives, otherwise we have 2 positive rows, which violates row PPO. By these considerations, is sign equivalent to
and we assume has this form. As before, can be obtained from by duplicating at least 2 distinct columns of . Observe that duplicating only columns 1 and 2 of violates Theorem 1.6. Duplicating columns 1 and 3 is sign equivalent to duplicating columns 1 and 4. Duplicating columns 2 and 3 is sign equivalent to duplicating columns 2 and 4. Thus, up to sign equivalence, contains one of
or
as a submatrix. Observe that is sign equivalent to (negate rows and , and negate columns and ; then appropriately permute rows and columns). By Examples 2.9 and 3.37, and allow row orthogonality. Since minimally allows orthogonality, is sign equivalent to or .
Remark 3.40.
Characterizing all sign patterns that minimally allow orthogonality may require a new approach. However, in doing so, we may learn a great deal about sign patterns that allow row orthogonality. Consider the sign pattern
Deleting any number of columns will contradict Theorem 1.6, so if allows row orthogonality, then it minimally allows orthogonality. Using the techniques described in Example 2.9, we were unable to find a row orthogonal realization of . It would be very interesting if this sign pattern does not allow row orthogonality. It is not too difficult to verify that satisfies the conditions of Theorem 1.6, so this would unveil a new necessary condition for sign patterns to allow row orthogonality.
Rows | Unique sign patterns (up to sign equivalence) |
---|---|
1 | |
2 | |
3 | |
4 | , , , |
, , , | |
5 | , , , |
, , , | |
, , |
4 Likelihood a random sign pattern allows row orthogonality
The question of finding the probability that vectors sampled from are linearly independent has attracted recent attention in the literature. This problem can equivalently be stated as asking for the probability that a random matrix in has rank (the literature is most interested in the case when ). In particular, Tikhomirov [12] answered this question in a strong form by showing that this probability is bounded below by whenever ; in particular, when the probability tends toward as tends toward .
In this section, we consider the adjacent problem of determining the threshold such that a random matrix in with allows row orthogonality with probability tending toward as tends toward .
Let and be functions from the non-negative integers to the reals. Then if , and if . An event happens with high probability as if . The union bound is the fact that the probability that at least one of a set of the events happens is at most the sum of the probabilities of the events.
For a probability distribution on a set , we write to mean that is distributed according to . If is a finite set, then we write to mean that is chosen uniformly from . We write to indicate that are distributed according to and are mutually independent from one another. For a positive integer , denotes the product distribution on where each entry is drawn independently from . Similarly, for positive integers and , denotes the product distribution on where each entry is drawn independently from . For an index set , let denote the set of vectors with entries in indexed by . We write to mean the product distribution on where each entry is drawn independently from .
We will need two forms of the Chernoff bound, which we state here.
Theorem 4.1.
Theorem 4.2.
[11, Remark 9.2] Suppose are independent random variables taking values from the set . Let . Then for any
For a fixed , let denote the distribution on where and . The main result of this section, Theorem 4.12, implies that for any fixed , there is a constant such that if where , then allows row orthogonality with high probability as . Before proving Theorem 4.12, we use Theorem 2.7 to obtain a slightly weaker result for nowhere zero sign patterns. We include this result since the proof is relatively short and highlights a substantially different approach from Theorem 4.12.
Theorem 4.3.
If with , then allows row orthogonality with high probability as .
Proof 4.4.
Let denote the th row of , so . Observe that and that for each . Set
and observe that since and so
Thus, if for all , then we may apply Theorem 2.7 to locate a set of orthogonal vectors such that . Thus, in order to conclude the proof, it suffices to show that for all with high probability as .
Since are independent, we may apply the Chernoff bound in Theorem 4.1 to bound
for any . By the union bound,
We now show how to improve Theorem 4.3 by using the SIPP. Recall that Theorem 3.22 states that the sign pattern requires o-SIPP. We say that a pair of negative -cycles are column-disjoint if the column indices of the negative -cycles are all distinct. Observe that any sign pattern that has a collection of column-disjoint negative -cycles between every pair of rows is sign equivalent to a superpattern of . So by Theorem 1.1 and Theorem 1.3, we have the following observation.
Observation 4.5.
Let be an sign pattern. If has a collection of column-disjoint negative 4-cycles between every pair of rows, then allows row orthogonality.
In the following proofs, we must condition on the outcome of a stochastic process. For those readers unfamiliar with these ideas, we recommend consulting [13, Chapter 9].
Lemma 4.6.
Fix any . If , then the probability that we can find distinct integers such that and for all is at least
Proof 4.7.
We employ the following greedy algorithm to find the required set of indices:
Initialize and . At time , do the following:
-
1.
Reveal .
-
2.
Attempt to locate some for which and . If such are found, then set and . If no such exist, then exit with failure.
If the above algorithm succeeds, then we have located the desired .
Let be the round on which the algorithm fails, setting if the algorithm succeeds. In order to complete the proof, we show that
Fix any and consider conditioning on the event . Since if and only if the algorithm has succeeded locating the set , we may condition on such an outcome. Now, conditioned on the algorithm locating , we observe that if and only if is nonnegative or nonpositive. Furthermore, before the th loop, no information is known about the vector and so . We may therefore bound
Since this bound is independent of , we may bound
We therefore conclude that
where the final inequality follows from the fact that .
We now use the above lemma to locate collections of column-disjoint negative -cycles.
Lemma 4.8.
Fix any and any . Assume and set Then contains column-disjoint negative -cycles between its first row and all other rows with probability at least
Proof 4.9.
Let be the diagonal matrix whose th diagonal entry is the th entry of . Observe that is sign equivalent to . Since and , . Thus contains column-disjoint negative -cycles between its first row and all other rows if and only if we can locate distinct so that and . As such, the conclusion follows from Lemma 4.6.
Lemma 4.10.
Fix a number . Assume , where for some . Then the probability that contains a collection of column-disjoint negative -cycles between every pair of rows is bounded below by
Proof 4.11.
In order to locate a collection of negative -cycles between every pair of rows of , we employ the following greedy algorithm:
Suppose that the rows of are . Initialize . At time do the following:
-
1.
Reveal .
-
2.
Find some with and set . If no such exists, then fail.
-
3.
Reveal
-
4.
Locate column-disjoint negative -cycles in between row and all rows , all of whose columns reside within . If such negative -cycles cannot be found, then fail.
If the above algorithm succeeds, then contains a collection of column-disjoint negative -cycles between every pair of rows.
Let denote the first round on which the algorithm fails, setting if the algorithm succeeds. In order to prove the claim, we argue that
Fix any . Let denote the event that the algorithm fails at step 2 on the th loop, and let denote the event that the algorithm fails at step 4 on the th loop. Certainly . We begin by bounding the probability of .
Consider conditioning on the event . Of course, if , then the algorithm has succeeded in locating the set . Furthermore, conditioned on and , observe that prior to the th loop of the algorithm, no entries within have been revealed; therefore . In particular, . Now, cannot be located if and only if . Additionally,
We can therefore fix a subset with . From the earlier observation, we know that and so we may bound
where the final inequality follows from the fact that and .
Next, we note that if , then and so
where the second inequality follows from the Chernoff bound in Theorem 4.2. Since this bound is independent of , we have argued that
(4.4) |
Next, we bound the probability of . In order for to hold, it must be the case that and that does not hold; in particular, the algorithm must have succeeded in locating the set . By construction, just after locating , no entries within have been revealed; therefore . Since is equal to the probability of not finding a collection of column disjoint negative 4-cycles between the first row of and the remaining rows, we may appeal to Lemma 4.8 to bound
Since this bound is independent of , we have shown that
(4.5) |
where denotes the event that does not occur.
Using this inequality, we finally bound
as needed.
Theorem 4.12.
For any fixed , if and
then allows row orthogonality with high probability as .
Proof 4.13.
Suppose that
where . Without loss of generality, we may additionally suppose that . Set which is certainly bounded above by for all sufficiently large since . Furthermore, by decreasing by some amount no more than , we may ensure that is an integer, the lower bound on remains true, , and . Now, since and is fixed, we have that
for all sufficiently large . Thus, we may apply Lemma 4.10 to learn that contains a collection of column-disjoint negative -cycles between every pair of rows (and hence has a row orthogonal realization) with probability at least
which tends to as since and .
We suspect that Theorem 4.12 is not best possible.
Question 4.14.
Determine the threshold such that if with , then has a row orthogonal realization with high probability as .
Theorem 4.12 implies that . Observe that and it is possible that this is the correct answer. As shown in the next theorem, the best known obstruction (see Theorem 1.6) does not block .
Theorem 4.15.
Let . Then with high probability as the matrix does not contain an submatrix such that and .
Proof 4.16.
Let denote the set of pairs , where , and the first entry of is . Observe that the map is a bijection between and the set of rank 1 matrices in . Thus the probability that has rank 1 is precisely .
The number of submatrices of is . By the union bound, the probability that contains an submatrix such that and is at most
We show that this sum tends toward as by showing that
for all , provided is sufficiently large. If or , then for sufficiently large
For , we have
Acknowledgements
The research of Z. Brennan, C. Cox, B. Curtis, E. Gomez-Leos, and C. Thompson was partially supported by NSF grant 1839918 and the authors thank the National Science Foundation.
References
- [1] N. Alon and J.H. Spencer. The probabilistic method. John Wiley & Sons, 2016.
- [2] R.F. Bailey and R. Craigen. On orthogonal matrices with zero diagonal. Electron. J. Linear Algebra, 35:307–318, 2019.
- [3] J. Bourgain, V.H. Vu and P.M. Wood. On the singularity probability of discrete random matrices, J.Functional Analysis 258 (2010), no. 2, 559–603. MR2557947
- [4] H. Cohn, A. Kumar and G. Minton. Optimal simplices and codes in projective spaces. Geometry & Topology, 20:1289–1357, 2016.
- [5] B.A. Curtis. Sign Patterns of Row Orthogonal Matrices. PhD thesis, 2020.
- [6] B.A. Curtis and B.L. Shader. Sign patterns of orthogonal matrices and the strong inner product property. Linear Algebra and its Applications, 592: 228–259, 2020.
- [7] M. Fiedler, “Problem 12,” in Proceedings: Theory of Graphs and Its Application. Prague: Publishing House of the Czechoslovakia Academy of Sciences, 1964, p. 160.
- [8] L. Hogben, J.C.-H. Lin, B.L. Shader. Inverse Problems and Zero Forcing for Graphs. American Mathematical Society (Mathematical Surveys and Monographs, 260), Providence, RI, 2022.
- [9] H. van der Holst, L. Lovász, and A. Schrijver. The Colin de Verdière graph parameter. In Graph Theory and Combinatorial Biology (L. Lovász, A. Gyárfás, G. Katona, A. Recski, and L. Székely, editors), János Bolyai Mathematical Society, Budapest, 1999, pp. 29–85.
- [10] C.R. Johnson and C. Waters. Sign patterns occurring in orthogonal matrices. Linear and Multilinear Algebra, 44: 287–299, 1998.
- [11] H. Tijms. Probability: A lively introduction. Cambridge University Press, 2018.
- [12] K. Tikhomirov. Singularity of random Bernoulli matrices. Ann. of Math. (2) 191 (2) 593 - 634, 2018.
- [13] D. Williams. Probability with martingales. Cambridge University Press, 1991.