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

Non-intersecting path explanation for block Pfaffians and applications into skew-orthogonal polynomials

Zong-Jun Yao Department of Mathematics, Sichuan University, Chengdu, 610064, China  and  Shi-Hao Li Department of Mathematics, Sichuan University, Chengdu, 610064, China shihao.li@scu.edu.cn
Abstract.

In this paper, we mainly consider a combinatoric explanation for block Pfaffians in terms of non-intersecting paths, as a generalization of results obtained by Stembridge. As applications, we demonstrate how are generating functions of non-intersecting paths related to skew orthogonal polynomials and their deformations, including a new concept called multiple partial-skew orthogonal polynomials.

Key words and phrases:
LGV lemma, Stembridge’s theorem, Pfaffians, orthogonal polynomials, skew-orthogonal polynomials
2020 Mathematics Subject Classification:
15B52, 15A15, 33E20

1. Introduction

Enumerative combinatorics is always playing a fundamental role in mathematics and physics. Among these combinatorial objects, the theory of non-intersecting paths is an extremely useful theory with applications in counting tilings, plane partitions, and tableaux. In particular, the famous Lindström-Gessel-Vionnet (LGV) lemma states that the generating function of non-intersecting paths between vertex sets 𝐮𝐯\mathbf{u}\to\mathbf{v}, where 𝐮\mathbf{u} and 𝐯\mathbf{v} have the same number of vertices, could be written as a determinant [24, 11]. This result has many applications such as non-intersecting random walks [17, 18], domino tiling models [8], dimer models [4], and so on. It turns out that the generating functions of non-intersecting paths are usually expressed in terms of structured determinants such as Hankel or Toeplitz determinants. Moreover, it has been shown in [26, 6] that the Hankel determinant of some special combinatorial numbers could be evaluated by using orthogonal polynomials, continued fraction and discrete integrable systems. Therefore, there is a unified relation between LGV lemma, integrable models, probability models, and orthogonal polynomials.

In [31], the author developed another algebraic tool—Pfaffian, to describe non-intersecting paths between a set of vertices to an interval. This combinatorial description for Pfaffian was also applied to count skew Young tableaux and plane partitions [27, 31]. A minor summation formula for Pfaffian [16] was given by using similar combinatorial technique, called the path switching involution in [21]. The minor summation formula is as an analogy of Cauchy-Binet formula for determinants, and it was applied to evaluate a certain Catalan-Hankel type Pfaffian

Pf((ji)μi+j+k)i,j=02n1,\displaystyle\operatorname{Pf}\left((j-i)\mu_{i+j+k}\right)_{i,j=0}^{2n-1}, (1.1)

where {μi}i\{\mu_{i}\}_{i\in\mathbb{N}} are moments sequence related to little qq-Jacobi polynomials [15]. In fact, Pfaffian expression (1.1) is an important quantity in random matrix theory, especially for the characterization of symplectic invariant ensemble [25]. By realizing that (1.1) is a normalization factor for certain skew-orthogonal polynomials, various Catalan-Hankel type Pfaffians were computed by using skew-orthogonal polynomials under Askey-Wilson scheme in [29].

Inspired by these results, we attempt to give a unified frame between non-intersecting paths, Pfaffian formulas, and skew-orthogonal polynomials. It was known from [31] that the generating function of non-intersecting paths from 𝐮I\mathbf{u}\to I could be written as a Pfaffian, namely,

GF[𝒫0(𝐮;I)]=Pf[QI(u(i),u(j))]i,j=12n,\displaystyle\operatorname{GF}[\mathscr{P}_{0}(\mathbf{u};I)]=\operatorname{Pf}\left[Q_{I}(u^{(i)},u^{(j)})\right]_{i,j=1}^{2n}, (1.2)

where 𝒫0(𝐮;I)\mathscr{P}_{0}(\mathbf{u};I) stands for all non-intersecting paths from an even-numbered vertex set 𝐮=(u(1),,u(2n))\mathbf{u}=(u^{(1)},\cdots,u^{(2n)}) to an interval II and QI(u(i),u(j))Q_{I}(u^{(i)},u^{(j)}) are some computable weights. Different from the Pfaffian formula considered in (1.1), we show that this generating function is related to another type of skew-orthogonal polynomials, which are related to orthogonal invariant ensemble in random matrix theory. Besides, if the vertex set 𝐮=(u(1),,u(2n+1))\mathbf{u}=(u^{(1)},\cdots,u^{(2n+1)}), then its corresponding generating function could be expressed as an augmented Pfaffian

GF[𝒫0(𝐮;I)]=Pf[QI(u(i),u(j))QI(u(i))QI(u(j))0]1i,j2n+1,\displaystyle\operatorname{GF}[\mathscr{P}_{0}(\mathbf{u};I)]=\operatorname{Pf}\left[\begin{array}[]{cc}Q_{I}(u^{(i)},u^{(j)})&Q_{I}(u^{(i)})\\ -Q_{I}(u^{(j)})&0\end{array}\right]_{1\leq i,j\leq 2n+1},

where QI(u(j))Q_{I}(u^{(j)}) is a weight function related to u(j)u^{(j)} only. Since skew-orthogonal polynomials are only related to the even-ordered Pfaffian formula (1.2), we demonstrate that this augmented Pfaffian should be related to partial-skew-orthogonal polynomials proposed in [5].

In recent years, matrix-valued orthogonal polynomials play an important role in probability and combinatoric models such as hexagon tilings [14] and periodic Aztec diamond models [8]. Despite of block determinants, Pfaffians of skew symmetric block matrices were also proved to be useful in practice [22]. It was shown in [31] that the generating function for 𝐮𝐯I\mathbf{u}\to\mathbf{v}\oplus I could be expressed as a block Pfaffian, with one block being all zeros. Here 𝐮\mathbf{u} and 𝐯\mathbf{v} are two sets of vertices and II is an interval of vertices. Our first result is to give a non-intersecting path interpretation for block Pfaffians

Pf[ABBC].\displaystyle\operatorname{Pf}\left[\begin{array}[]{cc}A&B\\ -B^{\top}&C\end{array}\right]. (1.5)

We show that this Pfaffian could be expressed as the generating function of non-intersecting paths 𝒫0(J𝐮;𝐯I)\mathscr{P}_{0}(J\oplus\mathbf{u};\mathbf{v}\oplus I), where 𝐮\mathbf{u} and 𝐯\mathbf{v} are two sets of vertices and II and JJ are two intervals of vertices. Besides, we show that this generating function is related to a 2-component skew-orthogonal polynomials, which could be used to describe a non-intersecting random walk with two different sources. Moreover, a concept of multiple partial-skew-orthogonal polynomials is proposed if the number of vertices in 𝐮\mathbf{u} and 𝐯\mathbf{v} are odd. Furthermore, we extend formula (1.5) to a more general case, where we consider the paths from {𝐮1,,𝐮n,I1,,Im}\{\mathbf{u}_{1},\cdots,\mathbf{u}_{n},I_{1},\cdots,I_{m}\} to {J1,,Jn,𝐯1,,𝐯m}\{J_{1},\cdots,J_{n},\mathbf{v}_{1},\cdots,\mathbf{v}_{m}\}. We show that this generating function could also be written as a block Pfaffian. We remark that in the reference [3], a non-intersecting path explanation for Pfaffians (1.5) was given, but it was shown by using Grassmann algebra in a cyclic digraph.

This paper is organized as follows. In Section 2, we recall the LGV lemma and show some applications. We connect the LGV lemma with several different orthogonal polynomials (including orthogonal polynomials on the real line and on the unit circle) and random walks (including discrete-time random walks and continuous-time random walks). In Section 3, we recall Stembridge’s results on the non-intersecting path explanation for Pfaffians. Their connections with skew-orthogonal polynomials and partial-skew-orthogonal polynomials are given. We show that this generating function is related to some random walk models as well. Thus we give a unified frame between non-intersecting paths, Pfaffian formulas and skew-orthogonal polynomials. We generalize Stembridge’s result in Section 4, where we consider paths from J𝐮𝐯IJ\oplus\mathbf{u}\to\mathbf{v}\oplus I. Multiple skew-orthogonal polynomials are introduced by following this generating function, with some applications in random walks starting from different sources. Moreover, we introduce a new concept of multiple partial-skew-orthogonal polynomials from this generating function. In Section 5, non-intersecting paths between several vertex sets and intervals are considered. This is the most general case between vertex sets and intervals.

2. Lindström-Gessel-Vionnet theorem and applications into orthogonal polynomials

In this part, we give some brief reviews on the connections between Lindström-Gessel-Vionnet lemma and orthogonal polynomials. There have been numerous references about this topic, and please refer to [12, 21, 30, 28] and references therein.

Definition 2.1.

Let VV be a set of vertices, EE be a set of directed edges between vertices in VV, and D=(V,E)D=(V,E) represents an acyclic graph formed by VV and EE. We assume that the weight of each edge should be greater than zero.

Definition 2.2 (𝒫(u;v)\mathscr{P}(u;v)).

Let uu and vv be two vertices in DD. We use 𝒫(u;v)\mathscr{P}(u;v) to denote the set of all paths from uu to vv in the graph DD.

Let PP be a path from uu to vv, that is, P𝒫(u;v)P\in\mathscr{P}(u;v), then the weight of PP denoted by ω(P)\omega(P), is defined as the product of the weights of all edges on this path. Moreover, the weight between two vertices uu and vv is defined as h(u,v)=P𝒫(u;v)ω(P)h(u,v)=\sum\limits_{P\in{\mathscr{P}}(u;v)}{\omega(P)}.

Definition 2.3 (𝒫(𝐮;𝐯)\mathscr{P}(\mathbf{u};\mathbf{v}), 𝒫0(𝐮;𝐯)\mathscr{P}_{0}(\mathbf{u};\mathbf{v})).

Let 𝐮=(u(1),u(2),,u(r))\mathbf{u}=(u^{(1)},u^{(2)},\cdots,u^{(r)}) and 𝐯=(v(1),v(2),,v(r))\mathbf{v}=(v^{(1)},v^{(2)},\cdots,v^{(r)}) be two sequences of vertices in VV, which are composed by rr points in an ascending order, namely, we assume that u(1)<<u(r)u^{(1)}<\cdots<u^{(r)} and v(1)<<v(r)v^{(1)}<\cdots<v^{(r)}. We denote 𝒫(𝐮;𝐯)\mathscr{P}(\mathbf{u};\mathbf{v}) by the set of all paths from 𝐮\mathbf{u} to 𝐯\mathbf{v} in the graph DD in the order u(1)v(1)u^{(1)}\to v^{(1)}, u(2)v(2)u^{(2)}\to v^{(2)}, …, and u(r)v(r)u^{(r)}\to v^{(r)}. Especially, we denote 𝒫0(𝐮;𝐯)\mathscr{P}_{0}(\mathbf{u};\mathbf{v}) as the set of all non-intersecting paths from 𝐮\mathbf{u} to 𝐯\mathbf{v}.

Let ξ=(P1,P2,,Pr)𝒫(𝐮;𝐯)\xi=(P_{1},P_{2},\cdots,P_{r})\in\mathscr{P}(\mathbf{u};\mathbf{v}) be an rr-path starting from 𝐮\mathbf{u} to 𝐯\mathbf{v}, where PiP_{i} is a path from uiu_{i} to viv_{i}. The weight of ξ\xi is defined to be ω(ξ)=ω(P1,P2,,Pn)=i=1rω(Pi)\omega(\xi)=\omega(P_{1},P_{2},\cdots,P_{n})=\prod_{i=1}^{r}\omega(P_{i}). Moreover, if we count the weight for all paths from 𝐮\mathbf{u} to 𝐯\mathbf{v}, then we have

h(𝐮,𝐯)=ξ𝒫(𝐮;𝐯)ω(ξ).\displaystyle h(\mathbf{u},\mathbf{v})=\sum_{\xi\in\mathscr{P}(\mathbf{u};\mathbf{v})}\omega(\xi).

In general, we refer this weight function as the generating function of all paths from 𝐮\mathbf{u} to 𝐯\mathbf{v}, and denote it as GF[𝒫(𝐮;𝐯)][\mathscr{P}(\mathbf{u};\mathbf{v})]. For non-intersecting paths, we have GF[𝒫0(𝐮;𝐯)]=ξ𝒫0(𝐮;𝐯)ω(ξ)[\mathscr{P}_{0}(\mathbf{u};\mathbf{v})]=\sum_{\xi\in\mathscr{P}_{0}(\mathbf{u};\mathbf{v})}\omega(\xi).

Definition 2.4 (DD-Compatible).

If II and JJ are ordered sets of vertices in a graph DD, then II is said to be D-compatible with JJ in the graph if and only if for any u<uu<u^{\prime} in II and any v>vv>v^{\prime} in JJ, every path P𝒫(u;v)P\in\mathscr{P}(u;v) intersects with every path Q𝒫(u;v)Q\in\mathscr{P}\left(u^{\prime};v^{\prime}\right).

Now, we could formally state the LGV lemma.

Theorem 2.5 ([24, 11, 31]).

Let 𝐮=(u(1),u(2),,u(r))\mathbf{u}=\left(u^{(1)},u^{(2)},\ldots,u^{(r)}\right) and 𝐯=(v(1),v(2),,v(r))\mathbf{v}=\left(v^{(1)},v^{(2)},\ldots,v^{(r)}\right) be two ordered rr-tuples of vertices in an acyclic directed graph DD. If 𝐮\mathbf{u} and 𝐯\mathbf{v} are D-compatible, then

GF[𝒫0(𝐮;𝐯)]=det[h(u(i),v(j))]1i,jr.\operatorname{GF}\left[\mathscr{P}_{0}(\mathbf{u};\mathbf{v})\right]=\operatorname{det}\left[h\left(u^{(i)},v^{(j)}\right)\right]_{1\leqslant i,j\leqslant r}. (2.1)

In literatures, there have been many applications of Lindstörm-Gessel-Vionnet lemma. For example, it has been applied to some tiling problem and random growth models [17, 8]. Originally, non-intersecting paths arose in matroid theory [24], which was later used to count tableaux and plane partitions [11]. In the references mentioned above, there is a specialized binomial determinant related to Toeplitz (respectively Hankel) determinant and orthogonal polynomials on the unit circle (respectively real line). Here we start with a non-intersecting random walk model and demonstrate how LGV lemma works in the random walk models and orthogonal polynomial theory.

Let’s consider NN independent simple random walks X1(t),,XN(t)X_{1}(t),\cdots,X_{N}(t) starting from positions 𝐚=(α1,,αN)\mathbf{a}=(\alpha_{1},\cdots,\alpha_{N}) at t=0t=0, and ending at positions 𝐛=(β1,,βN)\mathbf{b}=(\beta_{1},\cdots,\beta_{N}) at t=2Tt=2T with conditions α1<<αN\alpha_{1}<\cdots<\alpha_{N} and β1<<βN\beta_{1}<\cdots<\beta_{N}. There are two different models related to this setting. One is a continuous-time model called non-intersecting Brownian motion with configuration ×0\mathbb{R}\times\mathbb{R}_{\geq 0} [1, 9]. We first assume X(t)=(X1(t),,XN(t))X(t)=(X_{1}(t),\cdots,X_{N}(t)) to be NN independent standard Brownian motions with transition probability

p(x,y;t)=12πte(xy)22t,\displaystyle p(x,y;t)=\frac{1}{\sqrt{2\pi t}}e^{-\frac{(x-y)^{2}}{2t}},

which indicates the probability from xx to yy within time tt. Moreover, let’s assume that DN:={x1<<xN}ND_{N}:=\{x_{1}<\cdots<x_{N}\}\subset\mathbb{R}^{N} is a admissible configuration. In this case, non-intersecting means that X(t)DNX(t)\in D_{N} for all t(0,2T)t\in(0,2T). According to the Karlin-McGregor’s theorem [19], the conditional probability density function (PDF) for passing through 𝐱=(x1,,xN)\mathbf{x}=(x_{1},\cdots,x_{N}) at time tt and ending at β\beta, given the starting point α\alpha and the total duration of 2T2T, is expressed by

det(p(αi,xj;t))i,j=1Ndet(p(xj,βi;2Tt))i,j=1N.\displaystyle\operatorname{det}\left(p\left(\alpha_{i},x_{j};t\right)\right)_{i,j=1}^{N}\operatorname{det}\left(p\left(x_{j},\beta_{i};2T-t\right)\right)_{i,j=1}^{N}. (2.2)

This result could be also recognized as an application of LGV lemma, where det(p(αi,xj;t))i,j=1N\operatorname{det}\left(p\left(\alpha_{i},x_{j};t\right)\right)_{i,j=1}^{N} could be viewed as the generating function from 𝐚\mathbf{a} to 𝐱\mathbf{x}, and det(p(xj,βi;2Tt))i,j=1N\operatorname{det}\left(p\left(x_{j},\beta_{i};2T-t\right)\right)_{i,j=1}^{N} could be viewed as the generating function from 𝐱\mathbf{x} to 𝐛\mathbf{b}. To calculate the conditional probability of event 𝒩0\mathcal{N}_{0}, we normalize it by dividing a normalization factor for reaching 𝐛\mathbf{b} from 𝐚\mathbf{a} within the total time of 2T2T. In fact, this normalization factor is given by j=1Np(αjβj;2T)\prod_{j=1}^{N}p(\alpha_{j}-\beta_{j};2T), where p(x;T)p(x;T) is a normal distribution with mean zero and variance TT. Therefore, the conditional probability of the event 𝒩0\mathcal{N}_{0} could expressed by

α,β(𝒩0)=det[p(αjβk;2T)]j,k=1Nj=1Np(αjβj;2T).\mathbb{P}_{\alpha,\beta}(\mathcal{N}_{0})=\frac{\operatorname{det}\left[p(\alpha_{j}-\beta_{k};2T)\right]_{j,k=1}^{N}}{\prod_{j=1}^{N}p(\alpha_{j}-\beta_{j};2T)}.

In fact, the PDF (2.2) induces a bi-orthogonal system. We demonstrate this result in the following proposition.

Proposition 2.6.

If we denote ψi(x)=p(αi,x;t)\psi_{i}(x)=p(\alpha_{i},x;t) and ϕj(x)=p(x,βj;2Tt)\phi_{j}(x)=p(x,\beta_{j};2T-t), then we could introduce moments as

mαi,βj=ψi(x)ϕj(x)𝑑x.\displaystyle m_{\alpha_{i},\beta_{j}}=\int_{\mathbb{R}}\psi_{i}(x)\phi_{j}(x)dx.

Moreover, functions

Pn(x)=|mα1,β1mα1,β2mα1,βnmα2,β1mα2,β2mα2,βnmαn1,β1mαn1,β2mαn1,βnϕ1(x)ϕ2(x)ϕn(x)|\displaystyle P_{n}(x)=\left|\begin{array}[]{cccc}m_{\alpha_{1},\beta_{1}}&m_{\alpha_{1},\beta_{2}}&\cdots&m_{\alpha_{1},\beta_{n}}\\ m_{\alpha_{2},\beta_{1}}&m_{\alpha_{2},\beta_{2}}&\cdots&m_{\alpha_{2},\beta_{n}}\\ \vdots&\vdots&&\vdots\\ m_{\alpha_{n-1},\beta_{1}}&m_{\alpha_{n-1},\beta_{2}}&\cdots&m_{\alpha_{n-1},\beta_{n}}\\ \phi_{1}(x)&\phi_{2}(x)&\cdots&\phi_{n}(x)\end{array}\right|

and

Qn(x)=|mα1,β1mα1,βn1ψ1(x)mα2,β1mα2,βn1ψ2(x)mαn,β1mαn,βn1ψn(x)|\displaystyle Q_{n}(x)=\left|\begin{array}[]{cccc}m_{\alpha_{1},\beta_{1}}&\cdots&m_{\alpha_{1},\beta_{n-1}}&\psi_{1}(x)\\ m_{\alpha_{2},\beta_{1}}&\cdots&m_{\alpha_{2},\beta_{n-1}}&\psi_{2}(x)\\ \vdots&\vdots&&\vdots\\ m_{\alpha_{n},\beta_{1}}&\cdots&m_{\alpha_{n},\beta_{n-1}}&\psi_{n}(x)\\ \end{array}\right|

are orthogonal to each other with the uniform measure. In other words, we have

Pn(x)Qm(x)𝑑x=Hnδn,m,\displaystyle\int_{\mathbb{R}}P_{n}(x)Q_{m}(x)dx=H_{n}\delta_{n,m},

where

Hn=det[p(αjβk;2T)]j,k=1n1det[p(αjβk;2T)]j,k=1n.\displaystyle H_{n}=\det[p(\alpha_{j}-\beta_{k};2T)]_{j,k=1}^{n-1}\det[p(\alpha_{j}-\beta_{k};2T)]_{j,k=1}^{n}.

This is a general bi-orthogonal system [2]. If we consider confluent starting and ending point, then we could define multiple orthogonal polynomials and multiple orthogonal polynomials of mixed type [7]. To be precise, let’s consider a confluent case where NN non-intersecting Brownian motions start from kk different points αj\alpha_{j} with multiplicity aja_{j} for j=1,,kj=1,\cdots,k, and end at ll different points βj\beta_{j} with multiplicity bjb_{j} for j=1,,lj=1,\cdots,l, such that j=1kaj=j=1lbj=N\sum_{j=1}^{k}a_{j}=\sum_{j=1}^{l}b_{j}=N. Under this circumstance, we have the following proposition for multiple orthogonal polynomials of mixed type.

Proposition 2.7.

Let’s denote ψi(j)(x)=xiω1,j(x)\psi_{i}^{(j)}(x)=x^{i}\omega_{1,j}(x) for i=0,,aj1i=0,\cdots,a_{j}-1 and j=1,,kj=1,\cdots,k and ϕi(j)(x)=xiω2,j(x)\phi_{i}^{(j)}(x)=x^{i}\omega_{2,j}(x) for i=0,,bj1i=0,\cdots,b_{j}-1 and j=1,,lj=1,\cdots,l, where ω1,j(x)=p(αj,x;t)\omega_{1,j}(x)=p(\alpha_{j},x;t) and ω2,j(x)=p(x,βj;2Tt)\omega_{2,j}(x)=p(x,\beta_{j};2T-t). We define moments of mixed type by

mp+q(i,j)=xp+qω1,i(x)ω2,j(x)𝑑x,p=0,,ai1,q=0,,bj1\displaystyle m_{p+q}^{(i,j)}=\int_{\mathbb{R}}x^{p+q}\omega_{1,i}(x)\omega_{2,j}(x)dx,\quad p=0,\cdots,a_{i}-1,\,q=0,\cdots,b_{j}-1

for i=1,,k,j=1,,l.i=1,\cdots,k,\,j=1,\cdots,l. In this case, multiple orthogonal polynomials of mixed type could be written as

Paek,b(x)=|Aa1,b1(1,1)Aa1,bl(1,l)Aak1,b1(k,1)Aak1,bl(k,l)Φ1(x)Φl(x)|,Qa,bel(x)=|Aa1,b1(1,1)Aa1,bl1(1,l)Ψ1(x)Aak,b1(k,1)Aak,bl1(k,l)Ψk(x)|,\displaystyle P_{\vec{a}-e_{k},\vec{b}}(x)=\left|\begin{array}[]{ccc}A_{a_{1},b_{1}}^{(1,1)}&\cdots&A_{a_{1},b_{l}}^{(1,l)}\\ \vdots&&\vdots\\ A_{a_{k}-1,b_{1}}^{(k,1)}&\cdots&A_{a_{k}-1,b_{l}}^{(k,l)}\\ \Phi_{1}(x)&\cdots&\Phi_{l}(x)\end{array}\right|,\quad Q_{\vec{a},\vec{b}-e_{l}}(x)=\left|\begin{array}[]{cccc}A_{a_{1},b_{1}}^{(1,1)}&\cdots&A^{(1,l)}_{a_{1},b_{l}-1}&\Psi^{\top}_{1}(x)\\ \vdots&&\vdots&\vdots\\ A_{a_{k},b_{1}}^{(k,1)}&\cdots&A^{(k,l)}_{a_{k},b_{l}-1}&\Psi^{\top}_{k}(x)\end{array}\right|,

where

Ψj(x)=(1,x,,xaj1)ω1,j(x),Φj(x)=(1,x,,xbj1)ω2,j(x),Aai,bj(i,j)=(mp+q(i,j))p=0,,ai1q=0,,bj1.\displaystyle\Psi_{j}(x)=(1,x,\cdots,x^{a_{j}-1})\omega_{1,j}(x),\,\Phi_{j}(x)=(1,x,\cdots,x^{b_{j}-1})\omega_{2,j}(x),\,A_{a_{i},b_{j}}^{(i,j)}=(m_{p+q}^{(i,j)})_{p=0,\cdots,a_{i}-1\atop q=0,\cdots,b_{j}-1}.

Moreover, these polynomials satisfy the following mixed orthogonality

Pa,b(x)xpω1,j(x)𝑑x=0,p=0,,aj1,j=1,k,\displaystyle\int_{\mathbb{R}}P_{\vec{a},\vec{b}}(x)x^{p}\omega_{1,j}(x)dx=0,\quad p=0,\cdots,a_{j}-1,\,j=1\cdots,k,
Qa,b(x)xpω2,j(x)𝑑x=0,p=0,,bj1,j=1,l.\displaystyle\int_{\mathbb{R}}Q_{\vec{a},\vec{b}}(x)x^{p}\omega_{2,j}(x)dx=0,\quad p=0,\cdots,b_{j}-1,\,j=1\cdots,l.

Another application is to consider a discrete-time random walk, defined in the configuration space ×0\mathbb{Z}\times\mathbb{Z}_{\geq 0}. At each discrete time step, the simple random walk increase or decrease by one step with equal probability. If we denote αj=2xj\alpha_{j}=2x_{j} and βj=2yj\beta_{j}=2y_{j} for all j=1,2,,Nj=1,2,\cdots,N, and assume that all paths don’t intersect, then the number of all non-intersecting paths is given by a binomial determinant [20, 10]

det[(2TT+ykxj)]j,k=1N.\displaystyle\det\left[2T\choose T+y_{k}-x_{j}\right]_{j,k=1}^{N}.

Such a model is sometimes referred to as vicious random walkers which means that the walkers can not be in the same position at each discrete step. It was shown in [10] that this binomial determinant is related to symmetric Hahn polynomials. However, the following proposition demonstrates that such a binomial determinant could also be related to a Toeplitz determinant and orthogonal polynomials on the unit circle (OPUC).

Proposition 2.8.

([12] with a correction) The binomial coefficient could be written by a residue formula

(2TT+ykxj)=𝕋ω(z)z(ykxj)dz2πiz:=mykxj,ω(z)=(1+z)2TzT,\displaystyle{2T\choose T+y_{k}-x_{j}}=\int_{\mathbb{T}}\omega(z)z^{-(y_{k}-x_{j})}\frac{dz}{2\pi iz}:=m_{y_{k}-x_{j}},\quad\omega(z)=(1+z)^{2T}z^{-T},

where 𝕋\mathbb{T} denotes a unit circle. Moreover, if xj=yj=jx_{j}=y_{j}=j for j=1,2,,Nj=1,2,\cdots,N, then the binomial determinant could be expressed as a Toeplitz determinant det(mkj)j,k=1N\det(m_{k-j})_{j,k=1}^{N}.

Similar to the Proposition 2.6, if we define

Pn(x)=|m0m1mnm1m0mn1mn+1mn+2m11zzn|,\displaystyle P_{n}(x)=\left|\begin{array}[]{cccc}m_{0}&m_{1}&\cdots&m_{n}\\ m_{-1}&m_{0}&\cdots&m_{n-1}\\ \vdots&\vdots&&\vdots\\ m_{-n+1}&m_{-n+2}&\cdots&m_{1}\\ 1&z&\cdots&z^{n}\end{array}\right|,

then one could obtain the following orthogonal relation

𝕋ω(z)Pn(z)P¯m(z¯)dz2πiz=Hnδn,m,\displaystyle\int_{\mathbb{T}}\omega(z)P_{n}(z)\bar{P}_{m}(\bar{z})\frac{dz}{2\pi iz}=H_{n}\delta_{n,m},

where z¯\bar{z} is the complex conjugate of zz and Hn=det(mkj)k,j=0n1det(mkj)k,j=0n.H_{n}=\det(m_{k-j})_{k,j=0}^{n-1}\det(m_{k-j})_{k,j=0}^{n}. Regarding with general {xj,yj}\{x_{j},y_{j}\} and slat Toeplitz determinants, one could refer to [12, 13] for more examples.

3. Stembridge’s results on Pfaffians and applications into skew-orthogonal polynomials

In [31], Stembridge generalized the result of Lindström-Gessel-Viennot, by considering non-intersecting paths from a set of vertices to an ordered subset IVI\subset V, whose number of vertices is indefinite. In this section, let’s assume that 𝐮=(u(1),u(2),,u(r))\mathbf{u}=\left(u^{(1)},u^{(2)},\ldots,u^{(r)}\right) is a finite sequence of vertices in VV.

Definition 3.1 (𝒫(u;I)\mathscr{P}({u};I), 𝒫(𝐮;I)\mathscr{P}(\mathbf{u};I), 𝒫0(𝐮;I)\mathscr{P}_{0}(\mathbf{u};I)).

Let 𝒫(u;I)\mathscr{P}({u};I) be the set of all paths from uu to any vIv\in I, 𝒫(𝐮;I)\mathscr{P}(\mathbf{u};I) be the r-tuple set of paths (P1,,Pr)(P_{1},\cdots,P_{r}) where Pi𝒫(u(i),I)P_{i}\in\mathscr{P}(u^{(i)},I), and 𝒫0(𝐮;I)\mathscr{P}_{0}(\mathbf{u};I) be the non-intersecting paths in 𝒫(𝐮;I)\mathscr{P}(\mathbf{u};I).

Definition 3.2 (QI(𝐮)Q_{I}(\mathbf{u})).

QI(𝐮)Q_{I}(\mathbf{u}) refers to the generating function of all non-intersecting paths from 𝐮\mathbf{u} to II in the graph DD. Literally, we can write QI(𝐮)=QI(u(1),u(2),,u(r))=GF[𝒫0(𝐮;I)]Q_{I}(\mathbf{u})=Q_{I}\left(u^{(1)},u^{(2)},\ldots,u^{(r)}\right)=\operatorname{GF}\left[\mathscr{P}_{0}(\mathbf{u};I)\right].

Especially, if 𝐮={u}\mathbf{u}=\{u\}, which is a single vertex, then

QI(u)=vIh(u,v).\displaystyle Q_{I}(u)=\sum_{v\in I}h(u,v). (3.1)

If 𝐮=(u(1),u(2))\mathbf{u}=(u^{(1)},u^{(2)}), then according to LGV theorem, we have

QI(u(1),u(2))=v(1)<v(2)Idet(h(u(1),v(1))h(u(1),v(2))h(u(2),v(1))h(u(2),v(2)))=v(1),v(2)Ih(u(1),v(1))h(u(2),v(2))sgn(v(2)v(1)),\displaystyle\begin{aligned} Q_{I}(u^{(1)},u^{(2)})&=\sum_{v^{(1)}<v^{(2)}\in I}\det\left(\begin{array}[]{cc}h(u^{(1)},v^{(1)})&h(u^{(1)},v^{(2)})\\ h(u^{(2)},v^{(1)})&h(u^{(2)},v^{(2)})\end{array}\right)\\ &=\sum_{v^{(1)},v^{(2)}\in I}h(u^{(1)},v^{(1)})h(u^{(2)},v^{(2)})\text{sgn}(v^{(2)}-v^{(1)}),\end{aligned} (3.2)

where sgn is a sign function defined by

sgn(a)={1,if a>0,0,if a=0,1,if a<0.\displaystyle\text{sgn}(a)=\left\{\begin{array}[]{ll}1,&\text{if $a>0$},\\ 0,&\text{if $a=0$},\\ -1,&\text{if $a<0$}.\end{array}\right.
Remark 3.3.

The generating function for 2-points (3.2) naturally induces a skew-symmetric inner product. If we define

,:[x]×[x]f1(x),f2(x)×f1(x)f2(y)sgn(yx)w(x)w(y)𝑑x𝑑y,\displaystyle\begin{aligned} \langle\cdot,\cdot\rangle:\,&\mathbb{R}[x]\times\mathbb{R}[x]\to\mathbb{R}\\ &\langle f_{1}(x),f_{2}(x)\rangle\mapsto\int_{\mathbb{R}\times\mathbb{R}}f_{1}(x)f_{2}(y)\text{sgn}(y-x)w(x)w(y)dxdy,\end{aligned} (3.3)

then this inner product is skew symmetric, i.e. we have f1(x),f2(x)=f1(x),f2(x)\langle f_{1}(x),f_{2}(x)\rangle=-\langle f_{1}(x),f_{2}(x)\rangle. Moreover, if we take

f1(x)=h(u(1),x),f2(x)=h(u(2),x),w(x)=δxI,\displaystyle f_{1}(x)=h(u^{(1)},x),\quad f_{2}(x)=h(u^{(2)},x),\quad w(x)=\sum\delta_{x\in I},

then we see that QI(u(1),u(2))=f1(x),f2(x).Q_{I}(u^{(1)},u^{(2)})=\langle f_{1}(x),f_{2}(x)\rangle.

It was found by Stembridge that GF[𝒫0(𝐮;I)]\operatorname{GF}[\mathscr{P}_{0}(\mathbf{u};I)] could be written as a Pfaffian, whose elements are given by (3.1) and (3.2). To this end, let’s introduce the concept of 1-factor, and the graphic definition of Pfaffians.

Definition 3.4 (1-Factor).

Let 𝐮=(u(1),u(2),,u(2n))\mathbf{u}=\left(u^{(1)},u^{(2)},\ldots,u^{(2n)}\right). For each perfect matching for these vertices, the set of all pairings is called a 1-factor. Denote the set of all 1-factors for 𝐮\mathbf{u} as (𝐮)\mathscr{F}(\mathbf{u}). Specifically, if the set consists of the first 2n2n consecutive natural numbers, then the set of 1-factors can be simply denoted as 2n\mathscr{F}_{2n}.

For example, for a set 𝐮=(u(1),u(2),u(3),u(4))\mathbf{u}=\left({u^{(1)}},{u^{(2)}},{u^{(3)}},{u^{(4)}}\right), the perfect matching {{u(1),u(2)},{u(3),u(4)}}\left\{\left\{{u^{(1)},u^{(2)}}\right\},\left\{{u^{(3)},u^{(4)}}\right\}\right\} is a 1-factor in (𝐮)\mathscr{F}(\mathbf{u}).

Definition 3.5 (Crossing Number).

For any 1-factor π(𝐮)\pi\in\mathscr{F}(\mathbf{u}) for the set (u(1),u(2),,u(2n))\left(u^{(1)},u^{(2)},\ldots,u^{(2n)}\right), if we align vertices in a straight line with order u(1)<u(2)<<u(2n)u^{(1)}<u^{(2)}<\cdots<u^{(2n)}, and connect the paired ones with curves above the line, then the number of intersections of these curves is defined as the crossing number, denoted by cs(π)\operatorname{cs}(\pi). The sign of π\pi is defined as sgn(π)=(1)cs(π)\operatorname{sgn}(\pi)=(-1)^{\operatorname{cs}(\pi)}.

Definition 3.6 (Definition of Pfaffians).

If A=[ai,j]1i,j2nA=\left[a_{i,j}\right]_{1\leqslant i,j\leqslant 2n} is a skew symmetric matrix of order 2n2n such that ai,j=aj,ia_{i,j}=-a_{j,i}, then we can define the Pfaffian of AA by

Pf(A)=π2nsgn(π)(i,j)πaij.\displaystyle\operatorname{Pf}(A)=\sum_{\pi\in\mathscr{F}_{2n}}\operatorname{sgn}(\pi)\prod_{(i,j)\in\pi}a_{ij}. (3.4)

3.1. The generating function for non-intersecting paths from 𝐮I\mathbf{u}\to I

The following result was given by Stembridge in [31, Thm. 3.1].

Theorem 3.7.

Let 𝐮=(u(1),u(2),,u(2n))\mathbf{u}=\left(u^{(1)},u^{(2)},\ldots,u^{(2n)}\right) be 2n2n-tuple of vertices in an acyclic directed graph DD. If IVI\subset V is a totally ordered subset of vertices such that 𝐮\mathbf{u} is D-compatible with II, then

QI(u(1),u(2),,u(2n))=Pf[QI(u(i),u(j))]1i,j2n,Q_{I}\left(u^{(1)},u^{(2)},\ldots,u^{(2n)}\right)=\operatorname{Pf}\left[Q_{I}(u^{(i)},u^{(j)})\right]_{1\leqslant i,j\leqslant 2n},

where QI(u(i),u(j))Q_{I}(u^{(i)},u^{(j)}) is given by (3.2). Moreover, if 𝐮=(u(1),u(2),,u(2n+1))\mathbf{u}=(u^{(1)},u^{(2)},\cdots,u^{(2n+1)}), then the generating function of 𝐮I\mathbf{u}\to I is given by

QI(u(1),u(2),,u(2n+1))=Pf[QI(u(i),u(j))QI(u(i))QI(u(j))0]1i,j2n+1,\displaystyle Q_{I}\left(u^{(1)},u^{(2)},\cdots,u^{(2n+1)}\right)=\operatorname{Pf}\left[\begin{array}[]{cc}Q_{I}(u^{(i)},u^{(j)})&Q_{I}(u^{(i)})\\ -Q_{I}(u^{(j)})&0\end{array}\right]_{1\leq i,j\leq 2n+1},

where QI(u(i))Q_{I}(u^{(i)}) is given by (3.1) and QI(u(i),u(j))Q_{I}(u^{(i)},u^{(j)}) is given by (3.2).

Remark 3.8.

It should be remarked that if we consider the generating function for non-intersecting paths from JJ to 𝐯\mathbf{v}, then we have the expression

GF[𝒫0(J;𝐯)]=Q~J(v(r),,v(1)),\displaystyle\operatorname{GF}[\mathscr{P}_{0}(J;\mathbf{v})]=\tilde{Q}_{J}(v^{(r)},\cdots,v^{(1)}),

where

Q~J(v)=uJh(u,v),Q~J(v(2),v(1))=u(1),u(2)Jh(u(1),v(1))h(u(2),v(2))sgn(u(2)u(1)),\displaystyle\tilde{Q}_{J}(v)=\sum_{u\in J}h(u,v),\quad\tilde{Q}_{J}(v^{(2)},v^{(1)})=\sum_{u^{(1)},u^{(2)}\in J}h(u^{(1)},v^{(1)})h(u^{(2)},v^{(2)})\text{sgn}(u^{(2)}-u^{(1)}),

and GF[𝒫0(J;𝐯)]\operatorname{GF}[\mathscr{P}_{0}(J;\mathbf{v})] could be computed by above 1-vertex and 2-vertex formulas for any r+r\in\mathbb{N}_{+}.

3.2. The generating function of non-intersecting paths from 𝐮𝐯I\mathbf{u}\to\mathbf{v}\oplus I

In this part, let’s assume 𝐮=(u(1),u(2),,u(r))\mathbf{u}=(u^{(1)},u^{(2)},\cdots,u^{(r)}) and 𝐯=(v(1),v(2),,v(s))\mathbf{v}=(v^{(1)},v^{(2)},\cdots,v^{(s)}) with srs\leq r.

Definition 3.9 (𝐯I\mathbf{v}\oplus I).

𝐯I\mathbf{v}\oplus I is defined as a union of vertices 𝐯\mathbf{v} and II, where 𝐯\mathbf{v} and II are disjoint, and all vertices are arranged such that every v(i)v^{(i)} in 𝐯\mathbf{v} precedes vertices in II.

Definition 3.10 (𝒫0(𝐮;𝐯I)\mathscr{P}_{0}(\mathbf{u};\mathbf{v}\oplus I)).

𝒫0(𝐮;𝐯I)\mathscr{P}_{0}(\mathbf{u};\mathbf{v}\oplus I) is defined as the set of all non-intersecting paths (P1,,Pr)\left(P_{1},\ldots,P_{r}\right), where for 1is1\leq i\leq s, Pi𝒫(u(i);v(i))P_{i}\in\mathscr{P}(u^{(i)};v^{(i)}); while for s+1irs+1\leq i\leq r, Pi𝒫(u(i);I)P_{i}\in\mathscr{P}(u^{(i)};I).

Theorem 3.11 ([31] with a supplement statement).

Let 𝐮=(u(1),u(2),,u(r))\mathbf{u}=\left(u^{(1)},u^{(2)},\ldots,u^{(r)}\right) and 𝐯=(v(1),,v(s))\mathbf{v}=\left(v^{(1)},\ldots,v^{(s)}\right) be sequences of vertices in the acyclic directed graph DD, and suppose II is a totally ordered subset of VV such that 𝐮\mathbf{u} and 𝐯I\mathbf{v}\oplus I are DD-compatible. If r+sr+s is even, then

GF[𝒫0(𝐮;𝐯I)]=Pf[ABB𝟎],\displaystyle\operatorname{GF}\left[\mathscr{P}_{0}({\bf{u}};{\bf{v}}\oplus I)\right]={\mathop{\rm Pf}\nolimits}\left[\begin{array}[]{cc}A&B\\ -B^{\top}&\mathbf{0}\end{array}\right],

where

A=(QI(u(i),u(j)))1i,jr,B=(h(u(i),v(s+1j)))1ir, 1js.\displaystyle A=\left(Q_{I}(u^{(i)},u^{(j)})\right)_{1\leq i,j\leq r},\quad B=\left(h(u^{(i)},v^{(s+1-j)})\right)_{1\leq i\leq r,\,1\leq j\leq s}. (3.5)

Moreover, if r+sr+s is odd, then

GF[𝒫0(𝐮;𝐯I)]=Pf[A^B^B^0],\displaystyle\operatorname{GF}\left[\mathscr{P}_{0}(\mathbf{u};\mathbf{v}\oplus I)\right]=\operatorname{Pf}\left[\begin{array}[]{cc}\hat{A}&\hat{B}\\ -\hat{B}^{\top}&0\end{array}\right],

where

A^=(Aαα0),B^=(B0),α=(QI(u(i)))1ir,\displaystyle\hat{A}=\left(\begin{array}[]{cc}A&\vec{\alpha}\\ -\vec{\alpha}^{\top}&0\end{array}\right),\hat{B}=\left({\begin{array}[]{*{20}{c}}B\\ 0\end{array}}\right),\quad\vec{\alpha}=\left(Q_{I}(u^{(i)})\right)_{1\leq i\leq r},

and matrices AA and BB are given by (3.5).

Proof.

Here we give a proof when r+sr+s is odd, which is omitted in Stembridge’s paper. We assume that there is an auxiliary vertex u(r+1)u^{(r+1)} added in the graph DD. In order to satisfy DD-compatibility, we might as well assume that this vertex follows all vertices in 𝐮\mathbf{u} and II. Furthermore, we require that weights of u(r+1)u^{(r+1)} to itself is 11, and to other vertices are zero. Let’s denote 𝐮=(u(1),,u(r),u(r+1))\mathbf{u}^{*}=\left(u^{(1)},\ldots,u^{(r)},u^{(r+1)}\right), I=I{u(r+1)}I^{*}=I\cup\left\{u^{(r+1)}\right\}, then one could show that

GF[𝒫0(𝐮;𝐯I)]=GF[𝒫0(𝐮;𝐯I)],\displaystyle\mathrm{GF}\left[\mathscr{P}_{0}(\mathbf{u};\mathbf{v}\oplus I)\right]=\operatorname{GF}\left[\mathscr{P}_{0}\left(\mathbf{u}^{*};\mathbf{v}\oplus I^{*}\right)\right],

which gives the result. ∎

It should be remarked that the Theorem 3.11 comprises previous results in Theorem 2.5 and Theorem 3.7. If 𝐯\mathbf{v} is an empty set, then it degenerates to Theorem 3.7, while if II is an empty set and r=sr=s, then it degenerates to the LGV lemma. Therefore, we want to generalize Stembridge’s result to consider combinatoric explanations for full block Pfaffians. Before that, a comparison to Ishikawa-Wakayama’s result is made to show the strength of Stembridge’s method.

3.3. A comparison to Ishikawa-Wakayama’s result

In [16, Theorem 4.3], Ishikawa and Wakayama obtained similar results to Theorem 3.7, which reads

S(Im)Pf(ASS)πSmsgn(π)GF[𝒫0(uπ,I)]=Pf(HAH).\sum_{S\in{I\choose m}}\operatorname{Pf}\left(A_{S}^{S}\right)\sum_{\pi\in S_{m}}\operatorname{sgn}(\pi)\operatorname{GF}\left[\mathscr{P}_{0}\left(u^{\pi},I\right)\right]=\operatorname{Pf}(HAH^{\top}).

In the above formula, 𝐮=(u(1),,u(m))\mathbf{u}=\left(u^{(1)},\ldots,u^{(m)}\right) is a set of points consisting of an even number of elements, I={V1<<VN}I=\left\{V_{1}<\cdots<V_{N}\right\} is a finite set of even-number vertices, and (Im){I\choose m} represents the set of all subsets of II containing exactly mm elements. Moreover, A=(aViVj)1i<jNA=\left(a_{V_{i}V_{j}}\right)_{1\leq i<j\leq N} is a skew-symmetric matrix and ASSA_{S}^{S} is the submatrix of AA obtained by picking up the rows and columns indexed by S, and H=(h(u(i),Vj))1im,1jNH=\left(h\left(u^{(i)},V_{j}\right)\right)_{\begin{subarray}{c}1\leq i\leq m,1\leq j\leq N\end{subarray}}.

To relax the condition of DD-compatibility, Ishikawa and Wakayama imposed an additional condition that the set of vertices II should be finite and the number of vertices should be even. When 𝐮\mathbf{u} and II are DD-compatible, we can get the same result stated as in Theorem 3.7. Moreover, Ishikawa and Wakayama generalized Theorem 3.11 to [16, Theorem 4.4], where the DD-compatible condition is relaxed. For both (m+n)(m+n) and NN being even, 𝐯=(v(1),v(2),,v(n))\mathbf{v}=\left(v^{(1)},v^{(2)},\ldots,v^{(n)}\right), we have

s(Imn)Pf(ASS)πSmsgn(π)GF[𝒫0(uπ,S0I)]=Pf(HAHHJNJNH𝟎),\sum_{s\in{I\choose m-n}}\operatorname{Pf}\left(A_{S}^{S}\right)\sum_{\pi\in S_{m}}\operatorname{sgn}(\pi)\mathrm{GF}\left[\mathscr{P}_{0}\left(u^{\pi},S^{0}\oplus I\right)\right]=\operatorname{Pf}\left(\begin{array}[]{cc}HAH^{\top}&HJ_{N}\\ -J_{N}H^{\top}&\mathbf{0}\end{array}\right),

where

JN=(001010.100).J_{N}=\left(\begin{array}[]{cccc}0&\ldots&0&1\\ 0&\ldots&1&0\\ \vdots&.&\vdots&\vdots\\ 1&\ldots&0&0\end{array}\right).

Although the results of Stembridge and Ishikawa-Wakayama are both non-intersecting path explanations for Pfaffians, Stembridge derived the general generating function from 𝐮\mathbf{u} to II by using the 1-vertex generating function (3.1) and 2-vertex generating function (3.2), while Ishikawa-Wakayama derived the general generating function by the weight h(u(i),Vj)h\left(u^{(i)},V_{j}\right) between points in 𝐮\mathbf{u} and II, which contains more imformation between 𝐮\mathbf{u} and II. This is the reason why DD-compatibility could be relaxed. Furthermore, Ishikawa-Wakayama’s proof avoided internal pairings between points in 𝐮\mathbf{u}, which made it easier to find a map to offset all intersecting paths. Since we don’t want to make any constraints on the terminal interval II, we still adopt Stembridge’s method in the following sections.

3.4. Application of Stembridge’s theorem

There have been numerous application of Stembridge’s graphic explanation for Pfaffians. In [31, 27], it was used to count skew Young tableaux and the enumeration of plane partitions, and later it was used to demonstrate a minor summation formula for Pfaffian in [16]. In recent years, a fusion of combinatorial technique and statistical physics leads to more applications of Stembridge’s result. For example, Theorem 3.7 was used in a vicious random walker problem without return [10], and Theorem 3.11 was used in a free-boundary lozenge tiling problem [4, Section 4] where some triangular holes in the lattices were imposed. In this part, we demonstrate how to introduce skew-orthogonal polynomials by considering non-intersecting Brownian motions and applications of Stembridge’s results.

Let’s first consider NN independent simple random walks X(t)=(X1(t),,XN(t))X(t)=(X_{1}(t),\cdots,X_{N}(t)) starting from α=(α1,,αN)\vec{\alpha}=(\alpha_{1},\cdots,\alpha_{N}) as in Section 2. If we consider a continuous-time non-intersecting Brownian motion in the configuration space ×0\mathbb{R}\times\mathbb{R}_{\geq 0}, then it is known that at an arbitrary time TT, the distribution of arrival points at x=(x1,,xN)DN\vec{x}=(x_{1},\cdots,x_{N})\in D_{N} could be formulated by

ρT(x1,,xN)=1ZNdet(p(αj,xk;T))j,k=1N\displaystyle\rho_{T}(x_{1},\cdots,x_{N})=\frac{1}{Z_{N}}\det\left(p(\alpha_{j},x_{k};T)\right)_{j,k=1}^{N} (3.6)

with a normalization factor ZNZ_{N}. If N=2nN=2n is even, then Z2nZ_{2n} could be written by

Z2n=x1<<x2ndet(p(αj,xk;T))j,k=1Ndx=Pf[Q(αj,αk)]j,k=12n,\displaystyle Z_{2n}=\int_{x_{1}<\cdots<x_{2n}}\det\left(p(\alpha_{j},x_{k};T)\right)_{j,k=1}^{N}d\vec{x}=\operatorname{Pf}\left[Q_{\mathbb{R}}(\alpha_{j},\alpha_{k})\right]_{j,k=1}^{2n},

where

Q(αj,αk)=2p(αj,x;T)sgn(yx)p(αk,y;T)𝑑x𝑑y.\displaystyle Q_{\mathbb{R}}(\alpha_{j},\alpha_{k})=\int_{\mathbb{R}^{2}}p(\alpha_{j},x;T)\text{sgn}(y-x)p(\alpha_{k},y;T)dxdy.

Moreover, if N=2n+1N=2n+1 is odd, then

Z2n+1=x1<<x2n+1det(p(αj,xk;T))j,k=12N+1dx=Pf[Q(αj,αk)Q(αj)Q(αk)0]j,k=12n+1\displaystyle Z_{2n+1}=\int_{x_{1}<\cdots<x_{2n+1}}\det\left(p(\alpha_{j},x_{k};T)\right)_{j,k=1}^{2N+1}d\vec{x}=\operatorname{Pf}\left[\begin{array}[]{cc}Q_{\mathbb{R}}(\alpha_{j},\alpha_{k})&Q_{\mathbb{R}}(\alpha_{j})\\ Q_{\mathbb{R}}(\alpha_{k})&0\end{array}\right]_{j,k=1}^{2n+1}

with Q(αj)=p(αj,x;T)𝑑xQ_{\mathbb{R}}(\alpha_{j})=\int_{\mathbb{R}}p(\alpha_{j},x;T)dx. It should be noted that the element in this Pfaffian coincide with the formula in the skew-symmetric inner product (3.3). Moreover, we have the following proposition.

Proposition 3.12.

Let’s define monic functions

P2n(x)\displaystyle P_{2n}(x) =1Z2nPf[Q2nv2n+1Ψ2n(x)v2n+10ψ2n+1(x)Ψ2n(x)ψ2n+1(x)0],\displaystyle=\frac{1}{Z_{2n}}\operatorname{Pf}\left[\begin{array}[]{ccc}Q_{2n}&v_{2n+1}&\Psi_{2n}(x)\\ -v_{2n+1}^{\top}&0&\psi_{2n+1}(x)\\ -\Psi^{\top}_{2n}(x)&-\psi_{2n+1}(x)&0\end{array}\right],
P2n+1(x)\displaystyle P_{2n+1}(x) =1Z2nPf[Q2nΨ2n(x)v2n+2Ψ2n(x)0ψ2n+2(x)v2n+2ψ2n+2(x)0],\displaystyle=\frac{1}{Z_{2n}}\operatorname{Pf}\left[\begin{array}[]{ccc}Q_{2n}&\Psi_{2n}(x)&v_{2n+2}\\ -\Psi_{2n}^{\top}(x)&0&\psi_{2n+2}(x)\\ -v_{2n+2}^{\top}&-\psi_{2n+2}(x)&0\end{array}\right],

where Q2n=(Q(αj,αk))j,k=12nQ_{2n}=(Q_{\mathbb{R}}(\alpha_{j},\alpha_{k}))_{j,k=1}^{2n}, Ψ2n=(p(α1,x;T),,p(α2n,x;T))\Psi_{2n}=(p(\alpha_{1},x;T),\cdots,p(\alpha_{2n},x;T))^{\top}, and v2n+1v_{2n+1} and v2n+2v_{2n+2} are two column vectors admitting the forms vi=(Q(α1,αi),,Q(α2n,αi))v_{i}=(Q_{\mathbb{R}}(\alpha_{1},\alpha_{i}),\cdots,Q_{\mathbb{R}}(\alpha_{2n},\alpha_{i}))^{\top} for i=2n+1, 2n+2i=2n+1,\,2n+2. Moreover, {Pk(x)}k\{P_{k}(x)\}_{k\in\mathbb{N}} satisfy skew orthogonality relations

2P2n(x)sgn(yx)P2m(y)𝑑x𝑑y=2P2n+1(x)sgn(yx)P2m+1(y)𝑑x𝑑y=0,\displaystyle\int_{\mathbb{R}^{2}}P_{2n}(x)\text{sgn}(y-x)P_{2m}(y)dxdy=\int_{\mathbb{R}^{2}}P_{2n+1}(x)\text{sgn}(y-x)P_{2m+1}(y)dxdy=0,
2P2n(x)sgn(yx)P2m+1(y)𝑑x𝑑y=Z2n+2Z2nδn,m.\displaystyle\int_{\mathbb{R}^{2}}P_{2n}(x)\text{sgn}(y-x)P_{2m+1}(y)dxdy=\frac{Z_{2n+2}}{Z_{2n}}\delta_{n,m}.

It should be remarked that there are only Z2nZ_{2n} involved under the frame of skew-orthogonal polynomials. To have a clear understanding of Z2n+1Z_{2n+1}, partial-skew-orthogonal polynomials should be introduced and we have the following proposition.

Proposition 3.13.

If we define monic functions

R2n(x)\displaystyle R_{2n}(x) =1Z2nPf[Q2n+1Ψ2n+1(x)ψ2n+1(x)0],\displaystyle=\frac{1}{Z_{2n}}\operatorname{Pf}\left[\begin{array}[]{cc}Q_{2n+1}&\Psi_{2n+1}(x)\\ -\psi^{\top}_{2n+1}(x)&0\end{array}\right],
R2n+1(x)\displaystyle R_{2n+1}(x) =1Z2n+1Pf[0w2n+20w2n+2Q2n+2Ψ2n+2(x)0Ψ2n+2(x)0],\displaystyle=\frac{1}{Z_{2n+1}}\operatorname{Pf}\left[\begin{array}[]{ccc}0&w_{2n+2}&0\\ -w_{2n+2}^{\top}&Q_{2n+2}&\Psi_{2n+2}(x)\\ 0&-\Psi_{2n+2}(x)&0\end{array}\right],

where Q2n+2=(Q(αj,αk))j,k=12n+2Q_{2n+2}=(Q_{\mathbb{R}}(\alpha_{j},\alpha_{k}))_{j,k=1}^{2n+2}, Ψ2n+2(x)=(p(α1,x;T),,p(α2n+2,x;T))\Psi_{2n+2}(x)=(p(\alpha_{1},x;T),\cdots,p(\alpha_{2n+2},x;T))^{\top} and w2n+2=(Q(α1),,Q(α2n+2))w_{2n+2}=(Q_{\mathbb{R}}(\alpha_{1}),\cdots,Q_{\mathbb{R}}(\alpha_{2n+2})). Moreover, {Rk(x)}k\{R_{k}(x)\}_{k\in\mathbb{N}} are determined by the following relations

2R2n(x)sgn(yx)p(αm,x;T)𝑑x𝑑y=Z2n+2Z2nδ2n+1,m,for m2n+1,\displaystyle\int_{\mathbb{R}^{2}}R_{2n}(x)\text{sgn}(y-x)p(\alpha_{m},x;T)dxdy=\frac{Z_{2n+2}}{Z_{2n}}\delta_{2n+1,m},\qquad\qquad\text{for $m\leq 2n+1$},
2R2n+1(x)sgn(yx)p(αm,x;T)𝑑x𝑑y=Z2n+2Z2n+1Q(αm),for m2n+1.\displaystyle\int_{\mathbb{R}^{2}}R_{2n+1}(x)\text{sgn}(y-x)p(\alpha_{m},x;T)dxdy=-\frac{Z_{2n+2}}{Z_{2n+1}}Q_{\mathbb{R}}(\alpha_{m}),\qquad\text{for $m\leq 2n+1$}.

4. Non-intersecting path explanations for block Pfaffians and applications

In this section, we generalize Stembridge’s theorem and obtain a block matrix representation for generating function of non-intersecting paths. We assume that these paths should start from and end at several different sets of vertices.

4.1. Pfaffian representations for J𝐮𝐯IJ\oplus\mathbf{u}\to\mathbf{v}\oplus I

To give a Pfaffian representation for paths J𝐮𝐯IJ\oplus\mathbf{u}\to\mathbf{v}\oplus I, we need to introduce the separation of a graph.

Definition 4.1 (DD-separated).

We call vertex sets J𝐮J\oplus\mathbf{u} and 𝐯I\mathbf{v}\oplus I to be DD-separated if paths from JJ to 𝐯\mathbf{v} and those from 𝐮\mathbf{u} to II are non-intersecting.

Definition 4.2 (𝒫0(J𝐮;𝐯I)\mathscr{P}_{0}(J\oplus\mathbf{u};\mathbf{v}\oplus I)).

If a graph DD is DD-separated, then 𝒫0(J𝐮;𝐯I)\mathscr{P}_{0}(J\oplus\mathbf{u};\mathbf{v}\oplus I) represents all non-intersecting paths J𝐮𝐯IJ\oplus\mathbf{u}\to\mathbf{v}\oplus I without any path from JIJ\to I. Moreover, paths should be consecutive; that is, if there is a path from u(1)v(k)u^{(1)}\to v^{(k)}, then we must have u(2)v(k+1)u^{(2)}\to v^{(k+1)} and so on, until that v(s)v^{(s)} is matched over.

Again, we assume that 𝐮=(u(1),,u(r))\mathbf{u}=(u^{(1)},\cdots,u^{(r)}) and 𝐯=(v(1),,v(s))\mathbf{v}=(v^{(1)},\cdots,v^{(s)}). If r+sr+s is even, then we use 𝒫0ee(J𝐮;𝐯I){}_{e}^{e}\mathscr{P}_{0}(J\oplus\mathbf{u};\mathbf{v}\oplus I) (respectively 𝒫0oo(J𝐮;𝐯I){}_{o}^{o}\mathscr{P}_{0}(J\oplus\mathbf{u};\mathbf{v}\oplus I) ) to denote the number of paths from J𝐯J\to\mathbf{v} and 𝐮I\mathbf{u}\to I are both even (respectively both odd). Here we don’t have the case where the number of paths from J𝐯J\to\mathbf{v} is even, and that from 𝐮I\mathbf{u}\to I is odd. Otherwise there will be some paths from JIJ\to I, which is contradicted with DD-separated . On the other hand, if r+sr+s is odd, then we use 𝒫0eo(J𝐮;𝐯I){}_{e}^{o}\mathscr{P}_{0}(J\oplus\mathbf{u};\mathbf{v}\oplus I) (respectively 𝒫0oe(J𝐮;𝐯I){}_{o}^{e}\mathscr{P}_{0}(J\oplus\mathbf{u};\mathbf{v}\oplus I)) to represent that there are even (respectively odd) paths from 𝐮I\mathbf{u}\to I and odd (respectively even) paths from J𝐯J\to\mathbf{v}. Let’s denote weight matrices

Q=(QI(u(i),u(j)))1i,jr,Q~=(Q~J(v(s+1i),v(s+1j)))1i,js,\displaystyle Q=\left(Q_{I}(u^{(i)},u^{(j)})\right)_{1\leq i,j\leq r},\quad\tilde{Q}=\left(\tilde{Q}_{J}(v^{(s+1-i)},v^{(s+1-j)})\right)_{1\leq i,j\leq s},
H=(h(u(i),v(s+1j)))1ir,1js,\displaystyle H=\left(h(u^{(i)},v^{(s+1-j)})\right)_{1\leq i\leq r,1\leq j\leq s},

and column vectors QI=(QI(u(i)))1irQ_{I}=\left(Q_{I}(u^{(i)})\right)_{1\leq i\leq r} and Q~J=(Q~J(v(j)))1js\tilde{Q}_{J}=(\tilde{Q}_{J}(v^{(j)}))_{1\leq j\leq s}.

Theorem 4.3.

Let 𝐮=(u(1),,u(r))\mathbf{u}=\left(u^{(1)},\ldots,u^{(r)}\right) and 𝐯=(v(1),,v(s))\mathbf{v}=\left(v^{(1)},\ldots,v^{(s)}\right) be sequences of vertices in an acyclic directed graph DD. Assume that II and JJ are totally ordered subsets of VV, such that J𝐮J\oplus\mathbf{u} and 𝐯I\mathbf{v}\oplus I are DD-compatible and DD-separated. If r+sr+s is even, then we have

GF[𝒫0ee(J𝐮;𝐯I)]=Pf[QHHQ~],\displaystyle\operatorname{GF}[{}_{e}^{e}\mathscr{P}_{0}(J\oplus\mathbf{u};\mathbf{v}\oplus I)]=\operatorname{Pf}\left[\begin{array}[]{cc}Q&H\\ -H^{\top}&\tilde{Q}\end{array}\right], (4.3)

and

GF[𝒫0oo(J𝐮;𝐯I)]=Pf[QQIH0QI000H0Q~Q~J00Q~J0].\displaystyle\operatorname{GF}[{}_{o}^{o}\mathscr{P}_{0}(J\oplus\mathbf{u};\mathbf{v}\oplus I)]=\operatorname{Pf}\left[\begin{array}[]{cccc}Q&Q_{I}&H&0\\ -Q_{I}^{\top}&0&0&0\\ -H^{\top}&0&\tilde{Q}&\tilde{Q}_{J}\\ 0&0&-\tilde{Q}_{J}^{\top}&0\end{array}\right]. (4.8)

On the other hand, if r+sr+s is odd, then

GF[𝒫0oe(J𝐮;𝐯I)]=Pf[QH0HQ~Q~J0Q~J0],\displaystyle\operatorname{GF}[{}_{o}^{e}\mathscr{P}_{0}(J\oplus\mathbf{u};\mathbf{v}\oplus I)]=\operatorname{Pf}\left[\begin{array}[]{ccc}Q&H&0\\ -H^{\top}&\tilde{Q}&\tilde{Q}_{J}\\ 0&-\tilde{Q}_{J}^{\top}&0\end{array}\right], (4.12)

and

GF[𝒫0eo(J𝐮;𝐯I)]=Pf[QQIHQIQ~0H00].\displaystyle\operatorname{GF}[{}_{e}^{o}\mathscr{P}_{0}(J\oplus\mathbf{u};\mathbf{v}\oplus I)]=\operatorname{Pf}\left[\begin{array}[]{ccc}Q&Q_{I}&H\\ -Q_{I}^{\top}&\tilde{Q}&0\\ -H^{\top}&0&0\end{array}\right]. (4.16)
Proof.

Let’s prove the equation (4.3), and (4.8)-(4.16) could be similarly verified. This proof is based on the path switching involution for non-intersecting paths, and is divided into three steps.

(1) Expanding the Pfaffian as a summation of configurations. By expanding (4.3), we have

Pf[QHHQ~]\displaystyle\operatorname{Pf}\left[\begin{array}[]{cc}Q&H\\ -H^{\top}&\tilde{Q}\end{array}\right] =πsgn(π)×\displaystyle=\sum_{\pi}\text{sgn}(\pi)\times
(u(i),u(j))πQI(u(i),u(j))(u(i),v(j))πh(u(i),v(j))(v(i),v(j))πQ~J(v(j),v(i)),\displaystyle\quad\prod_{(u^{(i)},u^{(j)})\in\pi}Q_{I}(u^{(i)},u^{(j)})\prod_{(u^{(i)},v^{(j)})\in\pi}h(u^{(i)},v^{(j)})\prod_{(v^{(i)},v^{(j)})\in\pi}\tilde{Q}_{J}(v^{(j)},v^{(i)}),

where π\pi belongs to the set (u(1),,u(r),v(s),,v(1))\mathscr{F}(u^{(1)},\cdots,u^{(r)},v^{(s)},\cdots,v^{(1)}). We recognize this formula as a product of path weights by rewriting it as

Pf[QHHQ~]=(π,P(1),,P(r),Q(1),,Q(s))sgn(π)ω(P(1),,P(r),Q(1),,Q(s)),\displaystyle\operatorname{Pf}\left[\begin{array}[]{cc}Q&H\\ -H^{\top}&\tilde{Q}\end{array}\right]=\sum_{(\pi,P^{(1)},\cdots,P^{(r)},Q^{(1)},\cdots,Q^{(s)})}\text{sgn}(\pi)\omega(P^{(1)},\cdots,P^{(r)},Q^{(1)},\cdots,Q^{(s)}), (4.19)

where (π,P(1),,P(r),Q(1),,Q(s)){(\pi,P^{(1)},\cdots,P^{(r)},Q^{(1)},\cdots,Q^{(s)})} is a configuration such that

  • If (u(i),v(j))π(u^{(i)},v^{(j)})\in\pi, then P(i)=Q(j)𝒫(u(i),v(j))P^{(i)}=Q^{(j)}\in\mathscr{P}(u^{(i)},v^{(j)});

  • If (u(i),u(j))π(u^{(i)},u^{(j)})\in\pi, then P(i)𝒫(u(i),I)P^{(i)}\in\mathscr{P}(u^{(i)},I), P(j)𝒫(u(j),I)P^{(j)}\in\mathscr{P}(u^{(j)},I), while P(i)P^{(i)} and P(j)P^{(j)} don’t intersect;

  • If (v(i),v(j))π(v^{(i)},v^{(j)})\in\pi, then Q(i)𝒫(J,v(i))Q^{(i)}\in\mathscr{P}(J,v^{(i)}), Q(j)𝒫(J,v(i))Q^{(j)}\in\mathscr{P}(J,v^{(i)}), while Q(i)Q^{(i)} and Q(j)Q^{(j)} don’t intersect.

We call these three conditions as the configuration conditions.

(2) Finding a sign-reversing involution to offset the intersecting paths. To make the RHS in (4.19) represent the summation of weights for all non-intersecting paths, we intend to show that there exists an involution which can offset the intersecting term. Since the graph is DD-separated, we know that if there are paths intersected, then the intersecting point belongs to P(i)P(j)P^{(i)}\cap P^{(j)} or Q(i)Q(j)Q^{(i)}\cap Q^{(j)} for some ii and jj. We show the first case and the latter could be similarly verified. Let’s find the smallest intersecting point ww^{*}111We call an intersecting point the smallest if it is the first intersecting point starting from u(1)u^{(1)}. If there isn’t any intersecting point in P(1)P^{(1)}, then we find it in P(2)P^{(2)}, and so on.(the biggest intersecting point ww^{*} in the latter case). Among the paths passing through ww^{*}, we assume that P(i)P^{(i)} and P(j)P^{(j)} are the two paths which admit two smallest indices. Thus we could construct an involution

ϕ:(π,P(1),,P(r),Q(1),,Q(s))(π¯,P¯(1),,P¯(r),Q¯(1),,Q¯(s)),\displaystyle\phi:(\pi,P^{(1)},\cdots,P^{(r)},Q^{(1)},\cdots,Q^{(s)})\to(\bar{\pi},\bar{P}^{(1)},\cdots,\bar{P}^{(r)},\bar{Q}^{(1)},\cdots,\bar{Q}^{(s)}),

where we define paths P¯(i)\bar{P}^{(i)} and P¯(j)\bar{P}^{(j)} as

P¯(i)=P(i)(w)P(j)(w),P¯(j)=P(j)(w)P(i)(w),\displaystyle\bar{P}^{(i)}=P^{(i)}(\to w^{*})P^{(j)}(w^{*}\to),\quad\bar{P}^{(j)}=P^{(j)}(\to w^{*})P^{(i)}(w^{*}\to),

and the other paths remain the same. At the same time, π¯\bar{\pi} is obtained by interchanging u(i)u^{(i)} and u(j)u^{(j)} in π\pi. Next, we show that (π¯,P¯(1),,P¯(r),Q¯(1),,Q¯(s))(\bar{\pi},\bar{P}^{(1)},\cdots,\bar{P}^{(r)},\bar{Q}^{(1)},\cdots,\bar{Q}^{(s)}) is still a configuration, and ϕ\phi is sign-reversing.

To prove this, we first demonstrate that (π¯,P¯(1),,P¯(r),Q¯(1),,Q¯(s))\left(\bar{\pi},\bar{P}^{(1)},\ldots,\bar{P}^{(r)},\bar{Q}^{(1)},\ldots,\bar{Q}^{(s)}\right) is still in the summation (4.19), which is equivalent to verify that (π¯,P¯(1),,P¯(r),Q¯(1),,Q¯(s))\left(\bar{\pi},\bar{P}^{(1)},\ldots,\bar{P}^{(r)},\bar{Q}^{(1)},\ldots,\bar{Q}^{(s)}\right) satisfies configuration conditions. Since the paths except P(i)P^{(i)} and P(j)P^{(j)} remain the same, it is necessary to consider cases in π¯\bar{\pi} which only contain u(i)u^{(i)} or u(j)u^{(j)}. If (u(i),u(k))π¯\left(u^{(i)},u^{(k)}\right)\in\bar{\pi}, then we have (u(j),u(k))π\left(u^{(j)},u^{(k)}\right)\in\pi, implying that P(j)𝒫(u(j);I),P(k)𝒫(u(k);I)P^{(j)}\in\mathscr{P}\left(u^{(j)};I\right),P^{(k)}\in\mathscr{P}\left(u^{(k)};I\right). Therefore, as π¯\bar{\pi} is obtained by interchanging u(i)u^{(i)} and u(j)u^{(j)} in π\pi, it is known that P¯(i)𝒫(u(i);I)\bar{P}^{(i)}\in\mathscr{P}\left(u^{(i)};I\right) and P¯(k)𝒫(u(k);I)\bar{P}^{(k)}\in\mathscr{P}\left(u^{(k)};I\right) as required. To show that P¯(i)\bar{P}^{(i)} and P¯(k)\bar{P}^{(k)} do not intersect, we verify it by contradiction. Assume that P¯(i)\bar{P}^{(i)} and P¯(k)\bar{P}^{(k)} intersect with each other at some point, then the intersection point could only be behind the smallest ww^{*}. However, since P¯(i)\bar{P}^{(i)} is the same as P(j)P^{(j)} after the point ww^{*}, it means that P(j)P^{(j)} and P(k)P^{(k)} intersect, which leads to a contradiction with (u(j),u(k))π(u^{(j)},u^{(k)})\in\pi. On the other hand, if (u(i),v(k))π¯\left(u^{(i)},v^{(k)}\right)\in\bar{\pi}, then (u(j),v(k))π\left(u^{(j)},v^{(k)}\right)\in\pi, and therefore P(j)𝒫(u(j),v(k))P^{(j)}\in\mathscr{P}\left(u^{(j)},v^{(k)}\right). From this fact, we know that P¯(i)𝒫(u(i),v(k))\bar{P}^{(i)}\in\mathscr{P}\left(u^{(i)},v^{(k)}\right) as required. To show the map is sign-reversing, by making use of DD-compatibility, we know that for any kk satisfying i<k<ji<k<j, P(k)P^{(k)} must intersect with P(i)P^{(i)} and P(j)P^{(j)}. Therefore, by using [31, Lem. 2.1], we can deduce that this involution mapping is sign-reversing.

(3) Simplification of the configuration. After offsetting intersecting terms, only non-intersecting terms are left. Therefore, (4.19) could be rewritten as the non-intersecting part of

 is oddπΛsgn(π)(u(i),u(j))π2QI(u(i),u(j))i=sh(u(i+1),v(i))(v(i),v(j))π3Q~J(v(j),v(i)),\displaystyle\sum_{\text{$\ell$ is odd}}\sum_{\pi\in\Lambda_{\ell}}\text{sgn}(\pi)\prod_{\left(u^{(i)},u^{(j)}\right)\in\pi_{2}^{\ell}}Q_{I}(u^{(i)},u^{(j)})\prod_{i=\ell}^{s}h(u^{(i+1-\ell)},v^{(i)})\prod_{\left(v^{(i)},v^{(j)}\right)\in\pi_{3}^{\ell}}\tilde{Q}_{J}(v^{(j)},v^{(i)}), (4.20)

where

Λ={π(u(1),,u(r),v(s),,v(1))|π=π1π2π3},\displaystyle\Lambda_{\ell}=\{\pi\in\mathscr{F}(u^{(1)},\cdots,u^{(r)},v^{(s)},\cdots,v^{(1)})|\pi=\pi_{1}^{\ell}\cup\pi_{2}^{\ell}\cup\pi_{3}^{\ell}\},
π1=i=s(u(i+1),v(i)),π2(u(s+2),,u(r)),π3(v(1),,v(1)).\displaystyle\pi_{1}^{\ell}=\cup_{i=\ell}^{s}(u^{(i+1-\ell)},v^{(i)}),\quad\pi_{2}^{\ell}\in\mathscr{F}(u^{(s+2-\ell)},\cdots,u^{(r)}),\quad\pi_{3}^{\ell}\in\mathscr{F}(v^{(1)},\cdots,v^{(\ell-1)}).
Refer to caption
Figure 1. A decomposition of 1-factors for pairings in J𝐮𝐯IJ\oplus\mathbf{u}\to\mathbf{v}\oplus I

Moreover, by realizing

πΛlsgn(π)=sgn(π1)(sgn(π2))(sgn(π3))=1\displaystyle\sum_{\pi\in\Lambda_{l}}\operatorname{sgn}(\pi)=\operatorname{sgn}\left(\pi_{1}^{\ell}\right)\left(\sum\operatorname{sgn}\left(\pi_{2}^{\ell}\right)\right)\left(\sum\operatorname{sgn}\left(\pi_{3}^{\ell}\right)\right)=1

we know that (4.20) represents the weight of all non-intersecting paths from J𝐮𝐯IJ\oplus\mathbf{u}\to\mathbf{v}\oplus I, and there are even paths from 𝐮I\mathbf{u}\to I and even paths from J𝐯J\to\mathbf{v}. ∎

Let’s end with an example. If 𝐮=(u(1),u(2))\mathbf{u}=(u^{(1)},u^{(2)}) and 𝐯=(v(1),v(2))\mathbf{v}=(v^{(1)},v^{(2)}), then it is known that

Pf[0QI(u(1),u(2))h(u(1),v(2))h(u(1),v(1))QI(u(1),u(2))0h(u(2),v(2))h(u(2),v(1))h(u(1),v(2))h(u(2),v(2))0Q~J(v(2),v(1))h(u(1),v(1))h(u(2),v(1))Q~J(v(2),v(1))0]\displaystyle\operatorname{Pf}\left[\begin{array}[]{cccc}0&Q_{I}(u^{(1)},u^{(2)})&h(u^{(1)},v^{(2)})&h(u^{(1)},v^{(1)})\\ -Q_{I}(u^{(1)},u^{(2)})&0&h(u^{(2)},v^{(2)})&h(u^{(2)},v^{(1)})\\ -h(u^{(1)},v^{(2)})&-h(u^{(2)},v^{(2)})&0&\tilde{Q}_{J}(v^{(2)},v^{(1)})\\ -h(u^{(1)},v^{(1)})&-h(u^{(2)},v^{(1)})&-\tilde{Q}_{J}(v^{(2)},v^{(1)})&0\end{array}\right]
=QI(u(1),u(2))Q~J(v(2),v(1))h(u(1),v(2))h(u(2),v(1))+h(u(1),v(1))h(u(2),v(2)),\displaystyle\qquad=Q_{I}(u^{(1)},u^{(2)})\tilde{Q}_{J}(v^{(2)},v^{(1)})-h(u^{(1)},v^{(2)})h(u^{(2)},v^{(1)})+h(u^{(1)},v^{(1)})h(u^{(2)},v^{(2)}),

where the first term represents the weight of non-intersecting paths from 𝐮I\mathbf{u}\to I and J𝐯J\to\mathbf{v}, and the rest terms represent the weight of non-intersecting paths from u(1)v(1)u^{(1)}\to v^{(1)} and u(2)v(2)u^{(2)}\to v^{(2)}.

4.2. Applications of block Pfaffians and multiple skew-orthogonal polynomials

In this subsection, we show some applications of this block Pfaffians, which correspond to the multiple skew-orthogonal polynomial theory.

We start with the probability density function given by (3.6). If we consider there are only two starting points named a1a_{1} and a2a_{2}, and there are nin_{i} paths from αi\alpha_{i} for i=1,2i=1,2, then we have a confluent form for the PDF. By noting that

limα1,,αn1a1αn1+1,,αNa2det(p(αi,xj;T))i,j=1Ndet(xjk11p(a1,xj;T)xjk21p(a2,xj;T))j=1,,N,k1=1,,n1,k2=1,,n2,\displaystyle\lim_{\alpha_{1},\cdots,\alpha_{n_{1}}\to a_{1}\atop\alpha_{n_{1}+1},\cdots,\alpha_{N}\to a_{2}}\det(p(\alpha_{i},x_{j};T))_{i,j=1}^{N}\sim\det\left(\begin{array}[]{c}x_{j}^{k_{1}-1}p(a_{1},x_{j};T)\\ x_{j}^{k_{2}-1}p(a_{2},x_{j};T)\end{array}\right)_{j=1,\cdots,N,k_{1}=1,\cdots,n_{1},k_{2}=1,\cdots,n_{2}},

where N=n1+n2N=n_{1}+n_{2}, we know that the normalization factor in this case could be written as

Zn1,n2=x1<<xNdet(xjk11p(a1,xj;T)xjk21p(a2,xj;T))j=1,,N,k1=1,,n1,k2=1,,n2dx=Pf[Q2(u(i),u(j))i,j=1,,n1H(u(i),v(j))i=1,,n1,j=1,,n2H(v(i),u(j))i=1,,n2j=1,,n1Q~2(v(i),v(j))i,j=1,,n2]\displaystyle\begin{aligned} Z_{n_{1},n_{2}}&=\int_{x_{1}<\cdots<x_{N}}\det\left(\begin{array}[]{c}x_{j}^{k_{1}-1}p(a_{1},x_{j};T)\\ x_{j}^{k_{2}-1}p(a_{2},x_{j};T)\end{array}\right)_{j=1,\cdots,N,k_{1}=1,\cdots,n_{1},k_{2}=1,\cdots,n_{2}}d\vec{x}\\ &=\operatorname{Pf}\left[\begin{array}[]{cc}Q_{\mathbb{R}^{2}}(u^{(i)},u^{(j)})_{i,j=1,\cdots,n_{1}}&H(u^{(i)},v^{(j)})_{i=1,\cdots,n_{1},\atop j=1,\cdots,n_{2}}\\ -H(v^{(i)},u^{(j)})_{i=1,\cdots,n_{2}\atop j=1,\cdots,n_{1}}&\tilde{Q}_{\mathbb{R}^{2}}(v^{(i)},v^{(j)})_{i,j=1,\cdots,n_{2}}\end{array}\right]\end{aligned} (4.21)

if n1+n2n_{1}+n_{2} is even. Here we note that the partition function Zn1,n2Z_{n_{1},n_{2}} is in the form of equation (4.3). Moreover, in the above notations, we have

Q2(u(i),u(j))\displaystyle Q_{\mathbb{R}^{2}}(u^{(i)},u^{(j)}) =2xi1p(a1,x;T)sgn(yx)yj1p(a1,y;T)𝑑x𝑑y,\displaystyle=\int_{\mathbb{R}^{2}}x^{i-1}p(a_{1},x;T)\text{sgn}(y-x)y^{j-1}p(a_{1},y;T)dxdy,
Q~2(v(i),v(j))\displaystyle\tilde{Q}_{\mathbb{R}^{2}}(v^{(i)},v^{(j)}) =2xi1p(a2,x;T)sgn(yx)yj1p(a2,y;T)𝑑x𝑑y,\displaystyle=\int_{\mathbb{R}^{2}}x^{i-1}p(a_{2},x;T)\text{sgn}(y-x)y^{j-1}p(a_{2},y;T)dxdy,
H(u(i),v(j))\displaystyle H(u^{(i)},v^{(j)}) =2xi1p(a1,x;T)sgn(yx)yj1p(a2,y;T)𝑑x𝑑y,\displaystyle=\int_{\mathbb{R}^{2}}x^{i-1}p(a_{1},x;T)\text{sgn}(y-x)y^{j-1}p(a_{2},y;T)dxdy,
H(v(i),u(j))\displaystyle H(v^{(i)},u^{(j)}) =2xi1p(a2,x;T)sgn(yx)yj1p(a1,y;T)𝑑x𝑑y.\displaystyle=\int_{\mathbb{R}^{2}}x^{i-1}p(a_{2},x;T)\text{sgn}(y-x)y^{j-1}p(a_{1},y;T)dxdy.

This normalization factor induces the following definition of multiple skew-orthogonal functions.

Proposition 4.4.

Let’s denote mi,j=Q2(u(i),u(j))m_{i,j}=Q_{\mathbb{R}^{2}}(u^{(i)},u^{(j)}), m~i,j=Q~2(v(i),v(j))\tilde{m}_{i,j}=\tilde{Q}_{\mathbb{R}^{2}}(v^{(i)},v^{(j)}), hi,j=H(u(i),v(j))h_{i,j}=H(u^{(i)},v^{(j)}) and ψi,j(x)=xj1p(ai,x;T)\psi_{i,j}(x)=x^{j-1}p(a_{i},x;T). If n1+n2n_{1}+n_{2} is odd, then we could define functions

Rn1,n2(x)=Pf[[mi,j]i,j=1n1[hi,j]i=1,,n1j=1,,n2[ψ1,j(x)]j=1n1[hi,j]i=1,,n2j=1,,n1[m~i,j]i,j=1n2[ψ2,j(x)]j=1n2[ψ1,j(x)]j=1n1[ψ2,j(x)]j=1n20],\displaystyle R_{n_{1},n_{2}}(x)=\operatorname{Pf}\left[\begin{array}[]{ccc}\left[m_{i,j}\right]_{i,j=1}^{n_{1}}&\left[h_{i,j}\right]_{i=1,\cdots,n_{1}\atop j=1,\cdots,n_{2}}&\left[\psi_{1,j}(x)\right]_{j=1}^{n_{1}}\\ -\left[h_{i,j}\right]_{i=1,\cdots,n_{2}\atop j=1,\cdots,n_{1}}&\left[\tilde{m}_{i,j}\right]_{i,j=1}^{n_{2}}&\left[\psi_{2,j}(x)\right]_{j=1}^{n_{2}}\\ -\left[\psi_{1,j}(x)\right]_{j=1}^{n_{1}}&-\left[\psi_{2,j}(x)\right]_{j=1}^{n_{2}}&0\end{array}\right],
Rn1+1,n2(x)=Pf[[mi,j]i,j=1,,n11,n1+1[hi,j]i=1,,n11,n1+1j=1,,n2[ψ1,j(x)]j=1,,n11,n1+1[hi,j]i=1,,n2j=1,,n11,n1+1[m~i,j]i,j=1n2[ψ2,j(x)]j=1n2[ψ1,j(x)]j=1,,n11,n1+1[ψ2,j(x)]j=1n20],\displaystyle R_{n_{1}+1,n_{2}}(x)=\operatorname{Pf}\left[\begin{array}[]{ccc}\left[m_{i,j}\right]_{i,j=1,\cdots,n_{1}-1,n_{1}+1}&\left[h_{i,j}\right]_{i=1,\cdots,n_{1}-1,n_{1}+1\atop j=1,\cdots,n_{2}}&\left[\psi_{1,j}(x)\right]_{j=1,\cdots,n_{1}-1,n_{1}+1}\\ -[h_{i,j}]_{i=1,\cdots,n_{2}\atop j=1,\cdots,n_{1}-1,n_{1}+1}&\left[\tilde{m}_{i,j}\right]_{i,j=1}^{n_{2}}&\left[\psi_{2,j}(x)\right]_{j=1}^{n_{2}}\\ -\left[\psi_{1,j}(x)\right]_{j=1,\cdots,n_{1}-1,n_{1}+1}&-\left[\psi_{2,j}(x)\right]_{j=1}^{n_{2}}&0\end{array}\right],
Rn1,n2+1(x)=Pf[[mi,j]i,j=1n1[hi,j]i=1,,n1,j=1,,n21,n2+1[ψ1,j(x)]j=1n1[hi,j]i=1,,n21,n2+1j=1,,n1[m~i,j]i,j=1,,n21,n2+1[ψ2,j(x)]j=1,,n21,n2+1[ψ1,j(x)]j=1n1[ψ2,j(x)]j=1,,n21,n2+10],\displaystyle R_{n_{1},n_{2}+1}(x)=\operatorname{Pf}\left[\begin{array}[]{ccc}\left[m_{i,j}\right]_{i,j=1}^{n_{1}}&\left[h_{i,j}\right]_{i=1,\cdots,n_{1},\atop j=1,\cdots,n_{2}-1,n_{2}+1}&\left[\psi_{1,j}(x)\right]_{j=1}^{n_{1}}\\ -\left[h_{i,j}\right]_{i=1,\cdots,n_{2}-1,n_{2}+1\atop j=1,\cdots,n_{1}}&\left[\tilde{m}_{i,j}\right]_{i,j=1,\cdots,n_{2}-1,n_{2}+1}&\left[\psi_{2,j}(x)\right]_{j=1,\cdots,n_{2}-1,n_{2}+1}\\ -\left[\psi_{1,j}(x)\right]_{j=1}^{n_{1}}&-\left[\psi_{2,j}(x)\right]_{j=1,\cdots,n_{2}-1,n_{2}+1}&0\end{array}\right],

and they satisfy the following multiple skew orthogonality 222Here n=(n1,n2)\vec{n}=(n_{1},n_{2}) and eie_{i} is the unit vector whose iith element is 11, and the others are 0. For example, n+e1=(n1+1,n2)\vec{n}+e_{1}=(n_{1}+1,n_{2}).

2Rn1,n2(x)sgn(yx)yj1p(y,ai;T)𝑑x𝑑y=Zn+eiδj,ni+1,\displaystyle\int_{\mathbb{R}^{2}}R_{n_{1},n_{2}}(x)\text{sgn}(y-x)y^{j-1}p(y,a_{i};T)dxdy=Z_{\vec{n}+e_{i}}\delta_{j,n_{i}+1}, j=1,,ni+1,i=1,2,\displaystyle\quad j=1,\cdots,n_{i}+1,\quad i=1,2,
2Rn+ei(x)sgn(yx)yj1p(y,ai;T)𝑑x𝑑y=Zn+eiδj,ni,\displaystyle\int_{\mathbb{R}^{2}}R_{\vec{n}+e_{i}}(x)\text{sgn}(y-x)y^{j-1}p(y,a_{i};T)dxdy=Z_{\vec{n}+e_{i}}\delta_{j,n_{i}}, j=1,,ni+1,i=1,2,\displaystyle\quad j=1,\cdots,n_{i}+1,\quad i=1,2,

where Zn1,n2Z_{n_{1},n_{2}} is given by (4.21).

Instead, inspired by the generating function in the forms (4.12) and (4.16), we could introduce a new concept of multiple partial-skew-orthogonal polynomials. Namely, we could define another family of functions R~n+1,n2\tilde{R}_{n+1,n_{2}} and R~n1,n1+1\tilde{R}_{n_{1},n_{1}+1} to replace Rn1+1,n2R_{n_{1}+1,n_{2}} and Rn1,n2+1R_{n_{1},n_{2}+1} when n1+n2n_{1}+n_{2} is odd.

Proposition 4.5.

For n1+n2n_{1}+n_{2} being odd, we could define

R~n1+1,n2(x)=Pf[0Q(u(j))00Q(u(i))Q2(u(i),u(j))H(u(i),v(k))ψ1,i(x)0H(v(l),u(j))Q~2(v(l),v(k))ψ2,l(x)0ψ1,j(x)ψ2,k(x)0]i,j=1,,n1+1k,l=1,,n2,\displaystyle\tilde{R}_{n_{1}+1,n_{2}}(x)=\operatorname{Pf}\left[\begin{array}[]{cccc}0&Q_{\mathbb{R}}(u^{(j)})&0&0\\ -Q_{\mathbb{R}}(u^{(i)})&Q_{\mathbb{R}^{2}}(u^{(i)},u^{(j)})&H(u^{(i)},v^{(k)})&\psi_{1,i}(x)\\ 0&-H(v^{(l)},u^{(j)})&\tilde{Q}_{\mathbb{R}^{2}}(v^{(l)},v^{(k)})&\psi_{2,l}(x)\\ 0&-\psi_{1,j}(x)&-\psi_{2,k}(x)&0\end{array}\right]_{i,j=1,\cdots,n_{1}+1\atop k,l=1,\cdots,n_{2}},
R~n1+1,n2(x)=Pf[00Q~(v(k))00Q2(u(i),u(j))H(u(i),v(k))ψ1,i(x)Q~(v(l))H(v(l),u(j))Q~2(v(k),v(l))ψ2,l(x)0ψ1,j(x)ψ2,k(x)0]i,j=1,,n1k,l=1,,n2+1.\displaystyle\tilde{R}_{n_{1}+1,n_{2}}(x)=\operatorname{Pf}\left[\begin{array}[]{cccc}0&0&\tilde{Q}_{\mathbb{R}}(v^{(k)})&0\\ 0&Q_{\mathbb{R}^{2}}(u^{(i)},u^{(j)})&H(u^{(i)},v^{(k)})&\psi_{1,i}(x)\\ -\tilde{Q}_{\mathbb{R}}(v^{(l)})&-H(v^{(l)},u^{(j)})&\tilde{Q}_{\mathbb{R}^{2}}(v^{(k)},v^{(l)})&\psi_{2,l}(x)\\ 0&-\psi_{1,j}(x)&-\psi_{2,k}(x)&0\end{array}\right]_{i,j=1,\cdots,n_{1}\atop k,l=1,\cdots,n_{2}+1}.

Moreover, these functions satisfy the following partial-skew orthogonality

2Rn+e1(x)sgn(yx)yj1p(y,ai;T)𝑑x𝑑y=Q(u(j))Zn+e1δi,1,1jn1+1,\displaystyle\int_{\mathbb{R}^{2}}R_{\vec{n}+e_{1}}(x)\text{sgn}(y-x)y^{j-1}p(y,a_{i};T)dxdy=Q_{\mathbb{R}}(u^{(j)})Z_{\vec{n}+e_{1}}\delta_{i,1},\quad 1\leq j\leq n_{1}+1,
2Rn+e2(x)sgn(yx)yj1p(y,ai;T)𝑑x𝑑y=Q~(v(j))Zn+e1δi,2,1jn2+1.\displaystyle\int_{\mathbb{R}^{2}}R_{\vec{n}+e_{2}}(x)\text{sgn}(y-x)y^{j-1}p(y,a_{i};T)dxdy=\tilde{Q}_{\mathbb{R}}(v^{(j)})Z_{\vec{n}+e_{1}}\delta_{i,2},\quad 1\leq j\leq n_{2}+1.

As mentioned in [23], the advantage of partial-skew-orthogonality mainly lie in the existence of Pfaffian formula for odd order. We hope that these multiple partial-skew-orthogonal polynomials would be useful in the future study of classical integrable systems, including spectral transformations and corresponding Pfaffian τ\tau-functions.

5. Generating functions for non-intersecting paths between several sets and intervals

In this section, we consider non-intersecting paths between several sets and intervals. To this end, we consider different sets of vertices {𝐮i}i=1m\{\mathbf{u}_{i}\}_{i=1}^{m} and {𝐯i}i=1n\{\mathbf{v}_{i}\}_{i=1}^{n} and different intervals of vertices {Ii}i=1n\{I_{i}\}_{i=1}^{n} and {Ji}i=1m\{J_{i}\}_{i=1}^{m}. Here we assume that each 𝐮i\mathbf{u}_{i} has rir_{i} vertices labelled by 𝐮i=(ui(1),,ui(ri))\mathbf{u}_{i}=(u_{i}^{(1)},\cdots,u_{i}^{(r_{i})}) and each 𝐯j\mathbf{v}_{j} has sjs_{j} vertices labelled by 𝐯j=(vj(1),,vj(sj))\mathbf{v}_{j}=(v_{j}^{(1)},\cdots,v_{j}^{(s_{j})}). Moreover, we should assume that IiI_{i} and JjJ_{j} are both totally ordered sets of vertices.

Definition 5.1 (T(𝐮1,,𝐮m;I1,In)T\left(\mathbf{u}_{1},\ldots,\mathbf{u}_{m};I_{1},\ldots I_{n}\right)).

For mm different vertex sets 𝐮1,,𝐮m\mathbf{u}_{1},\cdots,\mathbf{u}_{m} and nn different vertices intervals I1,,InI_{1},\cdots,I_{n}, let’s define T(𝐮1,,𝐮m,I1,In)T\left(\mathbf{u}_{1},\ldots,\mathbf{u}_{m},I_{1},\ldots I_{n}\right) as direct sums of sets and intervals by inserting {Ii}i=1n\{I_{i}\}_{i=1}^{n} into 𝐮1𝐮m\mathbf{u}_{1}\oplus\cdots\oplus\mathbf{u}_{m}, and each vertex in IiI_{i} precedes those in IjI_{j} for i<ji<j.

For example, if there are two sets of vertices 𝐮1,𝐮𝟐\mathbf{u}_{1},\mathbf{u_{2}} and two intervals I1,I2I_{1},I_{2}, then

T(𝐮1,𝐮𝟐;I1,I2)={𝐮1𝐮2I1I2,𝐮1I1𝐮2I2,𝐮1I1I2𝐮2,\displaystyle T(\mathbf{u}_{1},\mathbf{u_{2}};I_{1},I_{2})=\{\mathbf{u}_{1}\oplus\mathbf{u}_{2}\oplus I_{1}\oplus I_{2},\mathbf{u}_{1}\oplus I_{1}\oplus\mathbf{u}_{2}\oplus I_{2},\mathbf{u}_{1}\oplus I_{1}\oplus I_{2}\oplus\mathbf{u}_{2},
I1𝐮1𝐮2I2,I1𝐮1I2𝐮2,I1I2𝐮1𝐮2}.\displaystyle I_{1}\oplus\mathbf{u}_{1}\oplus\mathbf{u}_{2}\oplus I_{2},I_{1}\oplus\mathbf{u}_{1}\oplus I_{2}\oplus\mathbf{u}_{2},I_{1}\oplus I_{2}\oplus\mathbf{u}_{1}\oplus\mathbf{u}_{2}\}.

Moreover, we can define Tτ(𝐮1,,𝐮m;I1,,In)T_{\tau}(\mathbf{u}_{1},\cdots,\mathbf{u}_{m};I_{1},\cdots,I_{n}) is a specific element in T(𝐮1,,𝐮m;I1,,In)T(\mathbf{u}_{1},\cdots,\mathbf{u}_{m};I_{1},\cdots,I_{n}) with ordering τ\tau. Given this τ\tau, we could define T^τ(J1,,Jm;𝐯1,,𝐯n)\hat{T}_{\tau}(J_{1},\cdots,J_{m};\mathbf{v}_{1},\cdots,\mathbf{v}_{n}) by replacing 𝐮i\mathbf{u}_{i} by JiJ_{i} for i=1,,mi=1,\cdots,m, and replacing IiI_{i} by 𝐯i\mathbf{v}_{i} for i=1,,ni=1,\cdots,n. For example, if we define

Tτ(𝐮1,𝐮2,I1,I2)=𝐮1I1𝐮2I2,\displaystyle T_{\tau}(\mathbf{u}_{1},\mathbf{u}_{2},I_{1},I_{2})=\mathbf{u}_{1}\oplus I_{1}\oplus\mathbf{u}_{2}\oplus I_{2},

then we have

T^τ(J1,J2;𝐯1,𝐯2)=J1𝐯1J2𝐯2.\displaystyle\hat{T}_{\tau}(J_{1},J_{2};\mathbf{v}_{1},\mathbf{v}_{2})=J_{1}\oplus\mathbf{v}_{1}\oplus J_{2}\oplus\mathbf{v}_{2}.

With this setting, we could define non-intersecting paths between several sets and intervals.

Definition 5.2 (𝒫0(Tτ(𝐮1,,𝐮m;I1,In);T^τ(J1,,Jm;𝐯1,𝐯n))\mathscr{P}_{0}(T_{\tau}\left(\mathbf{u}_{1},\cdots,\mathbf{u}_{m};I_{1},\cdots I_{n}\right);\hat{T}_{\tau}\left(J_{1},\cdots,J_{m};\mathbf{v}_{1},\cdots\mathbf{v}_{n}\right))).

Given τ\tau as a specific ordering, let’s denote 𝒫0(Tτ(𝐮1,,𝐮m;I1,,In);T^τ(J1,,Jm;𝐯1,,𝐯n))\mathscr{P}_{0}(T_{\tau}\left(\mathbf{u}_{1},\cdots,\mathbf{u}_{m};I_{1},\cdots,I_{n}\right);\hat{T}_{\tau}\left(J_{1},\cdots,J_{m};\mathbf{v}_{1},\cdots,\mathbf{v}_{n}\right)) as all disjoint paths where vertices in 𝐮i\mathbf{u}_{i} go to 𝐯j\mathbf{v}_{j} and JiJ_{i} for any i=1,,mi=1,\cdots,m and j=1,,nj=1,\cdots,n, while vertices in IjI_{j} go to the rest of 𝐯j\mathbf{v}_{j}. We further assume that the number of paths from IjI_{j} to 𝐯j\mathbf{v}_{j} and those from 𝐮i\mathbf{u}_{i} to JjJ_{j} are both even.

Definition 5.3 (Continuity of 1-Factors).

A 1-factor

π(u1(1),,u1(r1),,um(1),,um(rm),vn(sn),,vn(1),,v1(s1),,v1(1))\displaystyle\pi\in\mathscr{F}\left(u_{1}^{(1)},\cdots,u_{1}^{(r_{1})},\cdots,u_{m}^{(1)},\cdots,u_{m}^{(r_{m})},v_{n}^{(s_{n})},\cdots,v_{n}^{(1)},\cdots,v_{1}^{(s_{1})},\cdots,v_{1}^{(1)}\right) (5.1)

is said to be continuous if it satisfies the following two conditions: 1. If there is some up(k)u_{p}^{(k)} paired with vq(l)v_{q}^{(l)}, then up(k+1)u_{p}^{(k+1)} is paired with vq(l+1)v_{q}^{(l+1)}, then up(k1)u_{p}^{(k-1)} must be paired with vq(l1)v_{q}^{(l-1)}, until all elements in 𝐮p\mathbf{u}_{p} or 𝐯q\mathbf{v}_{q} are fully paired, and the rest in 𝐯𝐪\mathbf{v_{q}} and 𝐮𝐩\mathbf{u_{p}} are internally paired separately (see Figure 2.a) ; 2. Under the order τ\tau, if 𝐮i\mathbf{u}_{i} and 𝐮i+1\mathbf{u}_{i+1} are adjacent, and ui(ri)u_{i}^{(r_{i})} is paired with vq(l)v_{q}^{(l)}, then ui+1(1)u_{i+1}^{(1)} is paired with vq(l+1)v_{q}^{(l+1)} (see Figure 2.b). This is also true for 𝐯i\mathbf{v}_{i} and 𝐯i+1\mathbf{v}_{i+1}.

Refer to caption
(a) condition 1 of the continuity of 1-factors
Refer to caption
(b) condition 2 of the continuity of 1-factors
Figure 2. Graphic explanations for the continuity of 1-factors

It should be noted that a continuous 1-factor could be divided into three parts. Namely, for any π(u1(1),,u1(r1),,um(1),,um(rm),vn(sn),,vn(1),,v1(s1),,v1(1))\pi\in\mathscr{F}\left(u_{1}^{(1)},\cdots,u_{1}^{(r_{1})},\cdots,u_{m}^{(1)},\cdots,u_{m}^{(r_{m})},v_{n}^{(s_{n})},\cdots,v_{n}^{(1)},\cdots,v_{1}^{(s_{1})},\cdots,v_{1}^{(1)}\right), we have

π=σπσ1πσ2,\displaystyle\pi=\sigma\cup\pi_{\sigma_{1}}\cup\pi_{\sigma_{2}},

where σ=(up(i),vq(j))\sigma=\bigcup\left(u_{p}^{(i)},v_{q}^{(j)}\right), πσ1=t=1mπσ1(t)\pi_{\sigma_{1}}=\bigcup_{t=1}^{m}\pi_{\sigma_{1}}^{(t)}, πσ1(t)(𝐮t\(σ𝐮t))\pi_{\sigma_{1}}^{(t)}\in\mathscr{F}\left(\mathbf{u}_{t}\backslash\left(\sigma\cap\mathbf{u}_{t}\right)\right), and πσ2=t=1nπσ2(t)\pi_{\sigma_{2}}=\bigcup_{t=1}^{n}\pi_{\sigma_{2}}^{(t)} ,πσ2(t)(𝐯t\(σ𝐯t))\pi_{\sigma_{2}}^{(t)}\in\mathscr{F}\left(\mathbf{v}_{t}\backslash\left(\sigma\cap\mathbf{v}_{t}\right)\right). Here τ\tau means a continuous pairing between 𝐮\mathbf{u} and 𝐯\mathbf{v}, and πσ1\pi_{\sigma_{1}} (respectively πσ2\pi_{\sigma_{2}}) means a pairing between vertices in 𝐮\mathbf{u} (respectively in 𝐯\mathbf{v}). We denote all continuous 1-factors as a set Λσ\Lambda_{\sigma}.

Here we remark that 𝐮i\mathbf{u}_{i} and 𝐮i+1\mathbf{u}_{i+1} could not be merged into one vertex set, since there could be paths from 𝐮iIi\mathbf{u}_{i}\to I_{i} and from 𝐮i+1Ii+1\mathbf{u}_{i+1}\to I_{i+1}. These paths are different.

Definition 5.4 (DD-separated).

Two ordered sets of vertices Tτ(𝐮1,𝐮2,𝐮m,J1,J2,,Jn)T_{\tau}\left(\mathbf{u}_{1},\mathbf{u}_{2},\ldots\mathbf{u}_{m},J_{1},J_{2},\ldots,J_{n}\right) and T^τ(I1,I2,Im,𝐯1,𝐯2,,𝐯n)\hat{T}_{\tau}\left(I_{1},I_{2},\ldots I_{m},\mathbf{v}_{1},\mathbf{v}_{2},\ldots,\mathbf{v}_{n}\right) are DD-separated if any path from 𝐮i\mathbf{u}_{i} to IiI_{i} does not intersect with any path starting from 𝐮k\mathbf{u}_{k} (ki)(k\neq i), and any path from JjJ_{j} to 𝐯j\mathbf{v}_{j} does not intersect with any path to 𝐯l\mathbf{v}_{l} (lj)(l\neq j). Moreover, and any path from 𝐮i\mathbf{u}_{i} to IiI_{i} does not intersect with any path from JjJ_{j} to 𝐯j\mathbf{v}_{j} for all i=1,,mi=1,\cdots,m and j=1,,nj=1,\cdots,n.

Theorem 5.5.

Let 𝐮i=(ui(1),ui(2),,ui(ri))\mathbf{u}_{i}=\left(u^{(1)}_{i},u^{(2)}_{i},\ldots,u^{(r_{i})}_{i}\right) and 𝐯j=(vj(1),,vj(sj))\mathbf{v}_{j}=\left(v^{(1)}_{j},\ldots,v^{(s_{j})}_{j}\right) (1im,1jn)(1\leq i\leq m,1\leq j\leq n) be sequences of vertices in the acyclic directed graph DD, and assume that i=1mri+j=1nsj\sum_{i=1}^{m}r_{i}+\sum_{j=1}^{n}s_{j} is even. If for any 1im,1jn1\leq i\leq m,1\leq j\leq n, IjI_{j}, JiJ_{i} are totally ordered subsets of VV, such that Tτ(𝐮1,𝐮2,,𝐮m,I1,I2,In)T_{\tau}\left(\mathbf{u}_{1},\mathbf{u}_{2},\ldots,\mathbf{u}_{m},I_{1},I_{2},\ldots I_{n}\right) and T^τ(J1,J2,,Jm,𝐯1,𝐯2,𝐯n)\hat{T}_{\tau}\left(J_{1},J_{2},\ldots,J_{m},\mathbf{v}_{1},\mathbf{v}_{2},\ldots\mathbf{v}_{n}\right) are DD-compatible and DD-separated. Then

GF[𝒫0(Tτ(𝐮1,𝐮2,,𝐮m,I1,I2,In);T^τ(J1,J2,,Jm,𝐯1,𝐯2,𝐯n))]=Pf[QHHTQ~],\displaystyle{\rm{GF}}\left[\mathscr{P}_{0}\left(T_{\tau}\left(\mathbf{u}_{1},\mathbf{u}_{2},\ldots,\mathbf{u}_{m},I_{1},I_{2},\ldots I_{n}\right);\hat{T}_{\tau}\left(J_{1},J_{2},\ldots,J_{m},\mathbf{v}_{1},\mathbf{v}_{2},\ldots\mathbf{v}_{n}\right)\right)\right]=\operatorname{Pf}\left[\begin{array}[]{cc}{Q}&{H}\\ -{H}^{T}&{\tilde{{Q}}}\end{array}\right], (5.4)

where

Q=diag(QJ1(u1(k),u1(l))1k,lr1,,QJm(um(k),um(l))1k,lrm),\displaystyle{{Q}}=\text{diag}\left(Q_{J_{1}}(u_{1}^{(k)},u_{1}^{(l)})_{1\leq k,l\leq r_{1}},\cdots,Q_{J_{m}}(u_{m}^{(k)},u_{m}^{(l)})_{1\leq k,l\leq r_{m}}\right),
Q~=diag(Q~In(vn(k),vn(l))1k,lsn,,Q~I1(v1(k),v1(l))1k,ls1),\displaystyle{{\tilde{Q}}}=\text{diag}\left(\tilde{Q}_{I_{n}}(v_{n}^{(k)},v_{n}^{(l)})_{1\leq k,l\leq s_{n}},\cdots,\tilde{Q}_{I_{1}}(v_{1}^{(k)},v_{1}^{(l)})_{1\leq k,l\leq s_{1}}\right),
H=(H𝐮i,𝐯j)i=1,,m,j=n,,1,H𝐮i,𝐯j=(h(ui(k),vj(sj+1l)))k=1,,ri,l=1,,sj.\displaystyle{{H}}=\left(H_{\mathbf{u}_{i},\mathbf{v}_{j}}\right)_{i=1,\cdots,m,j=n,\cdots,1},\quad H_{\mathbf{u}_{i},\mathbf{v}_{j}}=\left(h(u_{i}^{(k)},v_{j}^{(s_{j}+1-l)})\right)_{k=1,\cdots,r_{i},l=1,\cdots,s_{j}}.
Proof.

First, we know that the right hand side of equation (5.4) could be expanded as

πsgn(π)((uk(i),uk(j))πQIk(uk(i),uk(j)))((uk(i),vl(j))πh(uk(i),vl(j)))((vl(j),vl(i))πQ~Jl(vl(j),vl(i)))\displaystyle\sum\limits_{\pi}{{\text{sgn}}}(\pi)\left(\prod\limits_{(u_{k}^{(i)},u_{k}^{(j)})\in\pi}Q_{I_{k}}(u_{k}^{(i)},u_{k}^{(j)})\right)\left(\prod\limits_{(u_{k}^{(i)},v_{l}^{(j)})\in\pi}h(u_{k}^{(i)},v_{l}^{(j)})\right)\left(\prod\limits_{(v_{l}^{(j)},v_{l}^{(i)})\in\pi}{\tilde{Q}}_{J_{l}}(v_{l}^{(j)},v_{l}^{(i)})\right)

according to the formula (3.4), where the summation is over all continuous 1-factors (5.1). Moreover, we could recognize this formula as a product of path weights and rewrite it as

Pf[QHHTQ~]=(π,𝐏1,,𝐏m,𝐐1,,𝐐n)sgn(π)ω(π,𝐏1,,𝐏m,𝐐1,,𝐐n),\displaystyle\operatorname{Pf}\left[\begin{array}[]{cc}Q&H\\ -H^{T}&\tilde{Q}\end{array}\right]=\sum_{\left(\pi,\mathbf{P}_{1},\ldots,\mathbf{P}_{m},\mathbf{Q}_{1},\ldots,\mathbf{Q}_{n}\right)}\operatorname{sgn}(\pi)\omega\left(\pi,\mathbf{P}_{1},\ldots,\mathbf{P}_{m},\mathbf{Q}_{1},\ldots,\mathbf{Q}_{n}\right), (5.7)

where 𝐏k\mathbf{P}_{k} and 𝐐k\mathbf{Q}_{k} are a set of paths, 𝐏k=(Pk(1),Pk(2),,Pk(rk))\mathbf{P}_{k}=\left(P_{k}^{(1)},P_{k}^{(2)},\ldots,P_{k}^{\left(r_{k}\right)}\right), 𝐐l=(Ql(1),Ql(2),,Ql(sl))\mathbf{Q}_{l}=\left(Q_{l}^{(1)},Q_{l}^{(2)},\ldots,Q_{l}^{\left(s_{l}\right)}\right), and (π,𝐏1,,𝐏m,𝐐1,,𝐐n)\left(\pi,\mathbf{P}_{1},\ldots,\mathbf{P}_{m},\mathbf{Q}_{1},\ldots,\mathbf{Q}_{n}\right) is an (r+s+1)(r+s+1)-tuple configuration such that

  • If (uk(i),vl(j))π\left(u_{k}^{(i)},v_{l}^{(j)}\right)\in\pi,then Pk(i)=Ql(j)𝒫(uk(i),vl(j))P_{k}^{(i)}=Q_{l}^{\left(j\right)}\in\mathscr{P}\left(u_{k}^{(i)},v_{l}^{(j)}\right);

  • If (uk(i),uk(j))π\left(u_{k}^{(i)},u_{k}^{(j)}\right)\in\pi,then Pk(i)𝒫(uk(i);Ik)P_{k}^{(i)}\in\mathscr{P}\left(u_{k}^{(i)};I_{k}\right),Pk(j)𝒫(uk(j);Ik)P_{k}^{(j)}\in\mathscr{P}\left(u_{k}^{(j)};I_{k}\right),while Pk(i)P_{k}^{(i)} and Pk(j)P_{k}^{(j)} don’t intersect;

  • If (vl(j),vl(i))π\left(v_{l}^{(j)},v_{l}^{(i)}\right)\in\pi,then Ql(j)𝒫(Jl;vl(j))Q_{l}^{(j)}\in\mathscr{P}\left(J_{l};v_{l}^{(j)}\right),Ql(i)𝒫(Jl;vl(i))Q_{l}^{(i)}\in\mathscr{P}\left(J_{l};v_{l}^{(i)}\right),while Ql(j)Q_{l}^{(j)} and Ql(i)Q_{l}^{(i)} don’t intersect.

Then, we seek for a sign-reversing involution to offset the intersecting paths. Since the graph is DD-separated, we know that if paths intersect, then the intersecting point belongs to Pk(i)Pl(j)P_{k}^{(i)}\cap P_{l}^{(j)} or Qk(i)Ql(j)Q_{k}^{(i)}\cap Q_{l}^{(j)} for some k,lk,l and i,ji,j. Similar to the proof in Theorem 4.3, we find the smallest intersection point ww^{*}333In this case, we find the smallest intersection point starting from paths in 𝐏1\mathbf{P}_{1}. If there is an intersection point in 𝐏1\mathbf{P}_{1}, then we find it from paths starting from u1(1)u_{1}^{(1)} and the rest is the same with the case in Theorem 4.3. If there isn’t any intersection point in 𝐏1\mathbf{P}_{1}, then we find it in 𝐏2\mathbf{P}_{2}, and so on., and assume that Pk(i)P_{k}^{(i)} and Pl(j)P_{l}^{(j)} are the two whose subcripts (priority) k,lk,l and supercripts (followed) i,ji,j are the smallest. Thus, we could construct an involution

ϕ:(π,𝐏1,,𝐏m,𝐐1,,𝐐n)(π,𝐏¯1,,𝐏¯m,𝐐¯1,,𝐐¯n)\displaystyle\phi:\left(\pi,\mathbf{P}_{1},\ldots,\mathbf{P}_{m},\mathbf{Q}_{1},\ldots,\mathbf{Q}_{n}\right)\to\left(\pi,\bar{\mathbf{P}}_{1},\ldots,\bar{\mathbf{P}}_{m},\bar{\mathbf{Q}}_{1},\ldots,\bar{\mathbf{Q}}_{n}\right)

where

P¯k(i)=Pk(i)(w)Pl(j)(w),P¯l(j)=Pl(j)(w)Pk(i)(w),\displaystyle\bar{P}_{k}^{(i)}=P_{k}^{(i)}(\to w^{*})P_{l}^{(j)}(w^{*}\to),\quad\bar{P}_{l}^{(j)}=P_{l}^{(j)}(\to w^{*})P_{k}^{(i)}(w^{*}\to),

and the rest paths remain invariant. Using the method shown in Theorem 4.3, we can verify that this map is a sign-reversing involution by the property of DD-separated graph.

By using the sign-reversing involutions to cancel the intersection terms, we find that only configurations containing continuous 1-factors are left. As we divide the continuous 1-factor into 3 parts, (5.4) could be rewritten as the non-intersecting part of

σπΛσsgn(π)(k=1ml=1n(uk(i),vl(j))σh(uk(i),vl(j)))×(k=1m(uk(i),uk(j))πσ1QIk(uk(i),uk(j)))(l=1n(vl(j),vl(i))πσ2Q~Jl(vl(j),vl(i)))\displaystyle\begin{aligned} \sum_{\sigma}\sum_{\pi\in\Lambda_{\sigma}}\operatorname{sgn}(\pi)&\left(\prod_{k=1}^{m}\prod_{l=1}^{n}\prod_{\left(u_{k}^{(i)},v_{l}^{(j)}\right)\in\sigma}h\left(u_{k}^{(i)},v_{l}^{(j)}\right)\right)\\ &\times\left(\prod_{k=1}^{m}\prod_{\left(u_{k}^{(i)},u_{k}^{(j)}\right)\in\pi_{\sigma_{1}}}Q_{I_{k}}\left(u_{k}^{(i)},u_{k}^{(j)}\right)\right)\left(\prod_{l=1}^{n}\prod_{\left(v_{l}^{(j)},v_{l}^{(i)}\right)\in\pi_{\sigma_{2}}}\tilde{Q}_{J_{l}}\left(v_{l}^{(j)},v_{l}^{(i)}\right)\right)\end{aligned} (5.8)
Refer to caption
Figure 3. An explanation of non-intersecting pairing between 𝐮1I1𝐮2J1𝐯1J2\mathbf{u}_{1}\oplus I_{1}\oplus\mathbf{u}_{2}\to J_{1}\oplus\mathbf{v}_{1}\oplus J_{2}.

Moreover, by realizing

πΛσsgn(π)=(sgn(πσ1(1)))(sgn(πσ1(m)))(sgn(πσ2(1)))(sgn(πσ2(n)))=1,\displaystyle\sum\limits_{\pi\in{\Lambda_{\sigma}}}{{\mathop{\rm sgn}}}(\pi)=\left({\sum{{\mathop{\rm sgn}}}\left({{\pi_{\sigma_{1}^{(1)}}}}\right)}\right)\cdots\left(\sum\text{sgn}\left(\pi_{\sigma_{1}^{(m)}}\right)\right)\left({\sum{{\mathop{\rm sgn}}}\left({{\pi_{\sigma_{2}^{(1)}}}}\right)}\right)\cdots\left({\sum{{\mathop{\rm sgn}}}\left({{\pi_{\sigma_{2}^{(n)}}}}\right)}\right)=1,

we know that (5.8) represents GF[𝒫0(Tτ(𝐮1,𝐮m,J1,,Jn);T^τ(I1,Im,𝐯1,,𝐯n))]\operatorname{GF}\left[\mathscr{P}_{0}\left(T_{\tau}\left(\mathbf{u}_{1},\ldots\mathbf{u}_{m},J_{1},\ldots,J_{n}\right);\hat{T}_{\tau}\left(I_{1},\ldots I_{m},\mathbf{v}_{1},\ldots,\mathbf{v}_{n}\right)\right)\right], and there are even paths from 𝐮kIk\mathbf{u}_{k}\to I_{k} and even paths from Jl𝐯lJ_{l}\to\mathbf{v}_{l}.

Acknowledgement

We would like to thank the referee for useful comments. This work is partially funded by grants (NSFC12101432, NSFC12175155).

References

  • [1] J. Baik and Z. Liu. Discrete Toeplitz/Hankel determinants and the width of non-intersecting processes. Int. Math. Res. Not., 20 (2014), 5737-5768.
  • [2] A. Borodin. Biorthogonal ensembles. Nucl. Phys. B, 536 (1998) 704-732.
  • [3] S. Carrozza and A. Tanasa. Pfaffians and nonintersecting paths in graphs with cycles: Grassmann algebra methods. Adv. Appl. Math., 93 (2018), 108-120.
  • [4] M. Ciucu and C. Krattenthaler. The interaction of a gap with a free boundary in a two dimensional dimen system. Commun. Math. Phys., 302 (2011), 253-289.
  • [5] X. Chang, Y. He, X. Hu and S. Li. Partial-skew-orthogonal polynomials and related integrable lattices with Pfaffian τ\tau-function. Comm. Math. Phys., 364 (2018) 1069-1119.
  • [6] X. Chen, X. Chang, J. Sun, X. Hu and Y. Yeh. Three semi-discrete integrable systems related to orthogonal polynomials and their generalized determinant solutions. Nonlinearity, 28 (2015), 2279-2306.
  • [7] E. Daems and A. Kuijlaars. Multiple orthogonal polynomials of mixed type and non-intersecting Brownian motions. J. Approximation Theo., 146 (2007), 91-114.
  • [8] M. Duits and A. Kuijlaars. The two-periodic Aztec diamond and matrix valued orthogonal polynomials. J. Eur. Math. Soc., 23 (2021), 1075-1131.
  • [9] P. Forrester, S. Majumdar and G. Schehr. Non-intersecting Brownian walkers and Yang-Mills theory on the sphere. Nuclear Phys. B, 844 (2011), 500-526.
  • [10] P. Forrester and T. Nagao. Vicious random walkers and a discretization of Gaussian random matrix ensembles. Nuclear Phys. B, 620 (2002), 551-565.
  • [11] I. Gessel and G. Viennot. Binomial determinants, paths, and hook length formulae. Adv. Math., 58 (1985), 300-321.
  • [12] R. Gharakhloo and K. Liechty. Bordered and framed Toeplitz and Hankel determinants with applications in integrable probability. arXiv: 2401.01971.
  • [13] R. Gharakhloo and N. Witte. Modulated bi-orthogonal polynomials on the unit circle: the 2jk2j-k and j2kj-2k systems. Constr. Approx., 58 (2023), 1-74.
  • [14] A. Groot and A. Kuijlaars. Matrix-valued orthogonal polynomials related to hexagon tilings. J. Approximation Thoery, 270 (2021), 105619.
  • [15] M. Ishikawa, H. Tagawa and J. Zeng. Pfaffian decomposition and a Pfaffian analogue of qq-Catalan Hankel determinants. J. Combin. Theory Ser. A., 120 (2013), 1263-1284.
  • [16] M. Ishikawa and M. Wakayama. Applications of minor summation formula III, Plücker relations, lattice paths and Pfaffian identities. J. Combin. Theory Ser. A., 113 (2006), 113-155.
  • [17] K. Johansson. Non-intersecting paths, random tilings and random matrices. Probab. Theory Related Fields., 123 (2002), 225-280.
  • [18] K. Johansson. Non-colliding Brownian motions and the extended tacnode process. Comm. Math. Phys., 319 (2013), 231-267.
  • [19] S. Karlin and J. McGregor. Coincidence probabilities. Pacific J. Math., 9 (1959), 1141-1164.
  • [20] C. Krattenthaler, A. Guttmann and X. Viennot. Vicious walkers, friendly walkers and Young tableaux. II. With a wall. J. Phys. A 33 (2000), 8835.
  • [21] C. Krattenthaler. Lattice path enumeration. Discrete Math. Appl., CRC Press, Boca Raton, FL, 2015, 589-678.
  • [22] S. Li, B. Shen, J. Xiang and G. Yu. Multiple skew orthogonal polynomials and 2-component Pfaff lattice hierarchy. Ann. Henri Poincaré, https://doi.org/10.1007/s00023-023-01382-2.
  • [23] S. Li and G. Yu. Christoffel transformations for (partial-)skew-orthogonal polynomials and applications. Adv. Math., 436 (2024), 109398.
  • [24] B. Lindström. On the vector representations of induced matroids. Bull. London Math. Soc., 5 (1973), 85-90
  • [25] M. Mehta and R. Wang. Calculation of a certain determinant. Comm. Math. Phys., 214 (2000), 227-232.
  • [26] Y. Nakamura and A. Zhedanov. Special solutions of the Toda chain and combinatorial numbers. J. Phys. A, 37 (2004), 5849-5862.
  • [27] S. Okada. On the generating functions for certain classes of plane partitions. J. Combin. Theory Ser. A, 51 (1989), 1-23.
  • [28] B. Sagan. The symmetric group. GTM Vol. 203, Springer, 2001.
  • [29] B. Shen, S. Li and G. Yu. Evaluations of certain Catalan-Hankel Pfaffians via classical skew orthogonal polynomials. J. Phys. A, 54 (2021) 264001.
  • [30] R. Stanley. Enumerative combinatorics, Vol.2, Cambridge University Press, 1999.
  • [31] J. Stembridge. Nonintersecting paths, Pfaffians, and plane partition. Adv. Math., 83 (1990), 96-131.