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

Characterization of the OU matrix of a braid diagram

Ayaka Shimizu and Yoshiro Yaguchi Osaka Central Advanced Mathematical Institute, Osaka Metropolitan University, Sugimoto, Osaka, 558-8585, Japan. Email: [email protected] Institute of Technology, Maebashi, Gunma, 371-0816, Japan. Email: [email protected]
Abstract

The OU matrix of a braid diagram is a square matrix that represents the number of over/under crossings of each pair of strands. In this paper, the OU matrix of a pure braid diagram is characterized for up to 5 strands. As an application, the crossing matrix of a positive pure braid is also characterized for up to 5 strands. Moreover, a standard form of the OU matrix is given and characterized for general braids of up to 5 strands.

1 Introduction

An nn-braid is an object in 3\mathbb{R}^{3} consisting of nn strands whose endpoints are fixed at horizontal bars at the top and bottom, where each strand runs from top to bottom without returning ([1]). We say that two nn-braids are equivalent (or same) if they are sent to each other by an ambient isotopy of 3{\mathbb{R}}^{3} fixing the top and bottom bars pointwise. In this paper, we call the equivalence class of a braid simply a braid. An nn-braid diagram is a regular projection of an nn-braid with horizontal bars in which each intersection point has the over/under information as shown in Figure 1.

Refer to caption
Figure 1: A 5-braid diagram BB and the OU matrix U(B)U(B).

Let s1,s2,,sns_{1},s_{2},\dots,s_{n} be the strands of an nn-braid diagram BB with subscripts from the left-hand side to the right-hand side on the top of the braid diagram as shown in Figure 1. The OU matrix, U(B)U(B), of BB is an n×nn\times n matrix with zero diagonal such that the (i,j)(i,j) entry is the number of crossings between sis_{i} and sjs_{j} where sis_{i} is over sjs_{j} as shown in Figure 1. (For more precise definition, see [9] or Section 2.) The OU matrix was introduced in [9] to discuss the “warping degree” of a braid diagram ([8]), which estimates the unknotting number of the closure of a braid diagram (see [6, 7] for the warping degree of a link diagram). It was also found that the determinant of the OU matrix acts effectively on layered braid diagrams ([9]). In this paper, we characterize the OU matrix of a pure braid diagram of up to 5×55\times 5.

We call a matrix whose entries are all integers (resp. non-negative integers) an integer matrix (resp. non-negative integer matrix). We call a matrix whose entries are all non-negative even integers a non-negative even matrix. We say that a square matrix MM is a zero-diagonal matrix when MM has zero diagonal, namely all the (i,i)(i,i) entries on the main diagonal are zero. We denote the (i,j)(i,j) entry of a matrix MM by M(i,j)M(i,j). The following properties of n×nn\times n zero-diagonal matrices, T0 and T1, are introduced in [2].

Definition 1 ([2]).

A zero-diagonal matrix MM is said to be T0 if whenever 1i<j<kn1\leq i<j<k\leq n, then M(i,j)=M(j,k)=0M(i,j)=M(j,k)=0 implies A(i,k)=0A(i,k)=0. A zero-diagonal matrix MM is said to be T1 if whenever 1i<j<kn1\leq i<j<k\leq n, then M(i,j),M(j,k)0M(i,j),~{}M(j,k)\neq 0 implies A(i,k)0A(i,k)\neq 0.

For example, the matrices

A=(0010000110010110),B=(0012001011012010)\displaystyle A=\begin{pmatrix}0&0&1&0\\ 0&0&0&1\\ 1&0&0&1\\ 0&1&1&0\end{pmatrix},\ B=\begin{pmatrix}0&0&1&2\\ 0&0&1&0\\ 1&1&0&1\\ 2&0&1&0\end{pmatrix}

are not T0 since they do not meet the condition with A(1,2)=A(2,3)=0A(1,2)=A(2,3)=0 and A(1,3)=1A(1,3)=1, B(1,2)=B(2,4)=0B(1,2)=B(2,4)=0 and B(1,4)=2B(1,4)=2. In this paper, when we say that a matrix MM is T0, it means that MM is a T0 zero-diagonal matrix. Let MTM^{T} denote the transpose of a matrix MM. In this paper, we prove the following theorem.

Theorem 1.

When n5n\leq 5, an n×nn\times n matrix MM is the OU matrix of some pure nn-braid diagram if and only if M+MTM+M^{T} is a non-negative even T0 matrix.

For an nn-braid diagram BB, the crossing matrix C(B)C(B) of BB is defined in [2] to be the n×nn\times n zero-diagonal matrix whose (i,j)(i,j) entry is the algebraic number (positive minus negative) of crossings of BB where the strand sis_{i} is over the strand sjs_{j} (see also [5]). If two braid diagrams BB and BB^{\prime} represent the same braid, then C(B)=C(B)C(B)=C(B^{\prime}) holds by definition, and therefore the crossing matrix is an invariant of equivalence classes of braids. This means that the crossing matrix C([B])C([B]) of the equivalence class [B][B] of BB is well defined. We denote C([B])C([B]) by C(B)C(B) for simplicity. In [2], it is proved that the crossing matrix can completely characterize positive braids of canonical length at most 2 (in terms of the canonical form as explained in [3, 10]), whereas it is also proved that it is impossible for length 3 or more.

As for the OU matrices, the equation U(B)=U(B)U(B)=U(B^{\prime}) does not hold for the braid diagrams BB and BB^{\prime} that represent the same braid because the OU matrix changes when two consecutive crossings of two strands are produced or reduced, that is, it changes by a Reidemeister move of type II. This implies that the OU matrix is not an invariant of equivalence classes of braids but an invariant of (planner isotopy class of) braid diagrams. For positive braids, it is shown in [9] that the OU matrix is an invariant by taking any positive diagram.

In this paper, we show that the equality C(B)=U(B)C(B)=U(B) holds for any positive braid diagram BB (Proposition 7) and hence for any positive braids as well. In [2], a conjecture is proposed about a necessary and sufficient condition for a matrix to be the crossing matrix C(B)C(B) of a positive pure braid BB, and it was proved that the conjecture is true when n3n\leq 3. In this paper, we show that the conjecture is true when n5n\leq 5 (Corollary 9) by discussing OU matrices. The conjecture is now open for n6n\geq 6, although an algorithm that exhibits all positive braids with a given crossing matrix, if any exist, or declares that there are none is known (see [5]).

The rest of the paper is organized as follows. In Section 2, we describe the definitions and properties of the OU matrix, CN matrix, and the crossing matrix. In Section 3, we investigate CN matrices whose entries are 0 or 2, introducing the BW-ladder diagram. In Section 4, we investigate the CN matrices of nn-pure braid diagrams for n5n\leq 5. In Section 5, we prove Theorem 1, namely, we characterize the OU matrices of pure braid diagrams up to 5×55\times 5. We also discuss the crossing matrices of positive pure braids and the OU matrices of braid diagrams that are not necessarily pure.

2 OU, CN and crossing matrices

This section includes an introductory overwiew of the OU, CN and crossing matrices. In Section 2.1, we review the definition of the OU matrix of a braid diagram which was defined in [9]. In Section 2.2, we define the CN matrix which will be used for the proof of the main theorem and see some properties. In Section 2.3, we review the definition of the crossing matrix of a braid which was defined in [2]. In Section 2.4, we discuss the relation between the OU matrices and the crossing matrices. In Section 2.5, we discuss matrices which realize such matrices.

2.1 OU matrix

In this subsection, we review the definition and some properties of the OU matrix defined in [9]. Let BB be an nn-braid diagram with the strands s1,s2,,sns_{1},s_{2},\dots,s_{n} that are ordered from the left-hand side to the right-hand side on the top of BB. Let π\pi be a permutation on (1,2,,n)(1,2,\dots,n). The OU matrix U(B,π)U(B,\pi) of BB with π\pi is an n×nn\times n zero-diagonal matrix MM such that M(i,j)M(i,j) is the number of crossings on the ithi^{th} strand that are over the jthj^{th} strand (iji\neq j), where the “kthk^{th} strand” stands for the strand sπ(k)s_{\pi(k)}, that is, the kthk^{th} strand according to the order π\pi. In this paper, we also denote U(B,id)U(B,id) simply by U(B)U(B) for the identity permutation idid. In Figure 2, the OU matrices of a braid diagram B1B_{1} with the permutations idid and π=(4,3,1,2)\pi=(4,3,1,2) are shown as U(B1)U(B_{1}) and U(B1,π)U(B_{1},\pi), respectively.

Refer to caption
Figure 2: Braid diagrams B1B_{1}, B2B_{2} and their OU, CN matrices.

For nn-braid diagrams BB and CC, we define the product BCBC by the nn-braid diagram obtained by placing BB above CC so that the endpoints of BB on the bottom bar coincide with those of CC on the top bar. We call this product the braid product. We have the following proposition.

Proposition 1 ([9]).

Let BB, CC be nn-braid diagrams, and let ρB\rho_{B} be the braid permutation of BB. Then U(BC,π)=U(B,π)+U(C,πρB)U(BC,\pi)=U(B,\pi)+U(C,\pi\rho_{B}) holds for any strand permutation π\pi. In particular, when BB is a pure braid, namely ρB=id\rho_{B}=id, we have U(BC,π)=U(B,π)+U(C,π)U(BC,\pi)=U(B,\pi)+U(C,\pi).

2.2 CN matrix

In this subsection, we define the CN matrix. Let BB be an nn-braid diagram with the strands s1,s2,,sns_{1},s_{2},\dots,s_{n} ordered from left to right on the top of BB. A crossing number matrix N(B)N(B), or CN matrix, of an nn-braid diagram BB is an n×nn\times n zero-diagonal matrix MM such that M(i,j)M(i,j) is the number of crossings between sis_{i} and sjs_{j} (iji\neq j). By definition, N(B)N(B) is a symmetric matrix, and N(B)=U(B)+(U(B))TN(B)=U(B)+\left(U(B)\right)^{T} holds for any BB (see Figure 2). From Proposition 1, we have the following proposition.

Proposition 2.

The addition N(BC)=N(B)+N(C)N(BC)=N(B)+N(C) holds when BB is a pure braid diagram.

We also have the following proposition for pure braid diagrams.

Proposition 3.

All the entries of the CN matrix N(B)N(B) of a braid diagram BB are even numbers if and only if BB is a pure braid diagram.

Proof.

A braid diagram BB has the identity braid permutation if and only if each pair of strands has the same relative position on the top and bottom, and that is equivalent to that each pair of strands has an even number of crossings between them. ∎

Definition 2.

For a square matrix MM, the reverse of MM, denoted by MM^{\prime}, is the matrix obtained from MM by reversing the order of the row and column.

Proposition 4.

If a matrix MM is the CN matrix of a braid diagram, then the reverse MM^{\prime} is also the CN matrix of a braid diagram.

Proof.

Let BB be an nn-braid diagram such that N(B)=MN(B)=M. Take the vertical rotation BB^{\prime}. Then, the order of the strands is reversed, and we have N(B)=MN(B^{\prime})=M^{\prime}. (See N(B1)N(B_{1}) and N(B2)N(B_{2}) in Figure 2, where B2=B1B_{2}=B_{1}^{\prime} and N(B2)=(N(B1))N(B_{2})=(N(B_{1}))^{\prime}.) ∎

Proposition 5.

If MM is the CN matrix of a braid diagram, then MM is T0.

Proof.

Let BB be a braid diagram such that N(B)=MN(B)=M. If M(i,j)0M(i,j)\neq 0 for a pair of ii and jj with i<ji<j, then BB has a crossing between the ithi^{th} and jthj^{th} strands by definition. Let cc be the upper-most crossing between the ithi^{th} and jthj^{th} strands. Let RR be the region bounded by the top horizontal bar, the ithi^{th} and jthj^{th} strands from the top to cc. Since RR is homeomorphic to the disc D2D^{2}, each kthk^{th} strand with i<k<ji<k<j has a positive even number of intersections with the boundary R\partial R, and exactly one of them is on the top bar and the others are on the ithi^{th} or jthj^{th} strands. Hence, the kthk^{th} strand has a crossing with the ithi^{th} or jthj^{th} strand, and we have M(i,k)>0M(i,k)>0 or M(k,j)>0M(k,j)>0. ∎

Since CN matrices are zero-diagonal symmetric matrices, we consider strictly upper triangular matrices instead of symmetric matrices in Sections 3 and 4 for simplicity.

2.3 Crossing matrix

As mentioned in Section 1, the crossing matrix is a matrix defined in [2] for braids. Let BB be an nn-braid diagram with strands s1,s2,,sns_{1},s_{2},\dots,s_{n} with subscripts from left to right at the top of the braid diagram. The crossing matrix C(B)C(B) of BB is an n×nn\times n zero-diagonal matrix such that each (i,j)(i,j) entry is the algebraic number of crossings, i.e., the number of positive crossings minus the number of negative crossings111The definition of positive/negative crossings in this paper is opposite to [2] and same to [5] or [9]., of BB where sis_{i} is over sjs_{j}. For a braid diagram BB, we denote [B][B] by the equivalent class of a braid that has BB as a diagram. We note that if two diagrams BB and BB^{\prime} represent the same braid in 3{\mathbb{R}}^{3}, i.e., [B]=[B][B]=[B^{\prime}], then C(B)=C(B)C(B)=C(B^{\prime}) (see [2, 5]). Therefore, for a braid b=[B]b=[B] that has a diagram BB, we can define the crossing matrix C(b)C(b) of bb by C(B)C(B). In [2], the crossing matrices of nn-braids and pure nn-braids are completely characterized for all nn\in\mathbb{N}. More specifically, it is proved in [2] that an n×nn\times n matrix MM is the crossing matrix of a pure braid if and only if MM is a zero-diagonal integer symmetric matrix. For positive pure braids, it is conjectured as follows.

Conjecture 1 ([2]).

An n×nn\times n matrix MM is the crossing matrix of some positive pure braid if and only if MM is a non-negative integer T0 symmetric matrix.

It has been shown that the conjecture is true when n3n\leq 3 in [2].

2.4 OU matrix and crossing matrix

In this subsection, we discuss the relation between the OU matrix and the crossing matrix. In [9], the following proposition was shown for positive pure braid diagrams.

Proposition 6 ([9]).

The OU matrix of a positive pure braid diagram is a symmetric matrix.

The OU matrix coincides with the crossing matrix for positive braid diagrams.

Proposition 7.

When BB is a positive braid diagram, the equality OU(B)=C(B)OU(B)=C(B) holds.

Proof.

Let kk be the number of crossings between a pair of strands sis_{i} and sjs_{j} where sis_{i} is over sjs_{j}. By the definition of the OU matrix, the (i,j)(i,j) enrty of the OU matrix U(B)U(B) is kk. The (i,j)(i,j) entry of the crossing matrix C(B)C(B) is also kk because all the crossings are positive crossings. ∎

Let n{\mathcal{B}}_{n} be the set of planar isotopy classes of nn-braid diagrams. Then, n{\mathcal{B}}_{n} forms a monoid under the braid product. Let 𝒫n{\mathcal{P}}_{n} be the submonoid of n{\mathcal{B}}_{n} which consists of planar isotopy classes of pure braid diagrams and let 𝒫n+{\mathcal{P}}_{n}^{+} be the submonoid of 𝒫n{\mathcal{P}}_{n} which consists of planar isotopy classes of positive pure braid diagrams. Let BnB_{n} be the nn-braid group, that is, the set of equivalent classes of nn-braids with the group operation naturally induced by the braid product. Let PnP_{n} be the subgroup of BnB_{n} which consists of pure nn-braids and let Pn+P_{n}^{+} be the monoid which consists of positive pure nn-braids.

Let XnX_{n} be the set of n×nn\times n zero-diagonal integer matrix and let Xn+X_{n}^{+} be the set of matrices MXnM\in X_{n} such that M(i,j)0M(i,j)\geq 0 for any i,ji,j. We note that (Xn,+)(X_{n},+) is a group whereas (Xn+,+)(X_{n}^{+},+) is a monoid. Let C:BnXnC:B_{n}\to X_{n} be the map defined by that C(b)=C(B)C(b)=C(B) is the crossing matrix of b=[B]Bnb=[B]\in B_{n}, let U:nXn+U:{\mathcal{B}}_{n}\to X_{n}^{+} be the map defined by that U(B)U(B) is the OU matrix of BnB\in{\mathcal{B}}_{n} and let N:nXn+N:{\mathcal{B}}_{n}\to X_{n}^{+} be the map defined by N(B)N(B) is the CN matrix of BnB\in{\mathcal{B}}_{n}.

For MXnM\in X_{n} and a permutation πSn=Sym{1,,n}\pi\in S_{n}=Sym\{1,\dots,n\}, let MπM^{\pi} be the n×nn\times n matrix obtained by rearranging the rows and columns of MM according to π\pi, i.e. the ithi^{th} row and column are π1(i)th\pi^{-1}(i)^{th} row and column of the original MM, respectively. We easily see that for MXnM\in X_{n} (resp. MXn+M\in X_{n}^{+}), it holds that MπXnM^{\pi}\in X_{n} (resp. MπXn+M^{\pi}\in X_{n}^{+}), and we obtain a group (resp. monoid) right action of SnS_{n} on XnX_{n}. Thus, we have semi-direct products XnSnX_{n}\rtimes S_{n} (as a group) and Xn+SnX_{n}^{+}\rtimes S_{n} (as a monoid) using the above right action of SnS_{n} on XnX_{n} and X+X^{+}.

For any two nn-braids aa and bBnb\in B_{n}, it holds that C(ab)=C(a)+C(b)πaC(ab)=C(a)+C(b)^{\pi_{a}} (see [2, 5]). Thus, the map C:BnXnC:B_{n}\to X_{n} induces a group homomorphism C~:BnXnSn\widetilde{C}:B_{n}\to X_{n}\rtimes S_{n} defined by C~(b)=(C(b),ρb)\widetilde{C}(b)=(C(b),\rho_{b}) for bBnb\in B_{n}, where ρbSn\rho_{b}\in S_{n} is the braid permutation of bb. As noted in [5], KerC~=[Pn,Pn]×{id}{\rm Ker}\ \widetilde{C}=[P_{n},P_{n}]\times\{id\}, where [Pn,Pn][P_{n},P_{n}] is the comutator subgroup of PnP_{n}. On the other hand, for two nn-braid diagrams A,BnA,B\in{\mathcal{B}}_{n}, it holds that U(AB)=U(A)+U(B)πAU(AB)=U(A)+U(B)^{\pi_{A}} (see [9]). Then for each braid diagram BB of bBnb\in B_{n} and each W{C,N}W\in\{C,N\}, the map W:nXn+W:{\mathcal{B}}_{n}\to X_{n}^{+} induces a monoid homomorphism W~:nXn+Sn\widetilde{W}:{\mathcal{B}}_{n}\to X_{n}^{+}\rtimes S_{n} defined by W~(B)=(W(B),ρb)\widetilde{W}(B)=(W(B),\rho_{b}). For each W{C,N}W\in\{C,N\}, it obviously holds that W~(B)=(O,id)\widetilde{W}(B)=(O,id) for BnB\in{\mathcal{B}}_{n}, where OO is the zero matrix, if and only if BB is a “trivial braid diagram”, that is, there are no crossings in BB.

2.5 Realizable matrices

We say that a matrix MM is OU-realizable (resp. CN-realizable) if there exists a braid diagram BB such that U(B)=MU(B)=M (resp. N(B)=MN(B)=M). We say that a matrix MM is C-realizable if there exists a braid bb such that C(b)=MC(b)=M. By Proposition 5, we have the following proposition.

Proposition 8.

A square matrix MM is not OU-realizable if M+MTM+M^{T} is not T0.

Example 1.

For

A=(0200100110000120),B=(0011000001001010),\displaystyle A=\begin{pmatrix}0&2&0&0\\ 1&0&0&1\\ 1&0&0&0\\ 0&1&2&0\end{pmatrix},\ B=\begin{pmatrix}0&0&1&1\\ 0&0&0&0\\ 0&1&0&0\\ 1&0&1&0\end{pmatrix},

the matrix AA is OU-realizable because A=U(B1)A=U(B_{1}) in Figure 2, whereas BB is not OU-realizable since B+BTB+B^{T} is not T0. In addition, AA and BB are not CN-realizable because they are not symmetric.

Proposition 9.

If an n×nn\times n symmetric matrix MM is CN-realizable, then any non-negative integer matrix M1M_{1} that satisfies M1+M1T=MM_{1}+M_{1}^{T}=M is OU-realizable.

Proof.

Let MM be a CN-realizable n×nn\times n matrix and let BB be an nn-braid diagram such that N(B)=MN(B)=M. Let M1M_{1} be a matrix such that M1+M1T=MM_{1}+M_{1}^{T}=M, that is, M1(i,j)+M1(j,i)=M(i,j)M_{1}(i,j)+M_{1}(j,i)=M(i,j), and M1(i,j)0M_{1}(i,j)\geq 0 for any ii and jj. For each pair of strands sis_{i} and sjs_{j} of BB, apply crossing changes if necessary so that sis_{i} is over sjs_{j} at M1(i,j)M_{1}(i,j) crossings and sis_{i} is under sjs_{j} at M1(j,i)M_{1}(j,i) crossings. Then we obtain a braid diagram BB^{\prime} such that U(B)=M1U(B^{\prime})=M_{1}. ∎

Since a crossing change does not change the braid permutation, we have the following corollary from the proof of Proposition 9.

Corollary 1.

If an n×nn\times n symmetric matrix MM is CN-realizable by a pure nn-braid diagram, then any non-negative integer matrix M1M_{1} that satisfies M1+M1T=MM_{1}+M_{1}^{T}=M is OU-realizable by a pure nn-braid diagram.

The following proposition allows us to extend the size of CN-realizable matrices.

Proposition 10.

Let M1M_{1} be a CN-realizable n×nn\times n matrix. Let M2M_{2} be an (n+1)×(n+1)(n+1)\times(n+1) matrix such that M2(i,j)=M1(i,j)M_{2}(i,j)=M_{1}(i,j) for each 1i,jn1\leq i,j\leq n and M2(i,n+1)=M2(n+1,j)=0M_{2}(i,n+1)=M_{2}(n+1,j)=0 for each 1i,jn1\leq i,j\leq n. Then M2M_{2} is also CN-realizable.

Proof.

Let B1B_{1} be an nn-braid diagram with N(B1)=M1N(B_{1})=M_{1}. Let B2B_{2} be an (n+1)(n+1)-braid diagram obtained from B1B_{1} by adding a straight strand on the right-hand side. Then, N(B2)=M2N(B_{2})=M_{2}. ∎

3 (0,2)(0,2)-CN matrix of pure braid diagram

From now on in Sections 3 and 4, we denote the CN matrices, which are zero-diagonal symmetric matrices, by strictly upper triangular matrices by replacing the entries below the main diagonal with 0, for simplicity. For example, we denote the CN matrix

(0202202002022020) by (0202002000020000).\displaystyle\begin{pmatrix}0&2&0&2\\ 2&0&2&0\\ 0&2&0&2\\ 2&0&2&0\end{pmatrix}\text{ by }\begin{pmatrix}0&2&0&2\\ 0&0&2&0\\ 0&0&0&2\\ 0&0&0&0\end{pmatrix}.
Definition 3.

We call a matrix whose entries are 0 or 22 a (0,2)(0,2)-matrix. In particular, when a CN matrix is a (0,2)(0,2)-matrix, we call it a (0,2)(0,2)-CN matrix.

In this section, we explore (0,2)(0,2)-CN matrices. In Section 3.1, we see some properties of the (0,2)(0,2)-CN matrices. In Section 3.2, we introduce a tool to construct a braid diagram that realizes the (0,2)(0,2)-CN matrix for some (0,2)(0,2)-matrices.

3.1 Properties of (0,2)(0,2)-CN matrix

We have the following proposition from Proposition 3.

Proposition 11.

When the CN matrix N(B)N(B) of a braid diagram BB is a (0,2)(0,2)-matrix, BB is a pure braid diagram.

Then we have the following proposition.

Proposition 12.

If n×nn\times n (0,2)(0,2)-matrices M1M_{1} and M2M_{2} are CN-realizable, then the sum M1+M2M_{1}+M_{2} is also CN-realizable.

Proof.

Let B1B_{1}, B2B_{2} be nn-braid diagrams such that N(B1)=M1N(B_{1})=M_{1}, N(B2)=M2N(B_{2})=M_{2}. Since M1M_{1}, M2M_{2} are (0,2)(0,2)-matrices, B1B_{1}, B2B_{2} are pure braid diagrams by Proposition 11. By Proposition 2, then, N(B1B2)=N(B1)+N(B2)=M1+M2N(B_{1}B_{2})=N(B_{1})+N(B_{2})=M_{1}+M_{2}. Hence, M1+M2M_{1}+M_{2} is CN-realizable, too. ∎

The following proposition implies that studying (0,2)(0,2)-CN matrices is essential for the study of CN matrices of pure braid diagram.

Proposition 13.

Let MM be an n×nn\times n strictly upper triangular matrix whose entries are non-negative even numbers. Let M02M^{02} be the (0,2)(0,2)-matrix obtained from MM by replacing all positive entries with 2. If M02M^{02} is CN-realizable, then MM is also CN-realizable by a pure braid diagram.

Proof.

Let BB be an nn-braid diagram such that N(B)=M02N(B)=M^{02}. We note that BB is a pure braid diagram by Proposition 11. For each entry M(i,j)M(i,j), if M(i,j)M02(i,j)M(i,j)\neq M^{02}(i,j), that is, M(i,j)4M(i,j)\geq 4, replace one of the crossings between sis_{i} and sjs_{j} of BB by M(i,j)1M(i,j)-1 half twists on BB, as shown in Figure 3. We note that this replacememt does not change the braid permutation because M(i,j)1M(i,j)-1 is also an odd number. Thus, we obtain another pure braid diagram BB^{\prime}, which satisfies N(B)=MN(B^{\prime})=M. ∎

Refer to caption
Figure 3: Replacement of a crossing with half twists.

3.2 BW-ladder diagram

In this subsection, we introduce the “BW-ladder diagram” to discuss the CN-realizablity for (0,2)(0,2)-matrices. We call a filled vertex a black vertex and an open vertex a white vertex. A BW-ladder diagram is a ladder-fashioned diagram consisting of some vertical segments and two types of horizontal edges, a black edge and a white edge which have black vertices and white vertices at the endpoints on vertical segments, respectively. In particular, we call a BW-ladder diagram with only black (resp. white) edges a B-ladder diagram (resp. W-ladder diagram). We consider white edges whose endpoints are on vertical segments which are next to each other only. We denote a black edge that has black vertices at the ithi^{th} and jthj^{th} vertical segments (i<ji<j) from the left-hand side by BjiB^{i}_{j}, as shown in Figure 4. Similarly, we denote a white edge that has white vertices at the ithi^{th} and (i+1)th(i+1)^{th} vertical segments by Wi+1iW^{i}_{i+1}. Each BW-ladder diagram can be represented by a word of BjiB^{i}_{j} and Wi+1iW^{i}_{i+1}.

Refer to caption
Figure 4: A black edge BjiB^{i}_{j} can be considered as a hook. A white edge Wi+1iW^{i}_{i+1} can be considered as a half twist. Undesired crossings are marked with a small circle.

For an n×nn\times n strictly upper triangular (0,2)(0,2)-matrix MM, a B-ladder diagram of MM is a B-ladder diagram with nn vertical segments that has a black edge BjiB^{i}_{j} if M(i,j)=2M(i,j)=2. Considering BjiB^{i}_{j} as a hook of the two segments where BjiB^{i}_{j} has roots, as shown in the left-hand side in Figure 4, we can restore MM from the B-ladder diagram. Moreover, considering Wi+1iW^{i}_{i+1} as a half-twist of the segments where Wi+1iW^{i}_{i+1} has roots, as shown in Figure 4, we consider the following local moves.

Definition 4.

A ladder move is one of the following moves L1 to L9 on a BW-ladder diagram. (See Figure 5.)

  • (L1)

    Bi+1iWi+1iWi+1iB^{i}_{i+1}\leftrightarrow W^{i}_{i+1}W^{i}_{i+1} for any ii.

  • (L2)

    BjiBlkBlkBjiB^{i}_{j}B^{k}_{l}\leftrightarrow B^{k}_{l}B^{i}_{j} for any i<j,k<li<j,\ k<l.

  • (L3)

    Wk+1kWl+1lWl+1lWk+1kW^{k}_{k+1}W^{l}_{l+1}\leftrightarrow W^{l}_{l+1}W^{k}_{k+1} when k+1<lk+1<l or l+1<kl+1<k.

  • (L4)

    Wi+1iWi+2i+1Wi+1iWi+2i+1Wi+1iWi+2i+1W^{i}_{i+1}W^{i+1}_{i+2}W^{i}_{i+1}\leftrightarrow W^{i+1}_{i+2}W^{i}_{i+1}W^{i+1}_{i+2} for any ii.

  • (L5)

    BjiWk+1kWk+1kBjiB^{i}_{j}W^{k}_{k+1}\leftrightarrow W^{k}_{k+1}B^{i}_{j} when j<kj<k, k+1<ik+1<i or i<k<k+1<ji<k<k+1<j.

  • (L6)

    BjiWi+1iWi+1iBji+1B^{i}_{j}W^{i}_{i+1}\leftrightarrow W^{i}_{i+1}B^{i+1}_{j} when i+1<ji+1<j.

  • (L7)

    Wi+1iBjiBji+1Wi+1iW^{i}_{i+1}B^{i}_{j}\leftrightarrow B^{i+1}_{j}W^{i}_{i+1} when i+1<ji+1<j.

  • (L8)

    BjiWjj1Wjj1Bj1iB^{i}_{j}W^{j-1}_{j}\leftrightarrow W^{j-1}_{j}B^{i}_{j-1} when i<j1i<j-1.

  • (L9)

    Wjj1BjiBj1iWjj1W^{j-1}_{j}B^{i}_{j}\leftrightarrow B^{i}_{j-1}W^{j-1}_{j} when i<j1i<j-1.

Refer to caption
Figure 5: Ladder moves.

The ladder move L1 means the replacement of a hook of two segments that are next to each other and a pair of half-twists between them. L2 means the change of the order of the black edges. L3 and L4 mean the transformations based on the braid relation. L5 means the change of the order of a black edge and a white edge whose endpoints are all different. L6 to L9 mean the transformation that a root of a hook moves along a half-twist, as shown in Figure 6.

Refer to caption
Figure 6: The ladder move L6L6.

Since a W-ladder diagram implies a braid diagram (without crossing information), we have the following proposition.

Proposition 14.

A strictly upper triangular (0,2)(0,2)-matrix MM is CN-realizable if a B-ladder diagram of MM can be transformed into a W-ladder diagram by a finite sequence of ladder moves.

Example 2.

The 5×55\times 5 (0,2)(0,2)-matrix shown in the left-hand side in Figure 7 is CN-realizable because the B-ladder diagram can be transformed into a W-ladder diagram by ladder moves.

Refer to caption
Figure 7: A sequence of ladder moves from a B-ladder diagram to a W-ladder diagram.

We have the following proposition.

Proposition 15.

The BW-ladder diagram Bll1Bll2Bll3Blk+1BlkB^{l-1}_{l}B^{l-2}_{l}B^{l-3}_{l}\dots B^{k+1}_{l}B^{k}_{l} can be transformed into Wll1Wl1l2Wl2l3Wk+2k+1Wk+1kWk+1kWk+2k+1Wl2l3Wl1l2Wll1W^{l-1}_{l}W^{l-2}_{l-1}W^{l-3}_{l-2}\dots W^{k+1}_{k+2}W^{k}_{k+1}W^{k}_{k+1}W^{k+1}_{k+2}\dots W^{l-3}_{l-2}W^{l-2}_{l-1}W^{l-1}_{l} by ladder moves.

Proof.

On Bll1Bll2Bll3Blk+1BlkB^{l-1}_{l}B^{l-2}_{l}B^{l-3}_{l}\dots B^{k+1}_{l}B^{k}_{l}, replace Bll1B^{l-1}_{l} with Wll1Wll1W^{l-1}_{l}W^{l-1}_{l} by the ladder move L1. Then Wll1Wll1Bll2Bll3Blk+1BlkW^{l-1}_{l}W^{l-1}_{l}B^{l-2}_{l}B^{l-3}_{l}\dots B^{k+1}_{l}B^{k}_{l} is transformed into
Wll1Bl1l2Bl1l3Bl1k+1Bl1kWll1W^{l-1}_{l}B^{l-2}_{l-1}B^{l-3}_{l-1}\dots B^{k+1}_{l-1}B^{k}_{l-1}W^{l-1}_{l} by applying L9 repeatedly. Replace Bl1l2B^{l-2}_{l-1} with Wl1l2Wl1l2W^{l-2}_{l-1}W^{l-2}_{l-1} by L1. Then we have Wll1Wl1l2Bl2l3Bl2k+1Bl2kWl1l2Wll1W^{l-1}_{l}W^{l-2}_{l-1}B^{l-3}_{l-2}\dots B^{k+1}_{l-2}B^{k}_{l-2}W^{l-2}_{l-1}W^{l-1}_{l} by L9. Repeat this procedure until Bk+1kB^{k}_{k+1} is replaced with Wk+1kWk+1kW^{k}_{k+1}W^{k}_{k+1}. (See Figure 8.) ∎

Refer to caption
Figure 8: A sequence from B54B43B52B^{4}_{5}B^{3}_{4}B^{2}_{5} to W54W43W32W32W43W54W^{4}_{5}W^{3}_{4}W^{2}_{3}W^{2}_{3}W^{3}_{4}W^{4}_{5}.

From Proposition 15, we have the following corollary.

Corollary 2.

Let MM be an n×nn\times n (0,2)(0,2)-matrix satisfying the following condition for some kk with 1kn1\leq k\leq n:

M(i,j)=0 for 1jn1\displaystyle M(i,j)=0\text{ for }1\leq j\leq n-1
M(i,n)={0 for 1ik1 or i=n2 for kin1\displaystyle M(i,n)=\left\{\begin{array}[]{ll}0\text{ for }1\leq i\leq k-1\text{ or }i=n\\ 2\text{ for }k\leq i\leq n-1\end{array}\right.

Then MM is CN-realizable.

Proof.

The matrix MM has the B-ladder diagram of Proposition 15 with l=nl=n, which can be transformed into a W-ladder diagram by ladder moves. ∎

From Propositions 10, 12 and Corollary 2, we have the following corollary.

Corollary 3.

Let M1M_{1} be a CN-realizable n×nn\times n strictly upper triangular (0,2)(0,2)-matrix. Then, an (n+1)×(n+1)(n+1)\times(n+1) matrix MM satisfying the following condition for some kk with 1kn1\leq k\leq n is also CN-realizable.

M(i,j)=M1(i,j) for 1jn\displaystyle M(i,j)=M_{1}(i,j)\text{ for }1\leq j\leq n
M(i,n+1)={0 for 1ik1 or i=n+12 for kin\displaystyle M(i,n+1)=\left\{\begin{array}[]{ll}0\text{ for }1\leq i\leq k-1\text{ or }i=n+1\\ 2\text{ for }k\leq i\leq n\end{array}\right.
Example 3.

The 6×66\times 6 matrix

M=(020200002020000202000022000002000000)\displaystyle M=\begin{pmatrix}0&2&0&2&0&0\\ 0&0&2&0&2&0\\ 0&0&0&2&0&2\\ 0&0&0&0&2&2\\ 0&0&0&0&0&2\\ 0&0&0&0&0&0\\ \end{pmatrix}

is CN-realizable because the 5×55\times 5 matrix from the first to the fifth rows and columns is CN-realizable as shown in Figure 7 and the sixth column meets the condition of Corollary 3 with n=5n=5, k=3k=3.

We also have the following proposition.

Proposition 16.

Any n×nn\times n T0 upper triangular (0,2)(0,2)-matrix with M(i,j)=0M(i,j)=0 for ji3j-i\geq 3 is CN-realizable.

Proof.

Suppose the B-ladder diagram with the following order of black edges,

B21B31B32B42B43Bk+1kBk+2kBk+2k+1Bn1n3Bn1n2Bnn2Bnn1,B^{1}_{2}B^{1}_{3}B^{2}_{3}B^{2}_{4}B^{3}_{4}\dots B^{k}_{k+1}B^{k}_{k+2}B^{k+1}_{k+2}\dots B^{n-3}_{n-1}B^{n-2}_{n-1}B^{n-2}_{n}B^{n-1}_{n},

where we ignore BjiB^{i}_{j} if M(i,j)=0M(i,j)=0. By the ladder move L1, we have

W21W21B31W32W32B42B43Wk+1kWk+1kBk+2kWk+2k+1Wk+2k+1Bnn2Wnn1Wnn1.W^{1}_{2}W^{1}_{2}B^{1}_{3}W^{2}_{3}W^{2}_{3}B^{2}_{4}B^{3}_{4}\dots W^{k}_{k+1}W^{k}_{k+1}B^{k}_{k+2}W^{k+1}_{k+2}W^{k+1}_{k+2}\dots B^{n-2}_{n}W^{n-1}_{n}W^{n-1}_{n}.

Since MM is T0, if Bk+2kB^{k}_{k+2} exists in the sequence, namely M(k,k+2)=2M(k,k+2)=2, then at least one of Wk+1kW^{k}_{k+1} or Wk+2k+1W^{k+1}_{k+2} exists in the sequence. Use the white edge, which is next to Bk+2kB^{k}_{k+2}, to apply L7 or L8 and obtain Bk+2k+1B^{k+1}_{k+2} or Bk+1kB^{k}_{k+1}. By L1, then, we obtain a W-ladder diagram. ∎

We have the following corollary.

Corollary 4.

Any n×nn\times n (0,2)(0,2)-matrix MM with M(i,j)=0M(i,j)=0 for ji2j-i\geq 2 is CN-realizable.

Proof.

Since any n×nn\times n matrix MM such that M(i,j)=0M(i,j)=0 for ji2j-i\geq 2 is T0, it follows from Proposition 16. ∎

4 CN matrices of pure braid diagrams up to 5×55\times 5

We investigate the (0,2)(0,2)-CN matrices of 2×22\times 2 to 4×44\times 4 in Section 4.1 and 5×55\times 5 in Section 4.2.

4.1 (0,2)(0,2)-CN matrices of 2×22\times 2 to 4×44\times 4

For 2×22\times 2 and 3×33\times 3 matrices, we have the following proposition.

Proposition 17.

Every 2×22\times 2 or 3×33\times 3 T0 upper triangular (0,2)(0,2)-matrix MM is CN-realizable.

Proof.

It follows from Proposition 16. ∎

For 4×44\times 4 matrices, we have the following proposition.

Proposition 18.

Every 4×44\times 4 T0 upper triangular (0,2)(0,2)-matrix MM is CN-realizable.

Proof.

By Corollary 3 and Proposition 17, if a 4×44\times 4 T0 upper triangular (0,2)(0,2)-matrix MM has the fourth column as (0,0,0,0)T(0,0,0,0)^{T}, (0,0,2,0)T(0,0,2,0)^{T}, (0,2,2,0)T(0,2,2,0)^{T} or (2,2,2,0)T(2,2,2,0)^{T}, then MM is CN-realizable. Therefore, it is sufficient to check the following four cases, which is 2342^{3}-4, Cases A, B, C, D that the fourth column is (0,2,0,0)T(0,2,0,0)^{T}, (2,0,0,0)T(2,0,0,0)^{T}, (2,0,2,0)T(2,0,2,0)^{T}, (2,2,0,0)T(2,2,0,0)^{T}, respectively.

  • Case A:

    Since the (1,4)(1,4) entry is 0, it is CN-realizable by Proposition 16.

  • Case B:

    By T0, the (1,2)(1,2) and (1,3)(1,3) entries are 22. Then the first row will be (0,2,2,2)(0,2,2,2). By Propositions 4, 17 and Corollary 3, it is CN-realizable.

  • Case C:

    By T0, the (1,2)(1,2) entry is 22. Divide the matrix as

    (02α2000000020000)+(000000β000000000),\displaystyle\begin{pmatrix}0&2&\alpha&2\\ 0&0&0&0\\ 0&0&0&2\\ 0&0&0&0\\ \end{pmatrix}+\begin{pmatrix}0&0&0&0\\ 0&0&\beta&0\\ 0&0&0&0\\ 0&0&0&0\\ \end{pmatrix},

    where α\alpha and β\beta are either 0 or 22. The first matrix is CN-realizable regardless of the value of α\alpha, as shown in Figure 9. The second matrix is also CN-realizable regardless of β\beta by Corollary 4. Hence, by Proposition 12, the sum of them is CN-realizable, too.

    Refer to caption
    Figure 9: Cases A to D. For Case C, ignore the purple-colored broken black edge when the (1,3)(1,3) entry is 0.
  • Case D:

    By T0, the (1,3)(1,3) and (2,3)(2,3) entries are 22. Divide the matrix as

    (0022002200000000)+(0β00000000000000),\displaystyle\begin{pmatrix}0&0&2&2\\ 0&0&2&2\\ 0&0&0&0\\ 0&0&0&0\\ \end{pmatrix}+\begin{pmatrix}0&\beta&0&0\\ 0&0&0&0\\ 0&0&0&0\\ 0&0&0&0\\ \end{pmatrix},

    where β\beta is 0 or 22. Since the first matrix is CN-realizable as shown in Figure 9 and the second matrix is CN-realizable by Corollary 4, the sum is also CN-realizable by Proposition 12.

4.2 (0,2)(0,2)-CN matrices of 5×55\times 5

In this subsection, we show the following proposition.

Proposition 19.

Every 5×55\times 5 T0 upper triangular (0,2)(0,2)-matrix MM is CN-realizable.

To prove Proposition 19, we prepare three lemmas below.

Lemma 1.

Let MM be a 5×55\times 5 T0 upper triangular (0,2)(0,2)-matrix such that M(i,j)=0M(i,j)=0 when i<Ii<I or j>Jj>J for some I,JI,J with 1I<J51\leq I<J\leq 5 and JI3J-I\leq 3. Then MM is CN-realizable.

Proof.

Consider the (JI+1)×(JI+1)(J-I+1)\times(J-I+1) matrix M1M_{1} such that M1(i,j)=M(i+I1,j+I1)M_{1}(i,j)=M(i+I-1,j+I-1). We note that M1M_{1} is T0 because MM is T0. Since JI+14J-I+1\leq 4, M1M_{1} is CN-realizable by Propositions 17 and 18. Let B1B_{1} be a (JI+1)(J-I+1)-braid diagram such that N(B1)=M1N(B_{1})=M_{1}. Add I1I-1 and 5J5-J straight strands to B1B_{1} on the left- and right-hand sides, respectively, as shown in Figure 10 (3). Thus, we obtain a 55-braid diagram BB such that N(B)=MN(B)=M. ∎

Refer to caption
Figure 10: (1): The case that I=2I=2, J=4J=4. (2): The case that I=1I=1, J=3J=3. (3): The case that I=3I=3, J=4J=4 and the construction of a braid diagram BB.
Lemma 2.

If a 5×55\times 5 T0 upper triangular (0,2)(0,2)-matrix MM has the configuration of one of aa to hh in Figure 11, then MM is CN-realizable.

Refer to caption
Figure 11: The configurations aa to hh and their reverses. The entries marked * is either 0 or 22.
Proof.

We use Proposition 12 and Lemma 1. For the configuration aa, divide the matrix as

(022α1α20000α3000020000200000)+(0000000β1β20000β300000000000),\displaystyle\begin{pmatrix}0&2&2&\alpha_{1}&\alpha_{2}\\ 0&0&0&0&\alpha_{3}\\ 0&0&0&0&2\\ 0&0&0&0&2\\ 0&0&0&0&0\end{pmatrix}+\begin{pmatrix}0&0&0&0&0\\ 0&0&\beta_{1}&\beta_{2}&0\\ 0&0&0&\beta_{3}&0\\ 0&0&0&0&0\\ 0&0&0&0&0\end{pmatrix},

where αi\alpha_{i} and βj\beta_{j} are either 0 or 2 (i,j=1,2,3i,j=1,2,3). The first matrix is CN-realizable as shown in Figure 12 regardless of the values of α1\alpha_{1}, α2\alpha_{2} and α3\alpha_{3}.

Refer to caption
Figure 12: Any colored broken black edge BjiB^{i}_{j} of B41B^{1}_{4}, B51B^{1}_{5} and B52B^{2}_{5} can be ignored when the corresponding (i,j)(i,j) entry of the matrix is 0.

The second matrix is CN-realizable by Lemma 1. Hence, the sum is also CN-realizable by Proposition 12. In the same way, we can see the CN-realizability for the configurations bb to ee as shown in Figure 13. We note that we use Lemma 1 and Proposition 12 twice for the configurations c,d,ec,d,e.

Refer to caption
Figure 13: Any colored broken black edge BjiB^{i}_{j} can be ignored when the corresponding (i,j)(i,j) entry of the matrix is 0.

The configuration ff realizes a CN matrix by Proposition 16. For the configurations gg and hh, the matrices

(0022200222000000000000000) and (0002000222000200000000000)\displaystyle\begin{pmatrix}0&0&2&2&2\\ 0&0&2&2&2\\ 0&0&0&0&0\\ 0&0&0&0&0\\ 0&0&0&0&0\\ \end{pmatrix}\text{ and }\begin{pmatrix}0&0&0&2&0\\ 0&0&2&2&2\\ 0&0&0&2&0\\ 0&0&0&0&0\\ 0&0&0&0&0\\ \end{pmatrix}

are CN-realizable as shown in Figure 14.

Refer to caption
Figure 14: The configurations gg and hh realize CN matrices.

Hence, the configurations gg and hh also realize CN-realizable matrices by Lemma 1 and Proposition 12. ∎

We note that the reverses aa^{\prime} to hh^{\prime} of the configurations aa to hh also realize a CN matrix by Proposition 4. We also have the following lemma.

Lemma 3.

If a 5×55\times 5 T0 upper triangular (0,2)(0,2)-matrix MM has one of the configurations shown in Figure 15, which we call configuration ii, then MM is CN-realizable.

Proof.

This follows from Proposition 4 and Corollary 3. ∎

Refer to caption
Figure 15: The configuration ii.

Now we prove Proposition 19.

Proof of Proposition 19.  By Corollary 3, it is sufficient to check the eleven cases, which is 2452^{4}-5, on the fifth column A to K in Figure 16. By T0, some of the entries in the first to fourth columns are automatically assigned, as shown in Figure 16.

Refer to caption
Figure 16: Cases A to K.
  • Case A:

    By the configuration ff, it is sufficient to suppose that the (1,4)(1,4) entry is 22. Then, by configuration cc, suppose that the (1,2)(1,2) entry is 0. By T0, then, the (2,4)(2,4) entry is 22, and observe that it has configuration dd. Hence, any matrix MM of Case A is CN-realizable.

  • Case B:

    By the configuration dd^{\prime}, suppose that the (0,3)(0,3) entry is 0. By configuration ii, suppose that the (1,4)(1,4) entry is 22. This has the configuration hh.

  • Case C:

    This has the configuration ii.

  • Case D:

    By configuration cc^{\prime}, suppose that the (1,4)(1,4) entry is 22. Then, by configuration ii, suppose the three subcases shown in Figure 17. The first subcase has the configuration hh. The second subcase has the braid diagram as shown in the figure. (See Example 7 for the construction of the braid diagram.) The third has the configuration dd^{\prime}.

  • Case E:

    By the configuration ii, suppose that the (1,4)(1,4) entry is 0. This has the configuration bb.

  • Case F:

    This has the configuration dd.

  • Case G:

    This has the configuration ee^{\prime}.

  • Case H:

    This has the configuration gg.

  • Case I:

    By the configuration aa, suppose that the (1,3)(1,3) entry is 0. By the configuration bb^{\prime}, suppose that the (1,4)(1,4) entry is 22. By T0, then, the (3,4)(3,4) entry is 22. This has the configuration ee^{\prime}.

  • Case J:

    By the configuration ee, suppose that the (1,4)(1,4) entry is 22. By the configuration hh, suppose that the (2,4)(2,4) entry is 22. Then, it has the configuration gg.

  • Case K:

    This has the configuration gg^{\prime}.

Refer to caption
Figure 17: Cases A, B, D, E, I and J.

Hence, every 5×55\times 5 T0 upper triangular (0,2)(0,2)-matrix is CN-realizable. ∎

By Propositions 17, 18 and 19, we have the following corollary.

Corollary 5.

When n5n\leq 5, any n×nn\times n T0 upper triangular (0,2)(0,2)-matrix is CN-realizable.

We also have the following corollary by Corollary 5 and Proposition 13.

Corollary 6.

When n5n\leq 5, any n×nn\times n non-negative even T0 upper triangular matrix is CN-realizable by a pure nn-braid diagram.

5 Proof of Theorem 1 and applications

We prove Theorem 1 in Section 5.1 and then discuss the crossing matrices of positive pure braids in Section 5.2, the OU matrices of braid diagrams that are not necessarily pure in Section 5.3. In this section, we consider a CN matrix as a symmetric matrix as defined in Section 2, not a strictly upper triangular matrix as in Sections 3 and 4.

5.1 Proof of Theorem 1

In this subsection, we prove Theorem 1. We have the following proposition.

Proposition 20.

When n5n\leq 5, an n×nn\times n matrix MM is the CN matrix of a pure nn-braid diagram if and only if MM is a non-negative even T0 symmetric matrix.

Proof.

By Propositions 3 and 5, the CN matrix N(B)N(B) of a pure nn-braid diagram BB is a non-negative even T0 symmetric matrix. On the other hand, a non-negative even T0 symmetric matrix is CN-realizable by a pure nn-braid diagram by Corollary 6. ∎

We prove Theorem 1.

Proof of Theorem 1.  By Propositions 3 and 5, the OU matrix U(B)U(B) of any pure nn-braid diagram BB is an n×nn\times n matrix such that U(B)+(U(B))TU(B)+(U(B))^{T} is a non-negative even T0 matrix.

Let MM be an n×nn\times n zero-diagonal matrix (n5n\leq 5) such that M+MTM+M^{T} is a non-negative even T0 matrix. Since M+MTM+M^{T} is CN-realizable by a pure nn-braid diagram by Corollary 6, MM is OU-realizable by a pure nn-braid diagram by Corollary 1. ∎

5.2 C-realizable matrix

In this subsection, we discuss the crossing matrices of positive pure braids. We have the following lemmas.

Lemma 4.

A zero-diagonal symmetric matrix MM is T0 if and only if M+MTM+M^{T} is T0.

Proof.

When MM is symmetric, i.e., M(i,j)=M(j,i)M(i,j)=M(j,i), M(i,j)=0M(i,j)=0 if and only if M(i,j)+MT(i,j)=0M(i,j)+M^{T}(i,j)=0. ∎

Lemma 5.

When n5n\leq 5, an n×nn\times n matrix MM is the OU matrix of a pure nn-braid diagram such that each pair of strands sis_{i} and sjs_{j} has the same number of crossings such that sis_{i} is over sjs_{j} and sjs_{j} is over sis_{i} if and only if MM is a non-negative integer T0 symmetric matrix.

Proof.

By Theorem 1, the OU matrix of a pure nn-braid MM (n5)(n\leq 5) satisfies that M+MTM+M^{T} is a non-negative even T0 n×nn\times n matrix. Moreover, MM is symmetric with the condition of the crossing number of sis_{i} and sjs_{j}, i.e., M(i,j)=M(j,i)M(i,j)=M(j,i). Hence MM itself is also T0 by Lemma 4.

On the other hand, let MM be an n×nn\times n non-negative integer T0 symmetric matrix (n5n\leq 5). Then M+MTM+M^{T} is also T0 by Lemma 4 and the entries of M+MTM+M^{T} are all non-negative even numbers since MM is symmetric. Therefore, M+MTM+M^{T} is CN-realizable by a pure nn-braid diagram by Proposition 20. Then MM is OU-realizable by a pure nn-braid diagram BB by Corollary 1. Moreover, BB is a braid diagram such that each pair of strands sis_{i}, sjs_{j} of BB has the same number of crossings such that sis_{i} is over sjs_{j} and sjs_{j} is over sis_{i} because M(i,j)=M(j,i)M(i,j)=M(j,i). ∎

We have the following corollaries.

Corollary 7.

When n5n\leq 5, the crossing matrix C(B)C(B) of any positive pure nn-braid diagram BB is an n×nn\times n non-negative integer T0 symmetric matrix.

Proof.

It follows from Lemma 5 since C(B)=U(B)C(B)=U(B) when BB is a positive braid diagram BB by Proposition 7. ∎

Corollary 8.

When n5n\leq 5, any n×nn\times n non-negative integer T0 symmetric matrix is C-realizable by a positive pure nn-braid.

Proof.

For the pure nn-braid diagram BB in the proof of Lemma 5 with U(B)=MU(B)=M, apply crossing changes if necessary to obtain a positive braid diagram BB^{\prime}. We note that BB^{\prime} is still a pure nn-braid diagram such that each pair of strands sis_{i} and sjs_{j} has the same number of crossings such that sis_{i} is over sjs_{j} and sjs_{j} is over sis_{i}. Then C(B)=U(B)=U(B)=MC(B^{\prime})=U(B^{\prime})=U(B)=M by Proposition 7 and Lemma 5. ∎

By Corollaries 7 and 8, we have the following corollary.

Corollary 9.

Conjecture 1 is true when n5n\leq 5.

5.3 OU matrix and CN matrix for a general braid type

A simple braid (or a permutation braid) is a braid that has a simple diagram, that is, each crossing is positive and each pair of strands has at most one crossing, which has been used in studies of the normal form of braids (see [3, 4, 10]). Let Σn\Sigma_{n} be the set of simple nn-braids. There is one-to-one correspondence from Σn\Sigma_{n} to Sn=Sym{1,,n}S_{n}=Sym\{1,\dots,n\} by taking their braid permutations ρ\rho. Let XnX_{n} be the set of n×nn\times n zero-diagonal integer matrices. As explained in [2, 5], using Lemma 9.1.6 in [10] directly, we have the following lemma.

Lemma 6.

The map C|Σn:ΣnXnC|_{\Sigma_{n}}:\Sigma_{n}\to X_{n} is injective and its image C(Σn)C(\Sigma_{n}) is the set of LXnL\in X_{n} satisfying the following conditions:

  • (i)

    L(i,j)=0L(i,j)=0 if iji\geq j,

  • (ii)

    L(i,j){0,1}L(i,j)\in\{0,1\},

  • (iii)

    LL is both T0 and T1, i.e., if whenever 1i<j<kn1\leq i<j<k\leq n, then L(i,j)=L(j,k)=pL(i,j)=L(j,k)=p implies L(i,k)=pL(i,k)=p for p=0,1p=0,1.

We call a square matrix satisfying the conditions (i)-(iii) in Lemma 6 a simple matrix. Take a simple diagram B0B_{0} of b0Σnb_{0}\in\Sigma_{n}. If a pair of strands sis_{i} and sj(i<j)s_{j}\ (i<j) of B0B_{0} has a crossing, then sis_{i} passes under sjs_{j}, and equivalently sjs_{j} passes over sis_{i} at the crossing. Thus, U(B0)=C(B0)(=C(b0))U(B_{0})=C(B_{0})(=C(b_{0})) for any simple diagram B0B_{0} of any b0Σnb_{0}\in\Sigma_{n}. We note that N(B0)=U(B0)+U(B0)TN(B_{0})=U(B_{0})+U(B_{0})^{T} and N(i,j)=U(i,j)N(i,j)=U(i,j) for any i,ji,j. By Lemma 6, we have the following lemma.

Lemma 7.

The map N|Σn:ΣnXnN|_{\Sigma_{n}}:\Sigma_{n}\to X_{n} is injective and its image N(Σn)N(\Sigma_{n}) is the set of LXnL\in X_{n} satisfying the following conditions:

  • (I)

    L(i,j)=L(j,i)L(i,j)=L(j,i) for any i,ji,j,

  • (II)

    L(i,j){0,1}L(i,j)\in\{0,1\},

  • (III)

    LL is T0 and T1.

We call a matrix satisfying the conditions (I)-(III) in Lemma 7 a double simple matrix. We also remark that C(Σn)Xn+C(\Sigma_{n})\subset X_{n}^{+} and C(Σn)Xn+C(\Sigma_{n})\subset X_{n}^{+}, where Xn+X_{n}^{+} is the set of matrices MXnM\in X_{n} such that M(i,j)0M(i,j)\geq 0 for any i,ji,j. In conclusion, we have the following corollary.

Corollary 10.

Let n5n\leq 5. For each braid bBnb\in B_{n},

  • (1)

    there exists a braid diagram BB of bb such that U(B)=M1+L1U(B)=M_{1}+L_{1} for some M1,L1Xn+M_{1},L_{1}\in X_{n}^{+} such that M1+M1T{M_{1}}+M_{1}^{T} is non-negative even T0 and L1L_{1} is simple.

  • (2)

    There exists a braid diagram BB of bb such that N(B)=M2+L2N(B)=M_{2}+L_{2} for some M2,L2Xn+M_{2},L_{2}\in X_{n}^{+} such that M2+M2TM_{2}+M^{T}_{2} is non-negative even T0 symmetric and L2L_{2} is double simple.

Proof.

Let ρbSn\rho_{b}\in S_{n} be the braid permutation of bb and let ρb+Σn\rho_{b}^{+}\in\Sigma_{n} be the permutation nn-braid corresponding to ρb\rho_{b}. Let a=b(ρb+)1Bna=b(\rho_{b}^{+})^{-1}\in B_{n}. It is easily seen that aPna\in P_{n}. Take a pure braid diagram AA of aa and take a simple diagram B0B_{0} of ρb+\rho_{b}^{+}. Let L1=U(B0),L2=N(B0)(=L1+L1T),M1=U(A)L_{1}=U(B_{0}),\ L_{2}=N(B_{0})(={L_{1}}+L_{1}^{T}),M_{1}=U(A) and M2=N(A)(=M1+M2T)M_{2}=N(A)(=M_{1}+M_{2}^{T}). Then, U(B)=M1+L1U(B)=M_{1}+L_{1} by Proposition 1 and N(B)=M2+L2N(B)=M_{2}+L_{2} by Proposition 2. We can see that M1+M1TM_{1}+M_{1}^{T} is non-negative even T0 by Theorem 1 and M2+M2TM_{2}+M_{2}^{T} is non-negative even T0 symmetric by Proposition 20. We can also see that L1L_{1} is simple by Lemma 6 and L2L_{2} is double simple by Lemma 7. ∎

Acknowledgment

The work of the first author was partially supported by the JSPS KAKENHI Grant Number JP21K03263. The work of the second author was partially supported by the JSPS KAKENHI Grant Number JP19K03508.

References

  • [1] E. Artin, Theory of braids, Ann. Math. 48 (1947), 101–126.
  • [2] J. Burillo, M. Gutierrez, S. Krstić and Z. Nitecki, Crossing matrices and Thurston’s normal form for braids, Topl. Appl. 118 (2002), 293–308.
  • [3] E. Elrifai, H. Morton, Algorithms for positive braids, Quart. J. Math. Oxford 45 (2)(1994), 479 497.
  • [4] F. A. Garside, The braid group and other groups, Quart J. Math. Oxford Ser. 20 (1969), 235–254.
  • [5] M. Gutierrez and Z. Nitecki, Crossing matrix of positive braids. (arXiv:1805.12189)
  • [6] A. Kawauchi, Lectures on knot theory (in Japanese), Kyoritsu Shuppan Co. Ltd. (2007).
  • [7] A. Shimizu, The warping degree of a link diagram, Osaka J. Math. 48 (2011), 209–231.
  • [8] A. Shimizu, A. Gill and S. Joshi, A note on the unknotting number and the region nuknotting number of weaving knots, preprint. (arXiv:2408.14938)
  • [9] A. Shimizu and Y. Yaguchi, Determinant of the OU matrix of a braid diagram, to appear in J. Knot Theory Ramifications.
  • [10] W. Thurston, Braid groups, in: D. Epsteinetal. (Eds.), Word Processing in Groups, Jonesand Bartlett, Boston, MA, 1992, pp.181–209.