This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Orthogonal realizations of random sign patterns and other applications of the SIPP

Zachary Brennan Department of Mathematics, Iowa State University (brennanz, cocox, bcurtis1, enriqueg, kph3, hogben, conorjt)@iastate.edu    Christopher Cox22footnotemark: 2    Bryan A. Curtis22footnotemark: 2 Corresponding author    Enrique Gomez-Leos22footnotemark: 2    Kimberly P. Hadaway22footnotemark: 2    Leslie Hogben22footnotemark: 2 American Institute of Mathematics, San Jose, CA 95112, USA ([email protected]).    Conor Thompson22footnotemark: 2
Abstract

A sign pattern is an array with entries in {+,,0}\{+,-,0\}. A matrix QQ is row orthogonal if QQT=IQQ^{T}=I. 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, 5×n5\times n 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.
{AMS}

15B10, 15B35, 15B52, 60B20.

1 Introduction

A sign pattern is an array with entries coming from the set {+,,0}\{+,-,0\}. 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 QQ is row orthogonal provided QQT=IQQ^{T}=I. The problem of determining whether an m×nm\times n 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 5×n5\times n 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 h(m)h(m) such that the probability of a random m×nm\times n sign pattern allowing row orthogonality goes to 11 as mm tends toward \infty for nh(m)n\geq h(m) (here random means ++ and - are equally likely and the probability of 0 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 {,0}\{*,0\}. We use the term pattern to mean a sign pattern or a zero-nonzero pattern. Given a real number aa,

sgn(a)={+if a>0if a<00if a=0andznz(a)={if a00if a=0.\operatorname{sgn}(a)=\begin{cases}+&\text{if }a>0\\ -&\text{if }a<0\\ 0&\text{if }a=0\end{cases}\qquad\text{and}\qquad\operatorname{znz}(a)=\begin{cases}*&\text{if }a\not=0\\ 0&\text{if }a=0.\end{cases}

The sign pattern and zero-nonzero pattern of a matrix A=[aij]A=[a_{ij}] are sgn(A)=[sgn(aij)]\operatorname{sgn}(A)=[\operatorname{sgn}(a_{ij})] and znz(A)=[znz(aij)]\operatorname{znz}(A)=[\operatorname{znz}(a_{ij})], respectively. The qualitative class of an m×nm\times n sign pattern SS is the set

𝒬(S)={Am×n:sgn(A)=S},\mathcal{Q}(S)=\{A\in\mathbb{R}^{m\times n}:\operatorname{sgn}(A)=S\},

and the qualitative class of an m×nm\times n znz pattern Z=[zij]Z=[z_{ij}] is the set

𝒬(Z)={Am×n:znz(A)=Z}.\mathcal{Q}(Z)=\{A\in\mathbb{R}^{m\times n}:\operatorname{znz}(A)=Z\}.

A matrix in the qualitative class 𝒬(P)\mathcal{Q}(P) is called a realization of the pattern PP. For a sign pattern SS, CSC_{S} denotes the unique (1,1,0)(1,-1,0)-matrix that is a realization of the sign pattern SS. Similarly, CZC_{Z} is the unique (1,0)(1,0)-matrix that is a realization of the zero-nonzero pattern ZZ. A superpattern of a sign pattern S=[sij]S=[s_{ij}] is a sign pattern R=[rij]R=[r_{ij}] of the same dimensions such that rij=sijr_{ij}=s_{ij} whenever sij{+,}s_{ij}\in\{+,-\}; if sij=0s_{ij}=0 then rij{+,,0}r_{ij}\in\{+,-,0\}.

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 m×nm\times n row orthogonal matrices is denoted by 𝒪(m,n)\mathcal{O}(m,n) and we write 𝒪(m)\mathcal{O}(m) as shorthand for 𝒪(m,m)\mathcal{O}(m,m). Note that every matrix Q𝒪(m)Q\in\mathcal{O}(m) is orthogonal, i.e., QTQ=QQT=IQ^{T}Q=QQ^{T}=I. The set of m×mm\times m real symmetric matrices is denoted by sym(m)\operatorname{sym}(m).

A zero matrix Om×nO\in\mathbb{R}^{m\times n} or zero vector 𝟎n{\bf 0}\in\mathbb{R}^{n} has every entry equal to zero. An m×nm\times n matrix or pattern is wide if mnm\leq n. 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 Am×nA\in\mathbb{R}^{m\times n} be a wide matrix. Then AA has the strong inner product property (SIPP) provided X=OX=O is the only symmetric matrix satisfying (XA)A=O(XA)\circ A=O [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 m×nm\times n sign pattern SS allows row orthogonality if there is a row orthogonal matrix Q𝒬(S)Q\in\mathcal{Q}(S) (equivalently, 𝒪(m,n)𝒬(S)\mathcal{O}(m,n)\cap\mathcal{Q}(S)\neq\emptyset). An m×nm\times n sign pattern SS allows o-SIPP if there is a row orthogonal matrix Q𝒬(S)Q\in\mathcal{Q}(S) 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 SS requires o-SIPP if every Q𝒬(S)𝒪(m,n)Q\in\mathcal{Q}(S)\cap\mathcal{O}(m,n) has the SIPP and 𝒬(S)𝒪(m,n)\mathcal{Q}(S)\cap\mathcal{O}(m,n)\neq\emptyset. Without the assumption that 𝒬(S)𝒪(m,n)\mathcal{Q}(S)\cap\mathcal{O}(m,n)\neq\emptyset, the all zeros pattern would vacuously require o-SIPP.

An m×nm\times n rectangular sign pattern SS is row potentially pairwise-orthogonal or row PPO if no row is a zero vector and for each pair (i,k)(i,k) with 1i<km1\leq i<k\leq m, there are realizations of row ii and row kk that are orthogonal. The term column PPO is defined analogously. A pair of rows ii and kk in a m×nm\times n rectangular sign pattern S=[sij]S=[s_{ij}] has a negative 44-cycle if there are two columns jj and \ell such that sijskl=+s_{ij}s_{kl}=+ and sijskl=s_{ij}s_{kl}=-, where multiplication on the set {0,+,}\{0,+,-\} is defined in the obvious way that conforms to real arithmetic. A pair of rows ii and kk in an m×nm\times n matrix or pattern P=[pij]P=[p_{ij}] is combinatorially orthogonal if pij0p_{ij}\neq 0 implies pkj=0p_{kj}=0 for every j=1,,nj=1,\dots,n.

A signed permutation matrix is a square (1,1,0)(1,-1,0)-matrix with exactly one nonzero entry in each row and column. Matrices A,Bm×nA,B\in\mathbb{R}^{m\times n} are sign equivalent if A=P1BP2A=P_{1}BP_{2}, where P1P_{1} and P2P_{2} are signed permutation matrices. Two sign patterns SS and SS^{\prime} are sign equivalent if CSC_{S} and CSC_{S^{\prime}} are sign equivalent.

For a vector 𝐯n{\bf v}\in\mathbb{R}^{n}, the support of 𝐯{\bf v}, denoted by supp(𝐯)\operatorname*{supp}({\bf v}), is the set of indices of nonzero entries of 𝐯{\bf v}. Let [n]={1,,n}[n]=\{1,\dots,n\}.

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 SS 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 Q𝒪(m,n)Q\in\mathcal{O}(m,n) has the SIPP and sgn(Q)=S\operatorname{sgn}(Q)=S, then every superpattern of SS 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 Am×nA\in\mathbb{R}^{m\times n} and p>np>n. Then AA has the SIPP if and only if the m×pm\times p matrix B=[A|O]B=\big{[}A\leavevmode\nobreak\ |\leavevmode\nobreak\ O\big{]} has the SIPP.

Corollary 1.3.

[6] If Q𝒪(m,n)Q\in\mathcal{O}(m,n) has the SIPP and sgn(Q)=S\operatorname{sgn}(Q)=S, then [S|O]\begin{bmatrix}{S\leavevmode\nobreak\ \big{|}\leavevmode\nobreak\ O}\end{bmatrix} 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 A,Bm×nA,B\in\mathbb{R}^{m\times n} be sign equivalent. Then AA has the SIPP if and only if BB has the SIPP.

Proposition 1.5.

[6] Let Q𝒪(m)Q\in\mathcal{O}(m). Then QQ has the SIPP if and only if QTQ^{T} 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 SS be a nowhere zero sign pattern and let RR be an r×sr\times s submatrix of SS. If r+sn+2r+s\geq n+2 and rankCR=1\operatorname{rank}C_{R}=1, then SS 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 rm(ϵ)r_{m}(\epsilon) used to quantify the notion of “nearby.”

Definition 2.1.

For an integer m1m\geq 1 and a real number 0ϵ<1m10\leq\epsilon<\frac{1}{m-1},

rm(ϵ)=1+ϵ(1(m2)ϵ)(1(m1)ϵ)1r_{m}(\epsilon)=\sqrt{{1+\epsilon\over(1-(m-2)\epsilon)(1-(m-1)\epsilon)}}-1

where 10\frac{1}{0} is interpreted as \infty, so r1(ϵ)=0r_{1}(\epsilon)=0 for all ϵ0\epsilon\geq 0.

For simplicity, we have defined the functions rm(ϵ)r_{m}(\epsilon) 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 r1(ϵ)=0r_{1}(\epsilon)=0, rm(ϵ)r_{m}(\epsilon) can be computed recursively for all m2m\geq 2 and all 0ϵ<1m10\leq\epsilon<\frac{1}{m-1} by

rm(ϵ)=1+ϵ1ϵ(rm1(ϵ1ϵ)+1)1.r_{m}(\epsilon)=\sqrt{\frac{1+\epsilon}{1-\epsilon}}\biggl{(}r_{m-1}\biggl{(}\frac{\epsilon}{1-\epsilon}\biggr{)}+1\biggr{)}-1.

Proof 2.3.

It is easy to verify the result for m=2m=2. For m3m\geq 3,

rm(ϵ)\displaystyle r_{m}(\epsilon) =1+ϵ(1(m2)ϵ)(1(m1)ϵ)1.\displaystyle=\sqrt{{1+\epsilon\over(1-(m-2)\epsilon)(1-(m-1)\epsilon)}}-1.
=1+ϵ1ϵ(1ϵ)((1ϵ)+ϵ)((1ϵ)(m3)ϵ)((1ϵ)(m2)ϵ)1\displaystyle=\sqrt{{1+\epsilon\over 1-\epsilon}}\sqrt{{(1-\epsilon)((1-\epsilon)+\epsilon)\over\bigl{(}(1-\epsilon)-(m-3)\epsilon\bigr{)}\bigl{(}(1-\epsilon)-(m-2)\epsilon\bigr{)}}}-1
=1+ϵ1ϵ1+ϵ1ϵ(1(m3)ϵ1ϵ)(1(m2)ϵ1ϵ)1\displaystyle=\sqrt{{1+\epsilon\over 1-\epsilon}}\sqrt{{1+{\epsilon\over 1-\epsilon}\over\bigl{(}1-(m-3){\epsilon\over 1-\epsilon}\bigr{)}\bigl{(}1-(m-2){\epsilon\over 1-\epsilon}\bigr{)}}}-1
=1+ϵ1ϵ(rm1(ϵ1ϵ)+1)1.\displaystyle=\sqrt{{1+\epsilon\over 1-\epsilon}}\biggl{(}r_{m-1}\biggl{(}{\epsilon\over 1-\epsilon}\biggr{)}+1\biggr{)}-1.

The next lemma provides the key step to go from approximately to exactly orthogonal.

Lemma 2.4.

Let mm be a positive integer, let 0ϵ<1m10\leq\epsilon<{1\over m-1} and fix any inner product space (Ω,,)(\Omega,\langle\cdot,\cdot\rangle). Additionally, let \lVert\cdot\rVert be any norm on Ω\Omega (possibly unrelated to ,\langle\cdot,\cdot\rangle). If 𝐱1,,𝐱mΩ{\bf x}_{1},\dots,{\bf x}_{m}\in\Omega satisfy

  1. 1.

    𝐱i,𝐱i=1\langle{\bf x}_{i},{\bf x}_{i}\rangle=1 for all i[m]i\in[m], and

  2. 2.

    |𝐱i,𝐱j|ϵ\lvert\langle{\bf x}_{i},{\bf x}_{j}\rangle\rvert\leq\epsilon for all ij[m]i\neq j\in[m],

then there exists 𝐱~1,,𝐱~mspan{𝐱1,,𝐱m}\widetilde{{\bf x}}_{1},\dots,\widetilde{{\bf x}}_{m}\in\operatorname*{span}\{{\bf x}_{1},\dots,{\bf x}_{m}\} satisfying

  1. 1.

    {𝐱~1,,𝐱~m}\{\widetilde{{\bf x}}_{1},\dots,\widetilde{{\bf x}}_{m}\} is orthonormal with respect to ,\langle\cdot,\cdot\rangle, and

  2. 2.

    𝐱i𝐱~irm(ϵ)𝐱i\lVert{\bf x}_{i}-\widetilde{{\bf x}}_{i}\rVert\leq r_{m}(\epsilon)\lVert{\bf x}_{i}\rVert for all i[m]i\in[m].

Proof 2.5.

We prove the result by induction on mm. The case m=1m=1 is immediate by taking 𝐱~1=𝐱1\widetilde{{\bf x}}_{1}={\bf x}_{1}. Let m2m\geq 2 be a positive integer and suppose the statement holds for m1m-1.

Without loss of generality, 𝐱m𝐱i\lVert{\bf x}_{m}\rVert\leq\lVert{\bf x}_{i}\rVert for all i{1,,m}i\in\{1,\dots,m\}. For i{1,,m1}i\in\{1,\ldots,m-1\}, let 𝐱i  =𝐱i,𝐱m𝐱m{\bf x}_{i}^{\mkern 3.0mu\vphantom{\perp}\vrule depth=0.0pt\mkern 4.0mu\vrule depth=0.0pt\mkern 3.0mu}=\langle{\bf x}_{i},{\bf x}_{m}\rangle{\bf x}_{m} and 𝐱i=𝐱i𝐱i  {\bf x}_{i}^{\perp}={\bf x}_{i}-{\bf x}_{i}^{\mkern 3.0mu\vphantom{\perp}\vrule depth=0.0pt\mkern 4.0mu\vrule depth=0.0pt\mkern 3.0mu}; then 𝐱i,𝐱m=0\langle{\bf x}_{i}^{\perp},{\bf x}_{m}\rangle=0. Since |𝐱i,𝐱m|ϵ\lvert\langle{\bf x}_{i},{\bf x}_{m}\rangle\rvert\leq\epsilon and 𝐱m,𝐱m=1\langle{\bf x}_{m},{\bf x}_{m}\rangle=1, we know that 0𝐱i  ,𝐱i  ϵ20\leq\langle{\bf x}_{i}^{\mkern 3.0mu\vphantom{\perp}\vrule depth=0.0pt\mkern 4.0mu\vrule depth=0.0pt\mkern 3.0mu},{\bf x}_{i}^{\mkern 3.0mu\vphantom{\perp}\vrule depth=0.0pt\mkern 4.0mu\vrule depth=0.0pt\mkern 3.0mu}\rangle\leq\epsilon^{2}. Therefore, the Pythagorean Theorem allows us to conclude that

(2.1) 𝐱i,𝐱i=𝐱i  ,𝐱i  +𝐱i,𝐱i1𝐱i,𝐱i1ϵ2.\langle{\bf x}_{i},{\bf x}_{i}\rangle=\langle{\bf x}_{i}^{\mkern 3.0mu\vphantom{\perp}\vrule depth=0.0pt\mkern 4.0mu\vrule depth=0.0pt\mkern 3.0mu},{\bf x}_{i}^{\mkern 3.0mu\vphantom{\perp}\vrule depth=0.0pt\mkern 4.0mu\vrule depth=0.0pt\mkern 3.0mu}\rangle+\langle{\bf x}_{i}^{\perp},{\bf x}_{i}^{\perp}\rangle\quad\implies\quad 1\geq\langle{\bf x}_{i}^{\perp},{\bf x}_{i}^{\perp}\rangle\geq 1-\epsilon^{2}.

Since ϵ<1\epsilon<1, we know that 𝐱i{\bf x}_{i}^{\perp} is non-zero. Let 𝐱^i\widehat{{\bf x}}_{i}^{\perp} denote the unit vector in the direction of 𝐱i{\bf x}_{i}^{\perp}, i.e., 𝐱^i=1𝐱i,𝐱i𝐱i\widehat{{\bf x}}_{i}^{\perp}=\frac{1}{\sqrt{\langle{\bf x}_{i}^{\perp},{\bf x}_{i}^{\perp}\rangle}}{\bf x}_{i}^{\perp}.

Since 𝐱i  =𝐱i,𝐱m𝐱m{\bf x}_{i}^{\mkern 3.0mu\vphantom{\perp}\vrule depth=0.0pt\mkern 4.0mu\vrule depth=0.0pt\mkern 3.0mu}=\langle{\bf x}_{i},{\bf x}_{m}\rangle{\bf x}_{m}, we see that 𝐱i  ϵ𝐱mϵ𝐱i\lVert{\bf x}_{i}^{\mkern 3.0mu\vphantom{\perp}\vrule depth=0.0pt\mkern 4.0mu\vrule depth=0.0pt\mkern 3.0mu}\rVert\leq\epsilon\lVert{\bf x}_{m}\rVert\leq\epsilon\lVert{\bf x}_{i}\rVert. Then the triangle inequality applied to 𝐱i=𝐱i+(𝐱i  ){\bf x}_{i}^{\perp}={\bf x}_{i}+(-{\bf x}_{i}^{\mkern 3.0mu\vphantom{\perp}\vrule depth=0.0pt\mkern 4.0mu\vrule depth=0.0pt\mkern 3.0mu}) implies that 𝐱i(1+ϵ)𝐱i\lVert{\bf x}_{i}^{\perp}\rVert\leq(1+\epsilon)\lVert{\bf x}_{i}\rVert. Together with (2.1), we have

(2.2) 𝐱^i1+ϵ1ϵ2𝐱i=1+ϵ1ϵ𝐱i.\lVert\widehat{{\bf x}}_{i}^{\perp}\rVert\leq\frac{1+\epsilon}{\sqrt{1-\epsilon^{2}}}\lVert{\bf x}_{i}\rVert=\sqrt{{1+\epsilon\over 1-\epsilon}}\lVert{\bf x}_{i}\rVert.

In particular,

(2.3) 𝐱i𝐱^i\displaystyle\lVert{\bf x}_{i}-\widehat{{\bf x}}_{i}^{\perp}\rVert \displaystyle\leq 𝐱i+𝐱i𝐱^i=𝐱i+(1𝐱i,𝐱i1)𝐱i\displaystyle\lVert{\bf x}_{i}^{\mkern 3.0mu\vphantom{\perp}\vrule depth=0.0pt\mkern 4.0mu\vrule depth=0.0pt\mkern 3.0mu}\rVert+\lVert{\bf x}_{i}^{\perp}-\widehat{{\bf x}}_{i}^{\perp}\rVert=\lVert{\bf x}_{i}^{\mkern 3.0mu\vphantom{\perp}\vrule depth=0.0pt\mkern 4.0mu\vrule depth=0.0pt\mkern 3.0mu}\rVert+\biggl{(}{1\over\sqrt{\langle{\bf x}_{i}^{\perp},{\bf x}_{i}^{\perp}\rangle}}-1\biggr{)}\lVert{\bf x}_{i}^{\perp}\rVert
\displaystyle\leq ϵ𝐱i+(11ϵ21)(1+ϵ)𝐱i=(1+ϵ1ϵ1)𝐱i,\displaystyle\epsilon\lVert{\bf x}_{i}\rVert+\biggl{(}{1\over\sqrt{1-\epsilon^{2}}}-1\biggr{)}(1+\epsilon)\lVert{\bf x}_{i}\rVert=\biggl{(}\sqrt{\frac{1+\epsilon}{1-\epsilon}}-1\biggr{)}\lVert{\bf x}_{i}\rVert,

where the second inequality follows by combining (2.1) and 𝐱i(1+ϵ)𝐱i\lVert{\bf x}_{i}^{\perp}\rVert\leq(1+\epsilon)\lVert{\bf x}_{i}\rVert.

Now, for any other j{1,,m1}j\in\{1,\ldots,m-1\} with jij\neq i, we have

𝐱i,𝐱j\displaystyle\langle{\bf x}_{i}^{\perp},{\bf x}_{j}^{\perp}\rangle =\displaystyle= 𝐱i,𝐱j𝐱i,𝐱m𝐱m,𝐱j\displaystyle\langle{\bf x}_{i},{\bf x}_{j}\rangle-\langle{\bf x}_{i},{\bf x}_{m}\rangle\langle{\bf x}_{m},{\bf x}_{j}\rangle
|𝐱i,𝐱j|\displaystyle\implies\lvert\langle{\bf x}_{i}^{\perp},{\bf x}_{j}^{\perp}\rangle\rvert \displaystyle\leq ϵ+ϵ2\displaystyle\epsilon+\epsilon^{2}
|𝐱^i,𝐱^j|\displaystyle\implies\lvert\langle\widehat{{\bf x}}_{i}^{\perp},\widehat{{\bf x}}_{j}^{\perp}\rangle\rvert \displaystyle\leq ϵ+ϵ2𝐱i,𝐱i𝐱j,𝐱jϵ+ϵ21ϵ2=ϵ1ϵ.\displaystyle{\epsilon+\epsilon^{2}\over\sqrt{\langle{\bf x}_{i}^{\perp},{\bf x}_{i}^{\perp}\rangle\langle{\bf x}_{j}^{\perp},{\bf x}_{j}^{\perp}\rangle}}\leq\frac{\epsilon+\epsilon^{2}}{1-\epsilon^{2}}=\frac{\epsilon}{1-\epsilon}.

Therefore, since 𝐱^1,,𝐱^m1\widehat{{\bf x}}_{1}^{\perp},\dots,\widehat{{\bf x}}_{m-1}^{\perp} are unit vectors by construction, we may apply the induction hypothesis to find an orthonormal set {𝐱~1,,𝐱~m1}span{𝐱^1,,𝐱^m1}\{\widetilde{{\bf x}}_{1},\dots,\widetilde{{\bf x}}_{m-1}\}\subseteq\operatorname*{span}\{\widehat{{\bf x}}_{1}^{\perp},\dots,\widehat{{\bf x}}_{m-1}^{\perp}\} such that

𝐱^i𝐱~irm1(ϵ1ϵ)𝐱^i\lVert\widehat{{\bf x}}_{i}^{\perp}-\widetilde{{\bf x}}_{i}\rVert\leq r_{m-1}\biggl{(}\frac{\epsilon}{1-\epsilon}\biggr{)}\lVert\widehat{{\bf x}}_{i}^{\perp}\rVert

for each i{1,,m1}i\in\{1,\ldots,m-1\}. By invoking additionally (2.2) and (2.3), we bound

𝐱i𝐱~i\displaystyle\lVert{\bf x}_{i}-\widetilde{{\bf x}}_{i}\rVert 𝐱i𝐱^i+𝐱^i𝐱~i(1+ϵ1ϵ1)𝐱i+rm1(ϵ1ϵ)𝐱^i\displaystyle\leq\lVert{\bf x}_{i}-\widehat{{\bf x}}_{i}^{\perp}\rVert+\lVert\widehat{{\bf x}}_{i}^{\perp}-\widetilde{{\bf x}}_{i}\rVert\leq\biggl{(}\sqrt{{1+\epsilon\over 1-\epsilon}}-1\biggr{)}\lVert{\bf x}_{i}\rVert+r_{m-1}\biggl{(}{\epsilon\over 1-\epsilon}\biggr{)}\lVert\widehat{{\bf x}}_{i}^{\perp}\rVert
(1+ϵ1ϵ1+1+ϵ1ϵrm1(ϵ1ϵ))𝐱i=rm(ϵ)𝐱i\displaystyle\leq\biggl{(}\sqrt{{1+\epsilon\over 1-\epsilon}}-1+\sqrt{{1+\epsilon\over 1-\epsilon}}r_{m-1}\biggl{(}{\epsilon\over 1-\epsilon}\biggr{)}\biggr{)}\lVert{\bf x}_{i}\rVert=r_{m}(\epsilon)\lVert{\bf x}_{i}\rVert

for all i{1,,m1}i\in\{1,\ldots,m-1\} where the last equality follows from Lemma 2.2.

Finally, let 𝐱~m=𝐱m\widetilde{{\bf x}}_{m}={\bf x}_{m}. Then 𝐱~1,,𝐱~m\widetilde{{\bf x}}_{1},\dots,\widetilde{{\bf x}}_{m} satisfy the claim, because 𝐱~1,,𝐱~m1span{𝐱^1,,𝐱^m1}span{𝐱1,,𝐱m}\widetilde{{\bf x}}_{1},\dots,\widetilde{{\bf x}}_{m-1}\in\break\operatorname*{span}\{\widehat{{\bf x}}_{1}^{\perp},\dots,\widehat{{\bf x}}_{m-1}^{\perp}\}\subseteq\operatorname*{span}\{{\bf x}_{1},\dots,{\bf x}_{m}\} by construction, and the former subspace is orthogonal to 𝐱m=𝐱~m{\bf x}_{m}=\widetilde{{\bf x}}_{m}.

Observe that the process used to create the vectors 𝐱~i\widetilde{{\bf x}}_{i} 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 \infty-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 5×n5\times n 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 𝐱=[x1,,xn]Tn{\bf x}=[x_{1},\dots,x_{n}]^{T}\in\mathbb{R}^{n}, define

δ(𝐱)=mini[n]|xi|maxj[n]|xj|=mini[n]|xi|𝐱.\delta({\bf x})={\min_{i\in[n]}\lvert x_{i}\rvert\over\max_{j\in[n]}\lvert x_{j}\rvert}={\min_{i\in[n]}\lvert x_{i}\rvert\over\lVert{\bf x}\rVert_{\infty}}.

Theorem 2.7.

Let 𝐱1,,𝐱mn{\bf x}_{1},\dots,{\bf x}_{m}\in\mathbb{R}^{n} be any non-zero vectors and let ϵ=maxij|𝐱i𝐱i2,𝐱j𝐱j2|,\epsilon=\max_{i\neq j}\big{\lvert}\big{\langle}{{\bf x}_{i}\over\lVert{\bf x}_{i}\rVert_{2}},{{\bf x}_{j}\over\lVert{\bf x}_{j}\rVert_{2}}\big{\rangle}\big{\rvert}, where ,\langle\cdot,\cdot\rangle is the standard Euclidean inner product. If

  1. 1.

    ϵ<1m1\epsilon<{1\over m-1}, and

  2. 2.

    rm(ϵ)<mini[m]δ(𝐱i)r_{m}(\epsilon)<\min_{i\in[m]}\delta({\bf x}_{i}),

then there exists an orthogonal set {𝐱~1,,𝐱~m}n\{\widetilde{{\bf x}}_{1},\dots,\widetilde{{\bf x}}_{m}\}\subseteq\mathbb{R}^{n} satisfying sgn(𝐱i)=sgn(𝐱~i)\operatorname{sgn}({\bf x}_{i})=\operatorname{sgn}(\widetilde{{\bf x}}_{i}) for all i[m]i\in[m].

Proof 2.8.

We apply Lemma 2.4 to the vectors 1𝐱i2𝐱i\frac{1}{\|{\bf x}_{i}\|_{2}}{\bf x}_{i}, specializing to the Euclidean inner product and the \infty-norm, to locate an orthonormal set {𝐱~1,,𝐱~m}n\{\widetilde{{\bf x}}_{1}^{\prime},\dots,\widetilde{{\bf x}}_{m}^{\prime}\}\subseteq\mathbb{R}^{n} such that

𝐱i𝐱i2𝐱~irm(ϵ)𝐱i𝐱i2𝐱i𝐱i2𝐱~irm(ϵ)𝐱i\biggl{\lVert}{{\bf x}_{i}\over\lVert{\bf x}_{i}\rVert_{2}}-\widetilde{{\bf x}}_{i}^{\prime}\biggr{\rVert}_{\infty}\leq r_{m}(\epsilon)\biggl{\lVert}{{\bf x}_{i}\over\lVert{\bf x}_{i}\rVert_{2}}\biggr{\rVert}_{\infty}\quad\implies\quad\lVert{\bf x}_{i}-\lVert{\bf x}_{i}\rVert_{2}\cdot\widetilde{{\bf x}}_{i}^{\prime}\rVert_{\infty}\leq r_{m}(\epsilon)\lVert{\bf x}_{i}\rVert_{\infty}

for all i[m]i\in[m]. In particular, setting 𝐱~i=𝐱i2𝐱~i\widetilde{{\bf x}}_{i}=\lVert{\bf x}_{i}\rVert_{2}\cdot\widetilde{{\bf x}}_{i}^{\prime} for each i[m]i\in[m], we know that {𝐱~1,,𝐱~m}\{\widetilde{{\bf x}}_{1},\dots,\widetilde{{\bf x}}_{m}\} is an orthogonal set and that

𝐱i𝐱~irm(ϵ)𝐱i<δ(𝐱i)𝐱i=minj[n]|(𝐱i)j|.\lVert{\bf x}_{i}-\widetilde{{\bf x}}_{i}\rVert_{\infty}\leq r_{m}(\epsilon)\lVert{\bf x}_{i}\rVert_{\infty}<\delta({\bf x}_{i})\lVert{\bf x}_{i}\rVert_{\infty}=\min_{j\in[n]}\lvert({\bf x}_{i})_{j}\rvert.

Since |xy|<|x|sgn(x)=sgn(y)\lvert x-y\rvert<\lvert x\rvert\implies\operatorname{sgn}(x)=\operatorname{sgn}(y) for any x,yx,y\in\mathbb{R}, we conclude that sgn(𝐱i)=sgn(𝐱~i)\operatorname{sgn}({\bf x}_{i})=\operatorname{sgn}(\widetilde{{\bf x}}_{i}) for all i[m]i\in[m].

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

S=[+++++++++++++++++++++++].S=\begin{bmatrix}-&-&-&+&+&+\\ +&+&-&+&+&+\\ +&+&+&-&-&+\\ +&+&+&+&+&-\\ +&+&+&+&+&+\end{bmatrix}.

Explicitly writing down a row orthogonal realization of SS would be difficult since this requires exact arithmetic. Despite this, it is not too difficult for a computer to find realizations of SS that are row orthogonal up to floating-point error. For example, the following matrix is such a realization for SS:

A=[0.07432940.6689650.2229880.3716470.07432940.5946350.1184150.594680.6653820.03600180.1956240.3873440.5118690.06205420.2067740.2590680.6465930.4540760.6659780.03899290.03961910.6810630.06575040.2919120.023190.2646910.6608170.06115890.5413880.442585].A^{\prime}=\begin{bmatrix}-0.0743294&-0.668965&-0.222988&0.371647&0.0743294&0.594635\\ 0.118415&0.59468&-0.665382&0.0360018&0.195624&0.387344\\ 0.511869&0.0620542&0.206774&-0.259068&-0.646593&0.454076\\ 0.665978&0.0389929&0.0396191&0.681063&0.0657504&-0.291912\\ 0.02319&0.264691&0.660817&0.0611589&0.541388&0.442585\\ \end{bmatrix}.

Of course, AA^{\prime} is not actually a row orthogonal matrix and so it does not directly demonstrate that SS has a row orthogonal realization; however, AA^{\prime} does satisfy the hypotheses of Theorem 2.7. In fact, by scaling and truncating AA^{\prime} appropriately, we find the following integer-valued matrix which satisfies the hypotheses of Theorem 2.7 as well:

A=[874254186513657342243567232871507344757323297376049].A=\begin{bmatrix}-8&-74&-25&41&8&65\\ 13&65&-73&4&22&43\\ 56&7&23&-28&-71&50\\ 73&4&4&75&7&-32\\ 3&29&73&7&60&49\end{bmatrix}.

Here and in similar examples, we use δ\delta to denote mini[m]δ(𝐫i)\min_{i\in[m]}\delta({\bf r}_{i}) where the 𝐫i{\bf r}_{i} are the rows of the matrix. To apply Theorem 2.7 to the matrix AA, observe that the value δ=373>.004\delta=\frac{3}{73}>.004 is obtained from row 55 of AA and the value ϵ=71146335965<0.006\epsilon={71\over\sqrt{146335965}}<0.006 is obtained from rows 11 and 44. Since r5r_{5} is increasing on its domain, r5(ϵ)<r5(0.006)<0.03<δr_{5}(\epsilon)<r_{5}(0.006)<0.03<\delta. We may therefore apply Theorem 2.7 to conclude that there exists a row orthogonal matrix with the same sign-pattern as AA.

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 m×nm\times n sign patterns that minimally allow orthogonality for m5m\leq 5.

3.1 Tools

Recall that an m×nm\times n matrix AA has the SIPP provided OO is the only symmetric matrix XX satisfying (XA)A=O(XA)\circ A=O. 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 QQ is an m×nm\times n full rank matrix with orthogonal rows that has the SIPP and DD is any m×mm\times m diagonal matrix with every diagonal entry nonzero. Then DQDQ has the SIPP. Furthermore, DD can be chosen so that DQDQ is row orthogonal.

Proof 3.2.

Let Xsym(m)X\in\operatorname{sym}(m) such that (XDQ)(DQ)=O(XDQ)\circ(DQ)=O. By algebraic manipulation,

O=(XDQ)(DQ)=(DXD)QQ.O=(XDQ)\circ(DQ)=(DXD)Q\circ Q.

Since DXDsym(m)DXD\in\operatorname{sym}(m) and QQ has the SIPP, it follows that DXD=ODXD=O, which implies X=OX=O. Thus DQDQ has the SIPP. Let 𝐫iT{\bf r}_{i}^{T} denote the iith row of QQ. Define D=diag(1𝐫1,,1𝐫m)D=\operatorname{diag}\left(\frac{1}{\|{\bf r}_{1}\|},\dots,\frac{1}{\|{\bf r}_{m}\|}\right) and Q^=DQ\widehat{Q}=DQ, so Q^𝒪(m,n)\widehat{Q}\in\mathcal{O}(m,n). 

The next three lemmas showcase additional hypotheses on AA that imply various entries in XX are 0.

Lemma 3.3.

Let Am×nA\in\mathbb{R}^{m\times n} be a wide matrix with full rank and let Xm×nX\in\mathbb{R}^{m\times n}. Suppose that every entry of row kk of AA is nonzero. If (XA)A=O(XA)\circ A=O, then every entry of row kk of XX is zero.

Proof 3.4.

Suppose (XA)A=O(XA)\circ A=O. Let 𝐫1T,,𝐫mT{\bf r}_{1}^{T},\dots,{\bf r}_{m}^{T} denote the rows of XX. Then (XA)A=O(XA)\circ A=O implies that 𝐫kTA=𝟎T{\bf r}_{k}^{T}A={\bf 0}^{T}. Since AA has full rank there exists a matrix BB such that AB=IAB=I and so 𝟎T=𝐫kTAB=𝐫kT{\bf 0}^{T}={\bf r}_{k}^{T}AB={\bf r}_{k}^{T}.

Lemma 3.5.

Suppose Q𝒪(m,n)Q\in\mathcal{O}(m,n), Xsym(m)X\in\operatorname{sym}(m) and (XQ)Q=O(XQ)\circ Q=O. Then IX=OI\circ X=O.

Proof 3.6.

Let Y=XQY=XQ and write Y=[yij]Y=[y_{ij}], X=[xij]X=[x_{ij}], and Q=[qij]Q=[q_{ij}]. Since QQ is row orthogonal, X=YQTX=YQ^{T}. The condition that (XQ)Q=O(XQ)\circ Q=O implies that yij=0y_{ij}=0 if qij0q_{ij}\not=0. Therefore,

xii=(YQT)ii=j=1myijqij=0.x_{ii}=(YQ^{T})_{ii}=\sum_{j=1}^{m}y_{ij}q_{ij}=0.

In other words, IX=OI\circ X=O.

Lemma 3.7.

Let Q=[qij]𝒪(m,n)Q=[q_{ij}]\in\mathcal{O}(m,n) and X=[xij]sym(m)X=[x_{ij}]\in\operatorname{sym}(m) satisfy the equation (XQ)Q=O(XQ)\circ Q=O. Suppose that the only two nonzero entries in column jj of QQ are qijq_{ij} and qkjq_{kj}. Then xik=xki=0x_{ik}=x_{ki}=0.

Proof 3.8.

Since (XQ)Q=O(XQ)\circ Q=O,

0=((XQ)Q)ij=(=1mxiqj)qij=(xiiqij+xikqkj)qij.0=((XQ)\circ Q)_{ij}=\left(\sum_{\ell=1}^{m}x_{i\ell}q_{\ell j}\right)q_{ij}=(x_{ii}q_{ij}+x_{ik}q_{kj})q_{ij}.

Since qij0q_{ij}\neq 0, xiiqij+xikqkj=0x_{ii}q_{ij}+x_{ik}q_{kj}=0. By Lemma 3.5, xii=0x_{ii}=0, so xikqkj=0x_{ik}q_{kj}=0. Since qkj0q_{kj}\neq 0, xik=0x_{ik}=0.

The next result extends one direction of [6, Proposition 3.9].

Lemma 3.9.

Suppose AA is a wide matrix partitioned as a 2×22\times 2 block matrix A=[A1A2A3A4]A=\left[\begin{array}[]{c|c}A_{1}&A_{2}\\ \hline\cr A_{3}&A_{4}\\ \end{array}\right] with A3,A4A_{3},A_{4} both nowhere zero (or vacuous). If A1A_{1} has the SIPP and AA is full rank, then AA has the SIPP.

Proof 3.10.

Let X=[X1X2X2TX4]X=\left[\begin{array}[]{c|c}X_{1}&X_{2}\\ \hline\cr X_{2}^{T}&X_{4}\end{array}\right] be a symmetric matrix such that (XA)A=O(XA)\circ A=O. By Lemma 3.3, [X2TX4]=O\left[\begin{array}[]{c|c}X_{2}^{T}&X_{4}\end{array}\right]=O. Therefore, X=[X1OOO]X=\left[\begin{array}[]{c|c}X_{1}&O\\ \hline\cr O&O\end{array}\right]. The equation (XA)A=O(XA)\circ A=O implies that (X1A1)A1=O(X_{1}A_{1})\circ A_{1}=O. Since X1X_{1} is symmetric and A1A_{1} has the SIPP, X1=OX_{1}=O and AA 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 \mathcal{M} and 𝒩\mathcal{N}, both in d\mathbb{R}^{d}, intersect transversally at a point 𝐱{\bf x} if 𝐱𝒩{\bf x}\in\mathcal{M}\cap\mathcal{N} and the intersection of the normal spaces of \mathcal{M} at 𝐱{\bf x} and of 𝒩\mathcal{N} at 𝐱{\bf x} contains only the zero vector. As the next result shows, a matrix Q𝒪(m,n)Q\in\mathcal{O}(m,n) having the SIPP is equivalent to 𝒬(sgn(Q))\mathcal{Q}(\operatorname{sgn}(Q)) and 𝒪(m,n)\mathcal{O}(m,n) intersecting transversally at QQ.

Theorem 3.11.

[6, Theorem 4.5] Let Q𝒪(m,n)Q\in\mathcal{O}(m,n) have sign pattern SS. The manifolds Q(S)Q(S) and 𝒪(m,n)\mathcal{O}(m,n) intersect transversally at QQ if and only if QQ has the SIPP. If QQ has the SIPP, then every superpattern of SS allows o-SIPP.

Theorem 3.13 improves the previous result by allowing us to control the relative magnitudes of the formerly zero entries in QQ when applying the SIPP. This requires the following theorem of van der Holst, Lovász, and Schrijver.

Theorem 3.12.

[9] Let 1(s)\mathcal{M}_{1}(s) and 2(t)\mathcal{M}_{2}(t) be smooth families of manifolds in d\mathbb{R}^{d}, and assume that 1(0)\mathcal{M}_{1}(0) and 2(0)\mathcal{M}_{2}(0) intersect transverally at 𝐲0{\bf y}_{0}. Then there is a neighborhood W2W\subseteq\mathbb{R}^{2} of the origin and a continuous function f:Wdf:W\to\mathbb{R}^{d} such that f(0,0)=𝐲0f(0,0)={\bf y}_{0} and for each ϵ=(ϵ1,ϵ2)W\epsilon=(\epsilon_{1},\epsilon_{2})\in W, 1(ϵ1)\mathcal{M}_{1}(\epsilon_{1}) and 2(ϵ2)\mathcal{M}_{2}(\epsilon_{2}) intersect transversally at f(ϵ)f(\epsilon).

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 ff as a continuous function of one variable from an interval about the origin to d\mathbb{R}^{d}.

Theorem 3.13.

Let Q𝒪(m,n)Q\in\mathcal{O}(m,n) have sign pattern SS. If QQ has the SIPP, then for all Am×nA\in\mathbb{R}^{m\times n} with AQ=OA\circ Q=O, for every ϵ\epsilon sufficiently small there is a matrix Mϵ𝒬(S)M_{\epsilon}\in\mathcal{Q}(S) such that Mϵ+ϵA𝒪(m,n)M_{\epsilon}+\epsilon A\in\mathcal{O}(m,n). Moreover, Mϵ+ϵAM_{\epsilon}+\epsilon A has the SIPP.

Proof 3.14.

Suppose that Q=[qij]Q=[q_{ij}] has the SIPP and let A=[aij]m×nA=[a_{ij}]\in\mathbb{R}^{m\times n} satisfy AQ=OA\circ Q=O. Define the smooth family of manifolds A(t)\mathcal{M}_{A}(t) by

A(t)={B=[bij]m×n:sgn(bij)=sgn(qij) if qij0, and bij=aijt if qij=0}\mathcal{M}_{A}(t)=\{B=[b_{ij}]\in\mathbb{R}^{m\times n}:\operatorname{sgn}(b_{ij})=\operatorname{sgn}(q_{ij})\text{ if }q_{ij}\not=0,\text{ and }b_{ij}=a_{ij}t\text{ if }q_{ij}=0\}

for t(1,1)t\in(-1,1). Since QQ has the SIPP, 𝒪(m,n)\mathcal{O}(m,n) and A(0)=𝒬(S)\mathcal{M}_{A}(0)=\mathcal{Q}(S) intersect transversally at QQ by Theorem 3.11. By Theorem 3.12, there exists a continuous function f:(1,1)m×nf:(-1,1)\to\mathbb{R}^{m\times n} such that f(0)=Qf(0)=Q and the manifolds A(ϵ)\mathcal{M}_{A}(\epsilon) and 𝒪(m,n)\mathcal{O}(m,n) intersect transversally at f(ϵ)f(\epsilon) for each ϵ>0\epsilon>0 sufficiently small. Since ff is continuous we may choose ϵ\epsilon small enough so that Mϵ:=f(ϵ)CZ𝒬(S)M_{\epsilon}:=f(\epsilon)\circ C_{Z}\in\mathcal{Q}(S), where ZZ is the zero-nonzero pattern of QQ and CZC_{Z} is the unique (1,0)(1,0)-matrix in 𝒬(Z)\mathcal{Q}(Z). To complete the proof, observe that f(ϵ)=Mϵ+ϵAf(\epsilon)=M_{\epsilon}+\epsilon A. Moreover, f(ϵ)f(\epsilon) has the SIPP by Theorem 3.11 since A(ϵ)\mathcal{M}_{A}(\epsilon) and 𝒪(m,n)\mathcal{O}(m,n) intersect transversally at f(ϵ)f(\epsilon).

We apply the previous theorem to prove the next result.

Proposition 3.15.

Let

S=[S1OS3S4]S=\left[\begin{array}[]{c|c}S_{1}&O\\ \hline\cr S_{3}&S_{4}\end{array}\right]

be a sign pattern that allows row orthogonality, and let S4S_{4}^{\prime} be a submatrix of S4S_{4} with the same number of rows as S4S_{4}. If S4S_{4}^{\prime} allows o-SIPP, then

S=[S1OS3S4]S^{\prime}=\left[\begin{array}[]{c|c}S_{1}&O\\ \hline\cr S_{3}&S_{4}^{\prime}\end{array}\right]

allows row orthogonality.

Proof 3.16.

Let QQ be a row orthogonal realization of SS. Then

Q=[Q1OQ3Q4],Q=\left[\begin{array}[]{c|c}Q_{1}&O\\ \hline\cr Q_{3}&Q_{4}\end{array}\right],

where the partition is consistent with that of SS. Assume that S4S_{4}^{\prime} allows o-SIPP. Then there exists a row orthogonal realization Q4Q_{4}^{\prime} of S4S_{4}^{\prime} with the SIPP. Then [OQ4]\left[\begin{array}[]{c|c}O&Q_{4}^{\prime}\end{array}\right] is row orthogonal and, by Lemma 1.2, has the SIPP. By Theorem 3.13, there exists an ϵ>0\epsilon>0 and a matrix MϵM_{\epsilon}^{\prime} such that [ϵQ3Mϵ]\left[\begin{array}[]{c|c}\epsilon Q_{3}&M_{\epsilon}^{\prime}\end{array}\right] is row orthogonal and sgn(Q4)=sgn(Mϵ)\operatorname{sgn}(Q_{4}^{\prime})=\operatorname{sgn}(M_{\epsilon}^{\prime}). Since QQ is row orthogonal, Q1Q3T=OQ_{1}Q_{3}^{T}=O. Therefore,

Q=[Q1OϵQ3Mϵ]Q^{\prime}=\left[\begin{array}[]{c|c}Q_{1}&O\\ \hline\cr\epsilon Q_{3}&M_{\epsilon}^{\prime}\end{array}\right]

is row orthogonal and sgn(Q)=S\operatorname{sgn}(Q^{\prime})=S^{\prime}.

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 SS play an important role in determining whether SS 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 n×nn\times n lower Hessenberg matrix

H=[110020(n1)11],H=\left[\begin{array}[]{ccccc}1&-1&0&\cdots&0\\ \vdots&\ddots&-2&\ddots&\vdots\\ \vdots&&\ddots&\ddots&0\\ \vdots&&&\ddots&-(n-1)\\ 1&\cdots&\cdots&\cdots&1\end{array}\right],

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 HH 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 k=1m,,n1k=1-m,\dots,n-1, the kk-th diagonal of an m×nm\times n matrix A=[aij]A=[a_{ij}] is the list of entries aija_{ij} such that ji=kj-i=k. The kk-th diagonal terminology is also applied to sign patterns.

Lemma 3.17.

Let A=[aij]m×nA=[a_{ij}]\in\mathbb{R}^{m\times n} be a wide matrix with full rank. Suppose that there is an integer kk such that 0kn10\leq k\leq n-1, each entry of AA on the rr-th diagonal is nonzero for 1mrk1-m\leq r\leq k, and each entry of AA on the rr-th diagonal is zero for k<rn1k<r\leq n-1. Then AA has the SIPP.

Proof 3.18.

Note that if k=n1k=n-1, then AA is nowhere zero and hence has the SIPP. Suppose that k<n1k<n-1. Let c=min{nk1,m}c=\min\{n-k-1,m\} and A=A[{1,,},{1,+k+1}]A_{\ell}=A[\{1,\ldots,\ell\},\{1\ldots,\ell+k+1\}] for =1,,c\ell=1,\ldots,c. We begin by successively showing that each AA_{\ell} has the SIPP. Since A1A_{1} contains a nonzero entry, Lemma 3.9 implies A1A_{1} has the SIPP. Suppose that AiA_{i} has the SIPP for some i{1,,c1}i\in\{1,\ldots,c-1\}. Then Lemma 3.9 and Lemma 1.2 imply that Ai+1A_{i+1} has the SIPP. If c=mc=m, then A=[AcO]A=\left[\begin{array}[]{c|c}A_{c}&O\end{array}\right] has the SIPP by Lemma 1.2. Otherwise, A=[AcB]A=\left[\begin{array}[]{c}A_{c}\\ \hline\cr B\end{array}\right], where BB is nowhere zero. By Lemma 3.9, AA has the SIPP.

Corollary 3.19.

Let SS be an m×nm\times n wide sign pattern. Suppose that there is an integer kk such that 0kn10\leq k\leq n-1, each entry of SS on the rr-th diagonal is nonzero for 1mrk1-m\leq r\leq k, and each entry of SS on the rr-th diagonal is zero for k<rn1k<r\leq n-1. If SS allows row orthogonality, then SS requires o-SIPP.

For sign equivalent matrices AA and BB, A𝒪(m,n)A\in\mathcal{O}(m,n) implies B𝒪(m,n)B\in\mathcal{O}(m,n) and AA has the SIPP implies BB has the SIPP. Thus the analogous statement with the upper part nonzero is also true.

Corollary 3.20.

Let SS be a wide m×nm\times n sign pattern. Suppose that there is an integer kk such that 1mk01-m\leq k\leq 0, each entry of SS on the rr-th diagonal is nonzero for krn1k\leq r\leq n-1, and each entry of SS on the rr-th diagonal is zero for 1mr<k1-m\leq r<k. If SS allows row orthogonality, then SS 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 ±1\pm 1. Matrices AA and BB are signature equivalent if there exist signature matrices D1D_{1} and D2D_{2} such that D1AD2=BD_{1}AD_{2}=B. Similarly, sign patterns SS and RR are signature equivalent if there exist signature matrices D1D_{1} and D2D_{2} such that D1CSD2=CRD_{1}C_{S}D_{2}=C_{R}. Theorem 5.7 in [6] states that a nonzero hollow matrix Q𝒪(n)Q\in\mathcal{O}(n) has the SIPP if and only if QQ is not signature equivalent to a symmetric hollow matrix. The following corollary is an immediate consequence.

Corollary 3.21.

Let SS be a nonzero hollow sign pattern that allows orthogonality. If SS is not signature equivalent to a symmetric hollow sign pattern, then SS requires o-SIPP. If SS is signature equivalent to a symmetric hollow sign pattern, then SS does not allow o-SIPP.

As an example, consider

S=[0++++0+++0++0].S=\left[\begin{array}[]{cccc}0&+&+&+\\ +&0&-&+\\ +&+&0&-\\ +&-&+&0\end{array}\right].

It is not difficult to verify that CSC_{S} has orthogonal rows, and hence SS allows orthogonality. Further, SS is not signature equivalent to a symmetric sign pattern. Thus, SS requires o-SIPP.

Let GG be a graph with vertex set V(G)={v1,,vm}V(G)=\{v_{1},\dots,v_{m}\} and edge set E(G)={e1,,e}E(G)=\{e_{1},\dots,e_{\ell}\}. The (vertex-edge) incidence matrix RG=[rij]R_{G}=[r_{ij}] of GG is the m×m\times\ell matrix that has rij=1r_{ij}=1 if viejv_{i}\in e_{j} and rij=0r_{ij}=0 otherwise. An orientation G\vec{G} of GG is the assignment of a direction to each edge. That is, the edge ej={vi,vk}e_{j}=\{v_{i},v_{k}\} is replaced by exactly one of the two arcs (vi,vk)(v_{i},v_{k}) or (vk,vi)(v_{k},v_{i}); the arc associated with eje_{j} is denoted by ej\vec{e}_{j}. The incidence matrix RG=[rij]R_{\vec{G}}=[r_{ij}] of an orientation G\vec{G} of GG is the m×m\times\ell matrix that has rij=1r_{ij}=-1 if ej=(vi,vk)\vec{e}_{j}=(v_{i},v_{k}), rij=1r_{ij}=1 if ej=(vk,vi)\vec{e}_{j}=(v_{k},v_{i}), and rij=0r_{ij}=0 otherwise.

Consider the complete graph KmK_{m} and an orientation Km\vec{K}_{m}. For m2m\geq 2, define the m×2(m2)m\times 2\binom{m}{2} matrix R(Km,Km)=[RKm|RKm]R(K_{m},\vec{K}_{m})=[R_{K_{m}}|R_{\vec{K}_{m}}] and its sign pattern S(Km,Km)=sgn(R(Km,Km))S(K_{m},\vec{K}_{m})=\operatorname{sgn}(R(K_{m},\vec{K}_{m})). The sign pattern S(Km,Km)S(K_{m},\vec{K}_{m}) was shown to have a row orthogonal realization with the SIPP in [5]. We now show that S(Km,Km)S(K_{m},\vec{K}_{m}) 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 m2m\geq 2 and any orientation Km\vec{K}_{m} of KmK_{m}, the sign pattern S(Km,Km)S(K_{m},\vec{K}_{m}) requires o-SIPP.

Proof 3.23.

For brevity, let S=S(Km,Km)S=S(K_{m},\vec{K}_{m}). It is not difficult to verify that CSC_{S} has orthogonal rows and hence SS allows row orthogonality. Let Q𝒬(S)Q\in\mathcal{Q}(S) be row orthogonal and suppose X=[xij]sym(m)X=[x_{ij}]\in\operatorname{sym}(m) satisfies (XQ)Q=O(XQ)\circ Q=O. By Lemma 3.5, xii=0x_{ii}=0 for i=1,,mi=1,\dots,m. For iki\neq k, there is a unique edge ej={vi,vk}e_{j}=\{v_{i},v_{k}\}. Then applying Lemma 3.7 to column jj gives that xik=0x_{ik}=0. Thus X=OX=O and QQ has the SIPP.

Let Q𝒪(m,n)Q\in\mathcal{O}(m,n). It is an immediate consequence of Theorem 1.1 that if QQ has a pair of combinatorially orthogonal rows, then QQ does not have the SIPP. Similarly, if m=nm=n and QQ has a pair of combinatorially orthogonal columns, then QQ 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

S=[+0++0++0+]andA=[101101110211].S=\left[\begin{array}[]{cccc}+&0&+&+\\ 0&+&+&-\\ 0&-&+&-\end{array}\right]\quad\text{and}\quad A=\left[\begin{array}[]{cccc}1&0&1&1\\ 0&1&1&-1\\ 0&-2&1&-1\end{array}\right].

Observe that AA has orthogonal rows. As we shall see, Corollary 3.27 implies SS 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 Am×nA\in\mathbb{R}^{m\times n} be a wide matrix. It is not difficult to verify that if AA is nowhere zero, then AA has the SIPP. Thus every nowhere zero sign pattern that allows row orthogonality requires o-SIPP. For each additional 0 entry in AA, the equation (XA)A=O(XA)\circ A=O imposes one fewer linear equation on the entries of XX. This suggests that the more 0 entries AA has, the larger the solution space to (XA)A=O(XA)\circ A=O is going to be, reducing the likelihood that AA 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 Q𝒪(m,n)Q\in\mathcal{O}(m,n) have the SIPP. Then the number of zero entries in QQ is at most nm12m(m+1)nm-\frac{1}{2}m(m+1).

The location of 0 entries in a sign pattern SS also play an important role in determining if SS requires o-SIPP.

Theorem 3.25.

Let Q=[qij]𝒪(m,n)Q=[q_{ij}]\in\mathcal{O}(m,n). Suppose that the zero entries of QQ are contained in at most three rows of QQ and that no pair of rows is combinatorially orthogonal. Then, QQ has the SIPP.

Proof 3.26.

Begin by assuming that the zero entries of QQ are contained in at most 1 row. Without loss of generality, the zero entries of QQ are contained in the first row and the (1,1)(1,1)-entry of QQ is nonzero. By Lemma 3.9, QQ has the SIPP.

Now assume that exactly kk rows of QQ contain a zero, where k{2,3}k\in\{2,3\}. Without loss of generality, the first kk rows each contain a zero. Let Xsym(m)X\in\operatorname{sym}(m) and suppose (XQ)Q=O(XQ)\circ Q=O. By Lemma 3.3, X=YOX=Y\oplus O, where Ysym(k)Y\in\operatorname{sym}(k). Observe that (YQ^)Q^=O(Y\hat{Q})\circ\hat{Q}=O, where Q^\hat{Q} is the submatrix of QQ formed from the first kk rows. Also, note that Lemma 3.5 implies YI=OY\circ I=O.

First, consider the case k=2k=2. Since the rows of Q^\hat{Q} are not combinatorially orthogonal, Q^\hat{Q} has a nowhere zero column. Since (YQ^)Q^=O(Y\hat{Q})\circ\hat{Q}=O, Lemma 3.7 implies that the off-diagonal entries of YY are zero. This, together with YI=OY\circ I=O, implies Y=OY=O.

Now consider the case k=3k=3. Suppose first that an off-diagonal entry of YY, without loss of generality the (1,2)(1,2)-entry, is zero, so Y=[00y100y2y1y20].Y=\left[\begin{array}[]{ccc}0&0&y_{1}\\ 0&0&y_{2}\\ y_{1}&y_{2}&0\\ \end{array}\right]. Since the rows of Q^\hat{Q} are not combinatorially orthogonal, Q^\hat{Q} has a column with nonzero entries in the first and third rows. Then (YQ^)Q^=O(Y\hat{Q})\circ\hat{Q}=O implies y1=0y_{1}=0; similarly y2=0y_{2}=0. So suppose that YY is a nonzero hollow matrix. Then by Lemma 3.7, and the preceding argument, no column of Q^\hat{Q} has exactly one zero entry. This, along with the fact that no pair of rows of Q^\hat{Q} are combinatorially orthogonal, implies Q^\hat{Q} has a nowhere zero column 𝐪j{\bf q}_{j}. Observe that (YQ^)Q^=O(Y\hat{Q})\circ\hat{Q}=O implies Y𝐪j=𝟎Y{\bf q}_{j}={\bf 0}. This is impossible since rank(Y)=3\operatorname{rank}(Y)=3. Thus, Y=OY=O and QQ has the SIPP.

Corollary 3.27.

Suppose SS is an m×nm\times n sign pattern that allows row orthogonality such that the zero entries of SS are contained in at most three rows of SS and that no pair of rows are combinatorially orthogonal. Then, SS requires o-SIPP.

As the next example illustrates, Corollary 3.27 cannot be extended to 4 or more rows.

Example 3.28.
\thlabel

ex:n-zero_n-rowDefine the n×(n+1)n\times(n+1) matrix

A=[0n23n1112120n213n112120n2113n12120n21113n23n2n201111212n201111212].A=\left[\begin{array}[]{cc|cccc|cc}0&\sqrt{n-2}&3-n&1&\cdots&1&\frac{1}{\sqrt{2}}&\frac{1}{\sqrt{2}}\\ 0&\sqrt{n-2}&1&3-n&\cdots&1&\frac{1}{\sqrt{2}}&\frac{1}{\sqrt{2}}\\ \vdots&\vdots&\vdots&\vdots&\ddots&\vdots&\vdots&\vdots\\ 0&\sqrt{n-2}&1&1&\cdots&3-n&\frac{1}{\sqrt{2}}&\frac{1}{\sqrt{2}}\\ \hline\cr 0&\sqrt{n-2}&1&1&\cdots&1&\frac{3-n}{\sqrt{2}}&\frac{3-n}{\sqrt{2}}\\ \hline\cr\sqrt{n-2}&0&1&1&\cdots&1&\frac{1}{\sqrt{2}}&\frac{1}{\sqrt{2}}\\ -\sqrt{n-2}&0&1&1&\cdots&1&\frac{1}{\sqrt{2}}&\frac{1}{\sqrt{2}}\end{array}\right].

Observe that the rows of AA are orthogonal. It can be verified that

X=[0011001111001100]n×nX=\begin{bmatrix}0&\cdots&0&1&-1\\ \vdots&\ddots&\vdots&\vdots&\vdots\\ 0&\cdots&0&1&-1\\ 1&\cdots&1&0&0\\ -1&\cdots&-1&0&0\end{bmatrix}\in\mathbb{R}^{n\times n}

satisfies (XA)A=O(XA)\circ A=O. Since X=XTX=X^{T} and rankA=n\operatorname{rank}A=n, it follows that AA 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

Q=[9900326262990032626200996232620099623262323262628846262323288462626262442]andX=[0011001111001100]O.Q=\left[\begin{array}[]{ccccccc}-9&9&0&0&3\sqrt{2}&-6\sqrt{2}&6\sqrt{2}\\ 9&-9&0&0&3\sqrt{2}&-6\sqrt{2}&6\sqrt{2}\\ 0&0&-9&9&-6\sqrt{2}&3\sqrt{2}&6\sqrt{2}\\ 0&0&9&-9&-6\sqrt{2}&3\sqrt{2}&6\sqrt{2}\\ 3\sqrt{2}&3\sqrt{2}&-6\sqrt{2}&-6\sqrt{2}&8&8&4\\ -6\sqrt{2}&-6\sqrt{2}&3\sqrt{2}&3\sqrt{2}&8&8&4\\ 6\sqrt{2}&6\sqrt{2}&6\sqrt{2}&6\sqrt{2}&4&4&2\\ \end{array}\right]\quad\text{and}\quad X=\left[\begin{array}[]{cccc}0&0&1&-1\\ 0&0&1&-1\\ 1&-1&0&0\\ 1&-1&0&0\\ \end{array}\right]\oplus O.

It is readily verified that the rows of QQ are orthogonal and (XQ)Q=O(XQ)\circ Q=O. Thus, sgn(Q)\operatorname{sgn}(Q) allows orthogonality but does not require o-SIPP.

By restricting the number of zero entries, we obtain the following result.

Proposition 3.30.

Let Q𝒪(m,n)Q\in\mathcal{O}(m,n) have at most four zero entries. Suppose that no pair of rows and no pair of columns columns of QQ are combinatorially orthogonal. Then QQ has the SIPP.

Proof 3.31.

Observe that if the zero entries of Q=[qij]Q=[q_{ij}] are contained in at most three rows, then Theorem 3.25 implies QQ has the SIPP. So, suppose that each zero entry of QQ is contained in a unique row.

Assume first that m=n=4m=n=4. If QQ has a nowhere zero column, then Proposition 1.5 and Theorem 3.25 imply QQ has the SIPP. Otherwise, without loss of generality, QQ is a nonzero hollow matrix. Theorem 3.2 in [2] guarantees that QQ is not symmetric. Thus, Corollary 3.21 implies QQ has the SIPP.

Now assume that n5n\geq 5 (and m4m\geq 4). Without loss of generality, QQ has the form

Q=[Q1Q2Q3Q4],Q=\left[\begin{array}[]{c|c}Q_{1}&Q_{2}\\ \hline\cr Q_{3}&Q_{4}\end{array}\right],

where all four zero entries of QQ are contained in Q1Q_{1} such that each column and each row of Q1Q_{1} contains a zero entry. Note that Q1Q_{1} has four rows, Q2Q_{2} is nowhere zero, and Q3Q_{3} and Q4Q_{4} may be vacuous. We proceed via cases.

We first resolve the case where Q1Q_{1} has exactly one column, i.e., Q1=𝟎Q_{1}={\bf 0}. Then Q2𝒪(4,n1)Q_{2}\in\mathcal{O}(4,n-1) is nowhere zero. Thus Q2Q_{2} has the SIPP and, by Lemma 3.9, QQ has the SIPP.

Suppose that Q1Q_{1} has at least two columns. Let Xsym(m)X\in\operatorname{sym}(m) and suppose (XQ)Q=O(XQ)\circ Q=O. By Lemma 3.3, X=YOX=Y\oplus O, where YY is symmetric and has four rows. Further, Lemma 3.5 implies YI=OY\circ I=O, i.e., YY has the form

Y=[0y1y2y3y10y4y5y2y40y6y3y5y60].Y=\left[\begin{array}[]{cccc}0&y_{1}&y_{2}&y_{3}\\ y_{1}&0&y_{4}&y_{5}\\ y_{2}&y_{4}&0&y_{6}\\ y_{3}&y_{5}&y_{6}&0\end{array}\right].

We now consider the case where Q1Q_{1} has a column 𝐜{\bf c} with exactly one zero entry. Without loss of generality, the zero appears in the last entry of 𝐜{\bf c}. Then (Y𝐜)𝐜=𝟎(Y{\bf c})\circ{\bf c}={\bf 0} implies y1=y2=y4=0y_{1}=y_{2}=y_{4}=0. It now follows from (YQ2)Q2=O(YQ_{2})\circ Q_{2}=O that y3=y5=y6=0y_{3}=y_{5}=y_{6}=0. Thus, Y=OY=O and QQ has the SIPP.

For the final case, assume that Q1Q_{1} 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,

Q^:=Q[{1,,5},{1,,n}]=[0q12𝐯1T0q22𝐯2Tq310𝐯3Tq410𝐯4Tq51q52𝐰T].\hat{Q}:=Q[\{1,\ldots,5\},\{1,\ldots,n\}]=\left[\begin{array}[]{cc|c}0&q_{12}&{\bf v}_{1}^{T}\\ 0&q_{22}&{\bf v}_{2}^{T}\\ q_{31}&0&{\bf v}_{3}^{T}\\ q_{41}&0&{\bf v}_{4}^{T}\\ \hline\cr q_{51}&q_{52}&{\bf w}^{T}\end{array}\right].

From (YQ1)Q1=O(YQ_{1})\circ Q_{1}=O it follows that y1=y6=0y_{1}=y_{6}=0. If yi=0y_{i}=0 for i{2,3,4,5}i\in\{2,3,4,5\}, then (YQ2)Q2=O(YQ_{2})\circ Q_{2}=O implies Y=OY=O. Suppose this is not the case, i.e., yi0y_{i}\not=0 for all i{2,3,4,5}i\in\{2,3,4,5\}. Then (YQ2)Q2=O(YQ_{2})\circ Q_{2}=O implies y2𝐯1T+y4𝐯2T=𝟎Ty_{2}{\bf v}_{1}^{T}+y_{4}{\bf v}_{2}^{T}={\bf 0}^{T}, and thus 𝐯2=a𝐯1{\bf v}_{2}=a{\bf v}_{1} for some nonzero value aa. Since QQ is row orthogonal

0\displaystyle 0 =\displaystyle= q12q22+𝐯1T𝐯2=q12q22+a(1q122),\displaystyle q_{12}q_{22}+{\bf v}_{1}^{T}{\bf v}_{2}=q_{12}q_{22}+a(1-q_{12}^{2}),
0\displaystyle 0 =\displaystyle= q22q52+𝐯2T𝐰=q22q52+a𝐯1T𝐰, and\displaystyle q_{22}q_{52}+{\bf v}_{2}^{T}{\bf w}=q_{22}q_{52}+a{\bf v}_{1}^{T}{\bf w},\text{ and}
0\displaystyle 0 =\displaystyle= q12q52+𝐯1T𝐰.\displaystyle q_{12}q_{52}+{\bf v}_{1}^{T}{\bf w}.

From the last two equations a=q22/q12a=q_{22}/q_{12}. Substituting this into the first equation implies q22=0q_{22}=0, a contradiction. Thus, Y=OY=O and QQ has the SIPP.

It is possible to show that if Q𝒪(m,n)Q\in\mathcal{O}(m,n) has at most five zero entries, no pair of rows and no pair of columns columns of QQ are combinatorially orthogonal, then QQ 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

S=[0++++++0++++0+++0+++0++++0].S=\left[\begin{array}[]{cccccc}0&+&+&+&+&+\\ +&0&+&-&+&-\\ +&+&0&+&-&-\\ +&-&+&0&-&+\\ +&+&-&-&0&+\\ +&-&-&+&+&0\end{array}\right].

Observe that CSC_{S} is a conference matrix, i.e., CSC_{S} is hollow, every off-diagonal entry is 11 or 1-1, and CSTCS=(n1)IC_{S}^{T}C_{S}=(n-1)I. Hence SS allows orthogonality. By Corollary 3.21, SS does not require o-SIPP. In fact, it is not difficult to see that CSC_{S} does not have the SIPP since the symmetric matrix X=CSX=C_{S} satisfies (XCS)CS=O(XC_{S})\circ C_{S}=O.

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 m×nm\times n nowhere zero sign patterns that minimally allow orthogonality for m5m\leq 5. Recall that a sign pattern SS minimally allows orthogonality provided SS allows row orthogonality and every sign pattern obtained from SS 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 SS be an m×nm\times n nowhere zero sign pattern. If m3m\leq 3, then SS minimally allows orthogonality if and only if n=3n=3 and SS is row and column PPO. If m=4m=4, then SS minimally allows orthogonality if and only if n=4n=4 and SS is row and column PPO, or SS is sign equivalent to

[+++++++++++++++].\left[\begin{array}[]{ccccc}-&-&+&+&+\\ +&+&-&-&+\\ +&+&+&+&-\\ +&+&+&+&+\end{array}\right].

We now determine all 5×n5\times n nowhere zero sign patterns that minimally allow orthogonality. The next theorem establishes the square case.

Theorem 3.34.

[5, Theorem 7.9] Let SS be a 5×55\times 5 nowhere zero sign pattern. Then SS allows orthogonality if and only if SS is row and column PPO.

Lemma 3.35.

Let SS be a 5×45\times 4 nowhere zero sign pattern. Then SS is sign equivalent to a sign pattern with at most 5 negative entries.

Proof 3.36.

By scaling the rows and columns of SS we can obtain the sign patterns

S1=[++++++R++]andS2=[+++++R++].S_{1}=\left[\begin{array}[]{c|ccc}+&+&+&+\\ \hline\cr+&\\ +&&\hbox{\multirowsetup$R$}\\ +&\\ +&\\ \end{array}\right]\quad\text{and}\quad S_{2}=\left[\begin{array}[]{c|ccc}-&+&+&+\\ \hline\cr+&\\ +&&\hbox{\multirowsetup$-R$}\\ +&\\ +&\\ \end{array}\right].

If RR contains at most 5 negative entries, then the proof is complete. So, suppose that RR has at least 6 negative entries.

First consider the case where RR has exactly 6 negative entries. If a row (or column) of RR has 3 negatives, then negating the corresponding row (or column) of S1S_{1} reduces the total number of negative entries to at most 5. Otherwise RR has two rows 𝐫1{\bf r}_{1} and 𝐫2{\bf r}_{2}, each containing exactly 2 negative entries, and a third row that contains a negative entry; let jj denote the column index of this entry. Observe that negating the rows of S1S_{1} corresponding to 𝐫1{\bf r}_{1} and 𝐫2{\bf r}_{2} does not change the total number of negative entries. Thus, we can scale the rows of S1S_{1} so that column jj has 3 negative entries. Negating column jj now reduces the total number of negatives to at most 5.

Now suppose that RR has at least 7 negative entries. Then S2S_{2} 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 R-R contains at least 3 negative entries, or if R-R has two rows that each contain exactly 2 negative entries. Thus, we may assume R-R has 1 row with exactly 2 negative entries and 2 columns with exactly 2 negative entries. Without loss of generality S2S_{2} is of the form

[++++++++++++++]or[++++++++++++++].\left[\begin{array}[]{c|ccc}-&+&+&+\\ \hline\cr+&-&-&+\\ +&+&+&-\\ +&+&-&+\\ +&-&+&+\end{array}\right]\quad\text{or}\quad\left[\begin{array}[]{c|ccc}-&+&+&+\\ \hline\cr+&-&-&+\\ +&+&+&-\\ +&+&+&-\\ +&+&-&+\end{array}\right].

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

A1=[4242974238242421229048578702473921263225364903104664394043054074957938412255301] and A2=[246246369123369123494254712773141742301142139675284107414564139224775136769231]A_{1}={\small\left[\begin{array}[]{cccccc}-424&-297&42&382&424&212\\ 290&48&-578&-70&247&392\\ 126&32&2&536&-490&310\\ 466&4&39&404&305&-407\\ 49&579&384&12&255&301\end{array}\right]}\text{ and }A_{2}={\small\left[\begin{array}[]{cccccc}-246&-246&369&123&369&123\\ 494&-254&7&127&7&314\\ 174&230&-11&-421&396&75\\ 284&107&414&56&-41&-392\\ 2&477&51&367&69&231\end{array}\right]}

to obtain row orthogonal matrices with the same sign patterns: For A1A_{1}, the value δ=1268>.003\delta=\frac{1}{268}>.003 is obtained from row 33 of A1A_{1} and the value ϵ=400409890583973409890583973<0.0007\epsilon={400\over 409890583973\sqrt{409890583973}}<0.0007 is obtained from rows 11 and 22. Thus ϵ<151\epsilon<\frac{1}{5-1}. Since r5r_{5} is increasing on its domain, r5(ϵ)<r5(0.0007)<0.003<δr_{5}(\epsilon)<r_{5}(0.0007)<0.003<\delta. For A2A_{2}, the value δ=2477>.004\delta=\frac{2}{477}>.004 is obtained from row 55 of A2A_{2} and the value ϵ=1395150118545<0.0009\epsilon={1\over 395150\sqrt{118545}}<0.0009 is obtained from rows 11 and 22. Thus ϵ<151\epsilon<\frac{1}{5-1} and r5(ϵ)<r5(0.0009)<0.004<δr_{5}(\epsilon)<r_{5}(0.0009)<0.004<\delta.

Theorem 3.38.

Let SS be a 5×n5\times n nowhere zero sign pattern. Then SS minimally allows orthogonality if and only if n=5n=5 and SS is row and column PPO, or SS is sign equivalent to

S1=[++++++++++++++++++++++++],S2=[+++++++++++++++++++++++],orS3=[+++++++++++++++++++++++].S_{1}={\small\left[\begin{array}[]{cccccc}-&-&+&+&+&+\\ +&+&-&-&+&+\\ +&+&+&+&-&+\\ +&+&+&+&+&-\\ +&+&+&+&+&+\end{array}\right]},\ S_{2}={\small\left[\begin{array}[]{cccccc}-&-&-&+&+&+\\ +&+&-&+&+&+\\ +&+&+&-&-&+\\ +&+&+&+&+&-\\ +&+&+&+&+&+\end{array}\right]},\ \text{or}\ S_{3}={\small\left[\begin{array}[]{cccccc}-&-&+&+&+&+\\ +&-&+&+&+&+\\ +&+&-&-&+&+\\ +&+&+&+&-&-\\ +&+&+&+&+&+\end{array}\right]}.

Proof 3.39.

Observe that each of the three patterns S1,S2S_{1},S_{2}, and S3S_{3} allows row orthogonality by Examples 2.9 and 3.37. Removing a column from one of S1,S2S_{1},S_{2} or S3S_{3} results in a 5×55\times 5 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 S1,S2S_{1},S_{2} and S3S_{3} results in a sign pattern that does not allow orthogonality. Thus, each of S1,S2S_{1},S_{2} and S3S_{3} minimally allows row orthogonality.

Assume that SS minimally allows orthogonality. Without loss of generality the first row and first column of SS have all positive entries. Suppose that SS has dd distinct columns 𝐜1,,𝐜d{\bf c}_{1},\ldots,{\bf c}_{d}. It is easy to see that d4d\geq 4: If SS had at most 3 distinct columns, then SS would have at most 44 distinct rows, contradicting the fact that SS is row PPO.

First consider the case where SS has d=5d=5 distinct columns. Since SS is row PPO,

R=[𝐜1𝐜2𝐜3𝐜4𝐜5]R=\left[\begin{array}[]{c|c|c|c|c}{\bf c}_{1}&{\bf c}_{2}&{\bf c}_{3}&{\bf c}_{4}&{\bf c}_{5}\\ \end{array}\right]

is row PPO. Observe that RR is column PPO since 𝐜1,,𝐜5{\bf c}_{1},\ldots,{\bf c}_{5} are distinct. By Theorem 3.34, RR allows orthogonality. Since SS minimally allows orthogonality, S=RS=R.

Now suppose that d>5d>5 and let

R=[𝐜1𝐜2𝐜3𝐜4𝐜5].R=\left[\begin{array}[]{c|c|c|c|c}{\bf c}_{1}&{\bf c}_{2}&{\bf c}_{3}&{\bf c}_{4}&{\bf c}_{5}\\ \end{array}\right].

As before, if RR is row PPO, then RR allows orthogonality. Since SS minimally allows orthogonality, it follows from the preceding argument that RR is not row PPO. Without loss of generality

R=[++++++++++++R^+],R=\left[\begin{array}[]{r|rrrr}+&+&+&+&+\\ +&+&+&+&+\\ \hline\cr+&\\ +&&\lx@intercol\hfil\hat{R}\hfil\lx@intercol\\ +&\\ \end{array}\right],

where the columns of R^\hat{R} are distinct and belong to the set

{[++],[++],[++],[+],[+],[+],[]}.\left\{\left[\begin{array}[]{ccc}+\\ +\\ -\end{array}\right],\left[\begin{array}[]{ccc}+\\ -\\ +\end{array}\right],\left[\begin{array}[]{ccc}-\\ +\\ +\end{array}\right],\left[\begin{array}[]{ccc}+\\ -\\ -\end{array}\right],\left[\begin{array}[]{ccc}-\\ +\\ -\end{array}\right],\left[\begin{array}[]{ccc}-\\ -\\ +\end{array}\right],\left[\begin{array}[]{ccc}-\\ -\\ -\end{array}\right]\right\}.

Either R^\hat{R} contains a column with exactly one negative entry or every column of R^\hat{R} has at least two negative entries. Observe that in the latter case, negating the last three rows of RR results in a column with all positive entries and a column with exactly one negative entry. Thus, we may assume

R=[++++++++++++++++].R=\left[\begin{array}[]{cc|ccc}+&+&+&+&+\\ +&+&+&+&+\\ \hline\cr+&+&+&-&*\\ +&+&-&*&*\\ +&-&*&*&*\end{array}\right].

Since SS allows row orthogonality, 𝐜j=(+,,,,)T{\bf c}_{j}=(+,-,*,*,*)^{T} for some j6j\geq 6. Observe that

[𝐜1𝐜2𝐜3𝐜4𝐜j]\left[\begin{array}[]{c|c|c|c|c}{\bf c}_{1}&{\bf c}_{2}&{\bf c}_{3}&{\bf c}_{4}&{\bf c}_{j}\\ \end{array}\right]

is row and column PPO and hence allows orthogonality by Theorem 3.34. This is a contradiction since SS minimally allows orthogonality. Thus, d5d\leq 5.

Finally, we consider the case where SS has exactly d=4d=4 distinct columns. Let R=[𝐜1𝐜2𝐜3𝐜4]R=\left[\begin{array}[]{c|c|c|c}{\bf c}_{1}&{\bf c}_{2}&{\bf c}_{3}&{\bf c}_{4}\end{array}\right]. By Lemma 3.35 we may assume that RR has at most 55 negative entries. Observe that at least 44 rows of RR contain a negative entry since SS is row PPO.

Suppose RR has exactly 44 negative entries. Then RR is sign equivalent to

[++++++++++++++++]\left[\begin{array}[]{cccc}-&+&+&+\\ +&-&+&+\\ +&+&-&+\\ +&+&+&-\\ +&+&+&+\end{array}\right]

and we assume RR has this form. Observe that SS can be obtained from RR by duplicating some of the columns. By Theorem 1.6 we must duplicate at least 2 distinct columns of RR to obtain SS. It follows that, up to sign equivalence, SS contains the submatrix

S1=[++++++++++++++++++++++++].S_{1}=\left[\begin{array}[]{cccccc}-&-&+&+&+&+\\ +&+&-&-&+&+\\ +&+&+&+&-&+\\ +&+&+&+&+&-\\ +&+&+&+&+&+\end{array}\right].

Suppose that RR has 55 negatives. Observe that RR cannot have exactly 1 negative per row, as this would contradict SS being row PPO. Further, at most 1 row of RR has 2 negatives, otherwise we have 2 positive rows, which violates row PPO. By these considerations, RR is sign equivalent to

[+++++++++++++++]\left[\begin{array}[]{cccc}-&-&+&+\\ +&-&+&+\\ +&+&-&+\\ +&+&+&-\\ +&+&+&+\\ \end{array}\right]

and we assume RR has this form. As before, SS can be obtained from RR by duplicating at least 2 distinct columns of RR. Observe that duplicating only columns 1 and 2 of RR 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, SS contains one of

S2=[+++++++++++++++++++++++],S3=[+++++++++++++++++++++++]S_{2}=\left[\begin{array}[]{cccccc}-&-&-&+&+&+\\ +&+&-&+&+&+\\ +&+&+&-&-&+\\ +&+&+&+&+&-\\ +&+&+&+&+&+\end{array}\right],\quad S_{3}=\left[\begin{array}[]{cccccc}-&-&+&+&+&+\\ +&-&+&+&+&+\\ +&+&-&-&+&+\\ +&+&+&+&-&-\\ +&+&+&+&+&+\end{array}\right]

or

S4=[++++++++++++++++++++++]S_{4}=\left[\begin{array}[]{cccccc}-&-&-&+&+&+\\ +&-&-&+&+&+\\ +&+&+&-&-&+\\ +&+&+&+&+&-\\ +&+&+&+&+&+\end{array}\right]

as a submatrix. Observe that S4S_{4} is sign equivalent to S2S_{2} (negate rows 1,21,2 and 55, and negate columns 4,54,5 and 66; then appropriately permute rows and columns). By Examples 2.9 and 3.37, S1,S2S_{1},S_{2} and S3S_{3} allow row orthogonality. Since SS minimally allows orthogonality, SS is sign equivalent to S1,S2,S_{1},S_{2}, or S3S_{3}.

Remark 3.40.

Characterizing all 6×n6\times n 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 6×86\times 8 sign pattern

S=[++++++++++++++++++++++++++++++++++].S=\left[\begin{array}[]{cccccccc}+&+&+&+&+&+&+&+\\ +&+&+&-&-&-&+&+\\ +&+&+&+&+&+&-&+\\ +&+&+&-&-&-&-&-\\ +&+&+&+&+&+&+&-\\ +&+&+&-&-&-&+&-\end{array}\right].

Deleting any number of columns will contradict Theorem 1.6, so if SS 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 SS. It would be very interesting if this sign pattern does not allow row orthogonality. It is not too difficult to verify that SS 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 [+]\left[\begin{array}[]{c}+\end{array}\right]
2 [+++]\left[\begin{array}[]{cc}+&-\\ +&+\end{array}\right]
3 [+++++++]\left[\begin{array}[]{ccc}+&-&+\\ +&+&-\\ +&+&+\end{array}\right]
4 [+++++++++++++]\left[\begin{array}[]{cccc}+&-&+&+\\ +&+&-&+\\ +&+&+&-\\ +&+&+&+\end{array}\right], [++++++++++++]\left[\begin{array}[]{cccc}+&-&-&+\\ +&+&-&+\\ +&+&+&-\\ +&+&+&+\end{array}\right], [++++++++++++]\left[\begin{array}[]{cccc}-&+&+&+\\ +&-&+&+\\ +&+&-&+\\ +&+&+&-\end{array}\right], [+++++++++++++++]\left[\begin{array}[]{ccccc}-&+&+&+&+\\ +&-&+&-&+\\ +&+&-&+&-\\ +&+&+&+&+\end{array}\right]
[++++++++++++++++++]\left[\begin{array}[]{rrrrr}+&-&-&+&+\\ +&+&-&-&+\\ +&+&+&-&-\\ +&+&+&+&-\\ +&+&+&+&+\end{array}\right], [++++++++++++++++++]\left[\begin{array}[]{rrrrr}-&-&+&+&+\\ +&-&-&+&+\\ +&+&-&-&+\\ +&+&+&+&-\\ +&+&+&+&+\end{array}\right], [++++++++++++++++++]\left[\begin{array}[]{rrrrr}-&-&+&+&+\\ +&-&-&+&+\\ +&+&-&+&+\\ +&+&+&-&+\\ +&+&+&+&-\end{array}\right],
5 [++++++++++++++++++]\left[\begin{array}[]{rrrrr}-&-&+&+&+\\ +&-&-&+&+\\ +&+&+&-&-\\ +&+&+&+&-\\ +&+&+&+&+\end{array}\right], [+++++++++++++++++++]\left[\begin{array}[]{rrrrr}-&-&+&+&+\\ +&-&-&+&+\\ +&+&+&-&+\\ +&+&+&+&-\\ +&+&+&+&+\end{array}\right], [+++++++++++++++++++]\left[\begin{array}[]{rrrrr}+&-&-&+&+\\ +&+&-&+&+\\ +&+&+&-&-\\ +&+&+&+&-\\ +&+&+&+&+\end{array}\right],
[+++++++++++++++++++]\left[\begin{array}[]{rrrrr}-&-&+&+&+\\ +&-&+&+&+\\ +&+&-&+&+\\ +&+&+&-&+\\ +&+&+&+&-\end{array}\right], [++++++++++++++++++++]\left[\begin{array}[]{rrrrr}-&+&+&+&+\\ +&-&+&+&+\\ +&+&-&+&+\\ +&+&+&-&+\\ +&+&+&+&-\end{array}\right], [+++++++++++++++++++++]\left[\begin{array}[]{rrrrr}+&-&+&+&+\\ +&+&-&+&+\\ +&+&+&-&+\\ +&+&+&+&-\\ +&+&+&+&+\end{array}\right],
[++++++++++++++++++++++++]\left[\begin{array}[]{cccccc}-&-&+&+&+&+\\ +&+&-&-&+&+\\ +&+&+&+&-&+\\ +&+&+&+&+&-\\ +&+&+&+&+&+\end{array}\right],  [+++++++++++++++++++++++]\left[\begin{array}[]{cccccc}-&-&-&+&+&+\\ +&+&-&+&+&+\\ +&+&+&-&-&+\\ +&+&+&+&+&-\\ +&+&+&+&+&+\end{array}\right], [+++++++++++++++++++++++]\left[\begin{array}[]{cccccc}-&-&+&+&+&+\\ +&-&+&+&+&+\\ +&+&-&-&+&+\\ +&+&+&+&-&-\\ +&+&+&+&+&+\end{array}\right]
Table 3.1: One representative of each sign-equivalence class of m×nm\times n nowhere zero sign patterns that minimally allow orthogonality for m5m\leq 5

4 Likelihood a random sign pattern allows row orthogonality

The question of finding the probability that mm vectors sampled from {±1}n\{\pm 1\}^{n} 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 {±1}m×n\{\pm 1\}^{m\times n} has rank mm (the literature is most interested in the case when m=nm=n). In particular, Tikhomirov [12] answered this question in a strong form by showing that this probability is bounded below by 1(12+o(1))m1-\bigl{(}{1\over 2}+o(1)\bigr{)}^{m} whenever nmn\geq m; in particular, when nmn\geq m the probability tends toward 11 as mm tends toward \infty.

In this section, we consider the adjacent problem of determining the threshold t(m)t(m) such that a random matrix in {+,}m×n\{+,-\}^{m\times n} with nt(m)n\geq t(m) allows row orthogonality with probability tending toward 11 as mm tends toward \infty.

Let f(n)f(n) and g(n)g(n) be functions from the non-negative integers to the reals. Then f(n)=o(g(n))f(n)=o(g(n)) if limnf(n)/g(n)=0\lim_{n\to\infty}{f(n)}/{g(n)}=0, and f(n)=ω(g(n))f(n)=\omega(g(n)) if g(n)=o(f(n))g(n)=o(f(n)). An event E=E(n)E=E(n) happens with high probability as nn\to\infty if Pr[E]=1o(1)\Pr[E]=1-o(1). 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 μ\mu on a set Ω\Omega, we write xμx\sim\mu to mean that xx is distributed according to μ\mu. If Ω\Omega is a finite set, then we write xΩx\sim\Omega to mean that xx is chosen uniformly from Ω\Omega. We write x1,,xnμx_{1},\dots,x_{n}\sim\mu to indicate that x1,,xnx_{1},\dots,x_{n} are distributed according to μ\mu and are mutually independent from one another. For a positive integer nn, μn\mu^{n} denotes the product distribution on Ωn\Omega^{n} where each entry is drawn independently from μ\mu. Similarly, for positive integers mm and nn, μm×n\mu^{m\times n} denotes the product distribution on Ωm×n\Omega^{m\times n} where each entry is drawn independently from μ\mu. For an index set α\alpha, let Ωα\Omega^{\alpha} denote the set of vectors with entries in Ω\Omega indexed by α\alpha. We write μα\mu^{\alpha} to mean the product distribution on Ωα\Omega^{\alpha} where each entry is drawn independently from μ\mu.

We will need two forms of the Chernoff bound, which we state here.

Theorem 4.1.

[1, Corollary A.1.2] Let XiX_{i}, 1in1\leq i\leq n, be mutually independent random variables with Pr[Xi=1]=Pr[Xi=1]=12\Pr[X_{i}=1]=\Pr[X_{i}=-1]=\frac{1}{2} and X=X1++Xn.X=X_{1}+\cdots+X_{n}. Let a>0a>0. Then

Pr[|X|>a]<2ea2/2n.\Pr[|X|>a]<2e^{-a^{2}/2n}.

Theorem 4.2.

[11, Remark 9.2] Suppose X1,,XnX_{1},\ldots,X_{n} are independent random variables taking values from the set {0,1}\{0,1\}. Let X=X1++XnX=X_{1}+\ldots+X_{n}. Then for any δ0\delta\geq 0

Pr[X(1δ)𝔼[X]]exp(δ2𝔼[X]/2).\Pr[X\leq(1-\delta)\mathbb{E}[X]]\leq\exp(-\delta^{2}\mathbb{E}[X]/2).

For a fixed 0p1/20\leq p\leq 1/2, let μp\mu_{p} denote the distribution on {0,±1}\{0,\pm 1\} where μp(1)=μp(1)=p\mu_{p}(1)=\mu_{p}(-1)=p and μp(0)=12p\mu_{p}(0)=1-2p. The main result of this section, Theorem 4.12, implies that for any fixed 0<p1/20<p\leq 1/2, there is a constant C=C(p)C=C(p) such that if Aμpm×nA\sim{\mu_{p}}^{m\times n} where nm2+Cmlogmn\geq m^{2}+Cm\log m, then sgn(A)\operatorname{sgn}(A) allows row orthogonality with high probability as mm\to\infty. 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 A{±1}m×nA\sim\{\pm 1\}^{m\times n} with n17m2logmn\geq 17m^{2}\log m, then sgn(A)\operatorname{sgn}(A) allows row orthogonality with high probability as mm\to\infty.

Proof 4.4.

Let 𝐱i{\bf x}_{i} denote the iith row of AA, so 𝐱1,,𝐱m{±1}n{\bf x}_{1},\dots,{\bf x}_{m}\sim\{\pm 1\}^{n}. Observe that 𝐱i2=n\lVert{\bf x}_{i}\rVert_{2}=\sqrt{n} and that δ(𝐱i)=1\delta({\bf x}_{i})=1 for each i[m]i\in[m]. Set

ϵ=174logmn,\epsilon=\sqrt{{\frac{17}{4}\log m\over n}},

and observe that 0ϵ12m0\leq\epsilon\leq{1\over 2m} since n17m2logmn\geq 17m^{2}\log m and so

rm(ϵ)=1+ϵ(1(m2)ϵ)(1(m1)ϵ)11+12m(1m22m)(1m12m)1=2m(m+1/2)(m+2)(m+1)1<1.r_{m}(\epsilon)=\sqrt{{1+\epsilon\over(1-(m-2)\epsilon)(1-(m-1)\epsilon)}}-1\leq\sqrt{{1+{1\over 2m}\over\bigl{(}1-{m-2\over 2m}\bigr{)}\bigl{(}1-{m-1\over 2m}\bigr{)}}}-1=2\sqrt{{m(m+1/2)\over(m+2)(m+1)}}-1<1.

Thus, if |𝐱i,𝐱j|ϵn\lvert\langle{\bf x}_{i},{\bf x}_{j}\rangle\rvert\leq\epsilon n for all ij[m]i\neq j\in[m], then we may apply Theorem 2.7 to locate a set of orthogonal vectors 𝐱~1,,𝐱~m\widetilde{{\bf x}}_{1},\dots,\widetilde{{\bf x}}_{m} such that sgn(𝐱~i)=sgn(𝐱i)\operatorname{sgn}(\widetilde{{\bf x}}_{i})=\operatorname{sgn}({\bf x}_{i}). Thus, in order to conclude the proof, it suffices to show that |𝐱i,𝐱j|ϵn\lvert\langle{\bf x}_{i},{\bf x}_{j}\rangle\rvert\leq\epsilon n for all ij[m]i\neq j\in[m] with high probability as mm\to\infty.

Since 𝐱1,,𝐱m{±1}n{\bf x}_{1},\dots,{\bf x}_{m}\sim\{\pm 1\}^{n} are independent, we may apply the Chernoff bound in Theorem 4.1 to bound

Pr[|𝐱i,𝐱j|>ϵn]<2eϵ2n/2\Pr\bigl{[}\lvert\langle{\bf x}_{i},{\bf x}_{j}\rangle\rvert>\epsilon n\bigr{]}<2e^{-\epsilon^{2}n/2}

for any ij{1,,m}i\neq j\in\{1,\dots,m\}. By the union bound,

Pr[|𝐱i,𝐱j|ϵn,ij[m]]\displaystyle\Pr\bigl{[}\lvert\langle{\bf x}_{i},{\bf x}_{j}\rangle\rvert\leq\epsilon n,\ \forall i\neq j\in[m]\bigr{]} 1i<j[m]Pr[|𝐱i,𝐱j|>ϵn]1(m2)2eϵ2n/2\displaystyle\geq 1-\sum_{i<j\in[m]}\Pr\bigl{[}\lvert\langle{\bf x}_{i},{\bf x}_{j}\rangle\rvert>\epsilon n\bigr{]}\geq 1-\binom{m}{2}2e^{-\epsilon^{2}n/2}
1m2eϵ2n/2=1m1/8=1o(1).\displaystyle\geq 1-m^{2}e^{-\epsilon^{2}n/2}=1-m^{-1/8}=1-o(1).

We now show how to improve Theorem 4.3 by using the SIPP. Recall that Theorem 3.22 states that the sign pattern S(Km,Km)S(K_{m},\vec{K}_{m}) requires o-SIPP. We say that a pair of negative 44-cycles are column-disjoint if the column indices of the negative 44-cycles are all distinct. Observe that any sign pattern that has a collection of column-disjoint negative 44-cycles between every pair of rows is sign equivalent to a superpattern of [S(Km,Km)O]\left[\begin{array}[]{c|c}S(K_{m},\vec{K}_{m})&O\end{array}\right]. So by Theorem 1.1 and Theorem 1.3, we have the following observation.

Observation 4.5.

Let SS be an m×nm\times n sign pattern. If SS has a collection of column-disjoint negative 4-cycles between every pair of rows, then SS 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 0<p1/20<p\leq 1/2. If 𝐱1,,𝐱mμpn{\bf x}_{1},\dots,{\bf x}_{m}\sim\mu_{p}^{n}, then the probability that we can find distinct integers i1,j1,i2,j2,,im,jm[n]i_{1},j_{1},i_{2},j_{2},\dots,i_{m},j_{m}\in[n] such that (𝐱k)ik=1({\bf x}_{k})_{i_{k}}=1 and (𝐱k)jk=1({\bf x}_{k})_{j_{k}}=-1 for all k[m]k\in[m] is at least

11p(1p)n2m+11-{1\over p}(1-p)^{n-2m+1}

Proof 4.7.

We employ the following greedy algorithm to find the required set Wm={i1,j1,,im,jm}W_{m}=\{i_{1},j_{1},\dots,i_{m},j_{m}\} of indices:

Initialize U0=[n]U_{0}=[n] and W0=W_{0}=\emptyset. At time t=1,,mt=1,\dots,m, do the following:

  1. 1.

    Reveal 𝐱t{\bf x}_{t}.

  2. 2.

    Attempt to locate some it,jtUt1i_{t},j_{t}\in U_{t-1} for which (𝐱t)it=1({\bf x}_{t})_{i_{t}}=1 and (𝐱t)jt=1({\bf x}_{t})_{j_{t}}=-1. If such it,jti_{t},j_{t} are found, then set Ut=Ut1{it,jt}U_{t}=U_{t-1}\setminus\{i_{t},j_{t}\} and Wt=Wt1{it,jt}W_{t}=W_{t-1}\cup\{i_{t},j_{t}\}. If no such it,jti_{t},j_{t} exist, then exit with failure.

If the above algorithm succeeds, then we have located the desired WmW_{m}.

Let τ\tau be the round on which the algorithm fails, setting τ=m+1\tau=m+1 if the algorithm succeeds. In order to complete the proof, we show that

Pr[τm]1p(1p)n2m+1.\Pr[\tau\leq m]\leq{1\over p}(1-p)^{n-2m+1}.

Fix any t[m]t\in[m] and consider conditioning on the event {τt}\{\tau\geq t\}. Since τt\tau\geq t if and only if the algorithm has succeeded locating the set Ut1U_{t-1}, we may condition on such an outcome. Now, conditioned on the algorithm locating Ut1U_{t-1}, we observe that τ=t\tau=t if and only if 𝐱t[Ut1]{\bf x}_{t}[U_{t-1}] is nonnegative or nonpositive. Furthermore, before the ttth loop, no information is known about the vector 𝐱t{\bf x}_{t} and so 𝐱t[Ut1]μpUt1{\bf x}_{t}[U_{t-1}]\sim\mu_{p}{}^{U_{t-1}}. We may therefore bound

Pr[τ=t|Ut1]=Pr[𝐱t[Ut1]{0,1}Ut1{0,1}Ut1|Ut1]2(1p)|Ut1|=2(1p)n2(t1).\Pr[\tau=t\ |\ U_{t-1}]=\Pr\bigl{[}{\bf x}_{t}[U_{t-1}]\in\{0,1\}^{U_{t-1}}\cup\{0,-1\}^{U_{t-1}}\ \bigm{|}\ U_{t-1}\bigr{]}\leq 2(1-p)^{\lvert U_{t-1}\rvert}=2(1-p)^{n-2(t-1)}.

Since this bound is independent of Ut1U_{t-1}, we may bound

Pr[τ=t]Pr[τ=t|τt]2(1p)n2(t1).\Pr[\tau=t]\leq\Pr[\tau=t\ |\ \tau\geq t]\leq 2(1-p)^{n-2(t-1)}.

We therefore conclude that

Pr[τm]\displaystyle\Pr[\tau\leq m] =\displaystyle= t=1mPr[τ=m]t=1m2(1p)n2(t1)=2k=0m1(1p)n2m+2+2k\displaystyle\sum_{t=1}^{m}\Pr[\tau=m]\leq\sum_{t=1}^{m}2(1-p)^{n-2(t-1)}=2\sum_{k=0}^{m-1}(1-p)^{n-2m+2+2k}
=\displaystyle= 2(1p)n2m+2(1p)n+2(2p)p2(2p)p(1p)n2m+21p(1p)n2m+1,\displaystyle 2{(1-p)^{n-2m+2}-(1-p)^{n+2}\over(2-p)p}\leq{2\over(2-p)p}(1-p)^{n-2m+2}\leq{1\over p}(1-p)^{n-2m+1},

where the final inequality follows from the fact that 1p2p12{1-p\over 2-p}\leq{1\over 2}.

We now use the above lemma to locate collections of column-disjoint negative 44-cycles.

Lemma 4.8.

Fix any 𝐱{±1}n{\bf x}\in\{\pm 1\}^{n} and any 0<p1/20<p\leq 1/2. Assume Aμpm×nA\sim\mu_{p}^{m\times n} and set B=[𝐱TA].B=\left[\begin{array}[]{c}{\bf x}^{T}\\ \hline\cr A\end{array}\right]. Then sgn(B)\operatorname{sgn}(B) contains column-disjoint negative 44-cycles between its first row and all other rows with probability at least

11p(1p)n2m+1.1-{1\over p}(1-p)^{n-2m+1}.

Proof 4.9.

Let DD be the diagonal matrix whose iith diagonal entry is the iith entry of 𝐱{\bf x}. Observe that sgn(B)\operatorname{sgn}(B) is sign equivalent to sgn(BD)=sgn([11AD])\operatorname{sgn}(BD)=\operatorname{sgn}\left(\left[\begin{array}[]{ccc}1&\cdots&1\\ \hline\cr&AD&\end{array}\right]\right). Since μp(1)=μp(1)\mu_{p}(1)=\mu_{p}(-1) and 𝐱{±1}n{\bf x}\in\{\pm 1\}^{n}, ADμpm×nAD\sim\mu_{p}^{m\times n}. Thus sgn(B)\operatorname{sgn}(B) contains column-disjoint negative 44-cycles between its first row and all other rows if and only if we can locate distinct i1,,im,j1,,jm[n]i_{1},\dots,i_{m},j_{1},\dots,j_{m}\in[n] so that (𝐰k)ik=1({\bf w}_{k})_{i_{k}}=1 and (𝐰k)jk=1({\bf w}_{k})_{j_{k}}=-1. As such, the conclusion follows from Lemma 4.6.

Lemma 4.10.

Fix a number 0<p1/20<p\leq 1/2. Assume Aμpm×nA\sim\mu_{p}^{m\times n}, where nm2+mr+2mpn\geq m^{2}+mr+{2m\over p} for some r{0,,m}r\in\{0,\dots,m\}. Then the probability that sgn(A)\operatorname{sgn}(A) contains a collection of column-disjoint negative 44-cycles between every pair of rows is bounded below by

1mem/8mp(1p)r.1-me^{-m/8}-{m\over p}(1-p)^{r}.

Proof 4.11.

In order to locate a collection of negative 44-cycles between every pair of rows of sgn(A)\operatorname{sgn}(A), we employ the following greedy algorithm:

Suppose that the rows of AA are 𝐱1,,𝐱m{\bf x}_{1},\dots,{\bf x}_{m}. Initialize U0=[n]U_{0}=[n]. At time t=1,,m1t=1,\dots,m-1 do the following:

  1. 1.

    Reveal 𝐱t[Ut1]{\bf x}_{t}[U_{t-1}].

  2. 2.

    Find some Wtsupp(𝐱i[Ut1])W_{t}\subseteq\operatorname*{supp}({\bf x}_{i}[U_{t-1}]) with |Wt|=2(mt)+r\lvert W_{t}\rvert=2(m-t)+r and set Ut=Ut1WtU_{t}=U_{t-1}\setminus W_{t}. If no such WtW_{t} exists, then fail.

  3. 3.

    Reveal A[{t+1,,m},Wt]A[\{t+1,\dots,m\},W_{t}]

  4. 4.

    Locate column-disjoint negative 44-cycles in sgn(A)\operatorname{sgn}(A) between row tt and all rows k>tk>t, all of whose columns reside within WtW_{t}. If such negative 44-cycles cannot be found, then fail.

If the above algorithm succeeds, then sgn(A)\operatorname{sgn}(A) contains a collection of column-disjoint negative 44-cycles between every pair of rows.

Let τ\tau denote the first round on which the algorithm fails, setting τ=m\tau=m if the algorithm succeeds. In order to prove the claim, we argue that

Pr[τm1]mem/8+mp(1p)r.\Pr[\tau\leq m-1]\leq me^{-m/8}+{m\over p}(1-p)^{r}.

Fix any t[m1]t\in[m-1]. Let 𝒮t\mathcal{S}_{t} denote the event that the algorithm fails at step 2 on the ttth loop, and let t\mathcal{F}_{t} denote the event that the algorithm fails at step 4 on the ttth loop. Certainly {τ=t}=𝒮tt\{\tau=t\}=\mathcal{S}_{t}\cup\mathcal{F}_{t}. We begin by bounding the probability of 𝒮t\mathcal{S}_{t}.

Consider conditioning on the event {τt}\{\tau\geq t\}. Of course, if τt\tau\geq t, then the algorithm has succeeded in locating the set Ut1U_{t-1}. Furthermore, conditioned on {τt}\{\tau\geq t\} and Ut1U_{t-1}, observe that prior to the ttth loop of the algorithm, no entries within A[{t,,m},Ut1]A[\{t,\dots,m\},U_{t-1}] have been revealed; therefore A[{t,,m},Ut1]μp{t,,m}×Ut1A[\{t,\dots,m\},U_{t-1}]\sim\mu_{p}^{\{t,\dots,m\}\times U_{t-1}}. In particular, 𝐱t[Ut1]μpUt1{\bf x}_{t}[U_{t-1}]\sim\mu_{p}^{U_{t-1}}. Now, WtW_{t} cannot be located if and only if |supp(𝐱t[Ut1])|<2(mt)+r\lvert\operatorname*{supp}({\bf x}_{t}[U_{t-1}])\rvert<2(m-t)+r. Additionally,

|Ut1|\displaystyle\lvert U_{t-1}\rvert =\displaystyle= nj=1t1(2(mj)+r)m2+rm+2mpj=1t1(2(mj)+r)\displaystyle n-\sum_{j=1}^{t-1}\bigl{(}2(m-j)+r\bigr{)}\geq m^{2}+rm+{2m\over p}-\sum_{j=1}^{t-1}\bigl{(}2(m-j)+r\bigr{)}
=\displaystyle= 2mp+m2+rm(t1)(2mt)(t1)r2mp.\displaystyle{2m\over p}+m^{2}+rm-(t-1)(2m-t)-(t-1)r\geq{2m\over p}.

We can therefore fix a subset UUt1U\subseteq U_{t-1} with |U|=2mp\lvert U\rvert={\left\lfloor{2m\over p}\right\rfloor}. From the earlier observation, we know that 𝐱t[U]μpU{\bf x}_{t}[U]\sim\mu_{p}^{U} and so we may bound

Pr[𝒮t|{τt},Ut1]\displaystyle\Pr[\mathcal{S}_{t}\ |\ \{\tau\geq t\},U_{t-1}] =\displaystyle= Pr[|supp(𝐱t[Ut1])|<2(mt)+r|{τt},Ut1]\displaystyle\Pr\bigl{[}\lvert\operatorname*{supp}({\bf x}_{t}[U_{t-1}])\rvert<2(m-t)+r\ \bigm{|}\ \{\tau\geq t\},U_{t-1}\bigr{]}
\displaystyle\leq Pr[|supp(𝐱t[U])|<2(mt)+r|{τt},Ut1]\displaystyle\Pr\bigl{[}\lvert\operatorname*{supp}({\bf x}_{t}[U])\rvert<2(m-t)+r\ \bigm{|}\ \{\tau\geq t\},U_{t-1}\bigr{]}
=\displaystyle= Pr𝐱μp2m/p[|supp𝐱|<2(mt)+r]\displaystyle\Pr_{{\bf x}\sim\mu_{p}^{\lfloor 2m/p\rfloor}}\bigl{[}\lvert\operatorname*{supp}{\bf x}\rvert<2(m-t)+r\bigr{]}
\displaystyle\leq Pr𝐱μp2m/p[|supp𝐱|<3m1],\displaystyle\Pr_{{\bf x}\sim\mu_{p}^{\lfloor 2m/p\rfloor}}\bigl{[}\lvert\operatorname*{supp}{\bf x}\rvert<3m-1\bigr{]},

where the final inequality follows from the fact that t1t\geq 1 and rmr\leq m.

Next, we note that if 𝐱μp2m/p{\bf x}\sim\mu_{p}^{\lfloor 2m/p\rfloor}, then 𝔼[|supp𝐱|]=2p2mp4m2p4m1{\mathbb{E}[\lvert\operatorname*{supp}{\bf x}\rvert]}=2p\left\lfloor{2m\over p}\right\rfloor\geq 4m-2p\geq 4m-1 and so

Pr𝐱μp2m/p[|supp𝐱|<3m1]\displaystyle\Pr_{{\bf x}\sim\mu_{p}^{\left\lfloor 2m/p\right\rfloor}}\bigl{[}\lvert\operatorname*{supp}{\bf x}\rvert<3m-1\bigr{]} \displaystyle\leq Pr𝐱μp2m/p[|supp𝐱|<𝔼[|supp𝐱|]m]\displaystyle\Pr_{{\bf x}\sim\mu_{p}^{\left\lfloor 2m/p\right\rfloor}}\bigl{[}\lvert\operatorname*{supp}{\bf x}\rvert<{\mathbb{E}[\lvert\operatorname*{supp}{\bf x}\rvert]}-m\bigr{]}
=\displaystyle= Pr𝐱μp2m/p[|supp𝐱|<(1m𝔼[|supp𝐱|])𝔼[|supp𝐱|]]\displaystyle\Pr_{{\bf x}\sim\mu_{p}^{\left\lfloor 2m/p\right\rfloor}}\left[\lvert\operatorname*{supp}{\bf x}\rvert<\left(1-\frac{m}{\mathbb{E}[\lvert\operatorname*{supp}{\bf x}\rvert]}\right)\mathbb{E}[\lvert\operatorname*{supp}{\bf x}\rvert]\right]
\displaystyle\leq exp(m22𝔼[|supp𝐱|])=exp(m24p2mp)em/8,\displaystyle\exp\biggl{(}-\frac{m^{2}}{2\mathbb{E}[|\operatorname*{supp}{\bf x}|]}\biggr{)}=\exp\biggl{(}-{m^{2}\over 4p\left\lfloor{2m\over p}\right\rfloor}\biggr{)}\leq{e^{-m/8}},

where the second inequality follows from the Chernoff bound in Theorem 4.2. Since this bound is independent of Ut1U_{t-1}, we have argued that

(4.4) Pr[𝒮t|τt]em/8.\Pr[\mathcal{S}_{t}\ |\ \tau\geq t]\leq e^{-m/8}.

Next, we bound the probability of t\mathcal{F}_{t}. In order for t\mathcal{F}_{t} to hold, it must be the case that τt\tau\geq t and that 𝒮t\mathcal{S}_{t} does not hold; in particular, the algorithm must have succeeded in locating the set WtW_{t}. By construction, just after locating WtW_{t}, no entries within A[{t+1,,m},Wt]A[\{t+1,\dots,m\},W_{t}] have been revealed; therefore A[{t+1,,m},Wt]μp{t+1,,m}×WtA[\{t+1,\dots,m\},W_{t}]\sim\mu_{p}^{\{t+1,\dots,m\}\times W_{t}}. Since Pr[t|Wt]\Pr[\mathcal{F}_{t}\ |\ W_{t}] is equal to the probability of not finding a collection of column disjoint negative 4-cycles between the first row of A[{t+1,,m},Wt]A[\{t+1,\dots,m\},W_{t}] and the remaining rows, we may appeal to Lemma 4.8 to bound

Pr[t|Wt]1p(1p)|Wt|2(mt)+11p(1p)r.\Pr[\mathcal{F}_{t}\ |\ W_{t}]\leq{1\over p}(1-p)^{\lvert W_{t}\rvert-2(m-t)+1}\leq{1\over p}(1-p)^{r}.

Since this bound is independent of WtW_{t}, we have shown that

(4.5) Pr[t|{τt},𝒮t¯]1p(1p)r,\Pr[\mathcal{F}_{t}\ |\ \{\tau\geq t\},\overline{\mathcal{S}_{t}}]\leq{1\over p}(1-p)^{r},

where 𝒮t¯\overline{\mathcal{S}_{t}} denotes the event that 𝒮t\mathcal{S}_{t} does not occur.

Combining (4.4) and (4.5) we have shown that

Pr[τ=t]\displaystyle\Pr[\tau=t] \displaystyle\leq Pr[τ=t|τt]=Pr[𝒮t|τt]+Pr[t|τt]\displaystyle\Pr[\tau=t\ |\ \tau\geq t]=\Pr[\mathcal{S}_{t}\ |\ \tau\geq t]+\Pr[\mathcal{F}_{t}\ |\ \tau\geq t]
\displaystyle\leq Pr[𝒮t|τt]+Pr[t|{τt},𝒮t¯]em/8+1p(1p)r,\displaystyle\Pr[\mathcal{S}_{t}\ |\ \tau\geq t]+\Pr[\mathcal{F}_{t}\ |\ \{\tau\geq t\},\overline{\mathcal{S}_{t}}]\leq e^{-m/8}+{1\over p}(1-p)^{r},

where the first equality holds since 𝒮t\mathcal{S}_{t} and t\mathcal{F}_{t} partition {τ=t}\{\tau=t\}.

Using this inequality, we finally bound

Pr[τm1]=t=1m1Pr[τ=t]t=1m1(em/8+1p(1p)r)mem/8+mp(1p)r,\Pr[\tau\leq m-1]=\sum_{t=1}^{m-1}\Pr[\tau=t]\leq\sum_{t=1}^{m-1}\left(e^{-m/8}+{1\over p}(1-p)^{r}\right)\leq me^{-m/8}+{m\over p}(1-p)^{r},\vspace{-5pt}

as needed.

Theorem 4.12.

For any fixed 0<p1/20<p\leq 1/2, if Aμpm×nA\sim\mu_{p}^{m\times n} and

nm2+mlog1/(1p)m+ω(m),n\geq m^{2}+m\log_{1/(1-p)}m+\omega(m),

then sgn(A)\operatorname{sgn}(A) allows row orthogonality with high probability as mm\to\infty.

Proof 4.13.

Suppose that

nm2+mlog1/(1p)m+f(m),n\geq m^{2}+m\log_{1/(1-p)}m+f(m),

where f(m)=ω(m)f(m)=\omega(m). Without loss of generality, we may additionally suppose that f(m)=o(m2)f(m)=o(m^{2}). Set r=log1/(1p)m+f(m)m2pr=\log_{1/(1-p)}m+\frac{f(m)}{m}-{2\over p} which is certainly bounded above by mm for all sufficiently large mm since f(m)=o(m2)f(m)=o(m^{2}). Furthermore, by decreasing f(m)f(m) by some amount no more than mm, we may ensure that rr is an integer, the lower bound on nn remains true, f(m)=ω(m)f(m)=\omega(m), and f(m)=o(m2)f(m)=o(m^{2}). Now, since f(m)=ω(m)f(m)=\omega(m) and 0<p1/20<p\leq 1/2 is fixed, we have that

nm2+mlog1/(1p)m+f(m)m2+mr+2mpn\geq m^{2}+m\log_{1/(1-p)}m+f(m)\geq m^{2}+mr+{2m\over p}

for all sufficiently large mm. Thus, we may apply Lemma 4.10 to learn that sgn(A)\operatorname{sgn}(A) contains a collection of column-disjoint negative 44-cycles between every pair of rows (and hence has a row orthogonal realization) with probability at least

1mem/8mp(1p)r=1mem/8mp(1p)log(1p)m(1p)f(m)m2p=1mem/81p(1p)f(m)m2p,1-me^{-m/8}-{m\over p}(1-p)^{r}=1-me^{-m/8}-{m\over p}(1-p)^{-\log_{(1-p)}m}(1-p)^{{f(m)\over m}-{2\over p}}=1-me^{-m/8}-{1\over p}(1-p)^{{f(m)\over m}-{2\over p}},

which tends to 11 as mm\to\infty since f(m)=ω(m)f(m)=\omega(m) and 0<p1/20<p\leq 1/2.

We suspect that Theorem 4.12 is not best possible.

Question 4.14.

Determine the threshold t(m)t(m) such that if S{+,}m×nS\sim\{+,-\}^{m\times n} with nt(m)n\geq t(m), then SS has a row orthogonal realization with high probability as mm\to\infty.

Theorem 4.12 implies that t(m)m2+mlog2m+ω(m)t(m)\leq m^{2}+m\log_{2}m+\omega(m). Observe that t(m)mt(m)\geq m 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 t(m)=mt(m)=m.

Theorem 4.15.

Let X{±1}m×mX\sim\{\pm 1\}^{m\times m}. Then with high probability as m{m}\to\infty the matrix XX does not contain an r×sr\times s submatrix YY such that r+s=m+2r+s={m}+2 and rankY=1\operatorname{rank}Y=1.

Proof 4.16.

Let Ω\Omega denote the set of pairs (𝐱,𝐲)({\bf x},{\bf y}), where 𝐱{±1}r{\bf x}\in\{\pm 1\}^{r}, 𝐲{±1}s{\bf y}\in\{\pm 1\}^{s} and the first entry of 𝐱{\bf x} is 11. Observe that the map (𝐱,𝐲)𝐱𝐲T({\bf x},{\bf y})\mapsto{\bf x}{\bf y}^{T} is a bijection between Ω\Omega and the set of rank 1 matrices in {±1}r×s\{\pm 1\}^{r\times s}. Thus the probability that Y{±1}r×sY\sim\{\pm 1\}^{r\times s} has rank 1 is precisely 2(r1)(s1)2^{-(r-1)(s-1)}.

The number of r×sr\times s submatrices of XX is (mr)(ms)\binom{m}{r}\binom{m}{s}. By the union bound, the probability that XX contains an r×sr\times s submatrix YY such that r+s=m+2r+s={m}+2 and rankY=1\operatorname{rank}Y=1 is at most

r=2m(mr)(mm+2r)2(r1)(m+1r)=k=1m1(mk+1)(mm+1k)2k(mk).\sum_{r=2}^{m}\binom{m}{r}\binom{m}{m+2-r}2^{-(r-1)(m+1-r)}=\sum_{k=1}^{m-1}\binom{m}{k+1}\binom{m}{m+1-k}2^{-k(m-k)}.

We show that this sum tends toward 0 as mm\to\infty by showing that

(mk+1)(mm+1k)<2k(mk)m2\binom{m}{k+1}\binom{m}{m+1-k}<\frac{2^{k(m-k)}}{m^{2}}

for all 1km11\leq k\leq m-1, provided mm is sufficiently large. If k2k\leq 2 or km2k\geq m-2, then for mm sufficiently large

(mk+1)(mm+1k)m4<2k(mk)m2.\binom{m}{k+1}\binom{m}{m+1-k}\leq m^{4}<\frac{2^{k(m-k)}}{m^{2}}.

For 3km33\leq k\leq m-3, we have

(mk+1)(mm+1k)22m<23(m3)m22k(mk)m2.\binom{m}{k+1}\binom{m}{m+1-k}\leq 2^{2m}<\frac{2^{3(m-3)}}{m^{2}}\leq\frac{2^{k(m-k)}}{m^{2}}.

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.