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

Simple Generating Functions for Certain Young Tableaux with Periodic Walls

Feihu Liu1 and Guoce Xin2,∗ 1,2School of Mathematical Sciences, Capital Normal University, Beijing 100048, PR China 1liufeihu7476@163.com & 2guoce_xin@163.com
(Date: January 26, 2024)
Abstract.

Recently, Banderier et. al. considered Young tableaux with walls, which are similar to standard Young tableaux, except that local decreases are allowed at some walls. We count the numbers f¯m(n)\overline{f}_{m}(n) of Young tableaux of shape 2×mn2\times mn with walls, that allow local decreases at the (jm+i)(jm+i)-th columns for all j=0,,n1j=0,\dots,n-1 and i=2,,mi=2,\dots,m. We find that they have nice generating functions (thanks to the OEIS) as follows.

F¯m(x)=n0f¯m(n)xn=k=1mC(ek2πimx1m)=exp(n1(2mn1mn1)xnn),\overline{F}_{m}(x)=\sum_{n\geq 0}\overline{f}_{m}(n)x^{n}=\prod_{k=1}^{m}C(e^{k\frac{2\pi i}{m}}x^{\frac{1}{m}})=\exp\left(\sum_{n\geq 1}\binom{2mn-1}{mn-1}\frac{x^{n}}{n}\right),

where C(x)=114x2xC(x)=\frac{1-\sqrt{1-4x}}{2x} is the well-known Catalan generating function. We prove generalizations of this result. Firstly, we use the Yamanouchi word to transform Young tableaux with horizontal walls into lattice paths. This results in a determinant formula. Then by lattice path counting theory, we obtain the generating functions Fr(x)F_{r}(x) for the number of lattice paths from (0,0)(0,0) to (nr,kn)(\ell n-r,kn) that never go above the path (NkE)n1NkEr(N^{k}E^{\ell})^{n-1}N^{k}E^{\ell-r}, where N,EN,E stand for north and east steps, respectively. We also obtain exponential formulas for F1(x)F_{1}(x) and F(x)F_{\ell}(x). The formula for F¯m(x)\overline{F}_{m}(x) is thus proved since it is just F1(x)F_{1}(x) specializes at k==mk=\ell=m.

* This work is partially supported by NSFC(12071311).

Mathematic subject classification: Primary 05A15; Secondary 05A10, 05E05.

Keywords: Rectangular Young tableau; Lattice path; Puiseux’s theorem; Symmetric function; Catalan number.

1. Introduction

Throughout this paper, \mathbb{C}, \mathbb{Z}, \mathbb{N} and \mathbb{P} denote the set of all complex numbers, all integers, non-negative integers and positive integers, respectively.

Recently, Banderier et. al. [1, 2] considered a variation of standard Young tableaux, called Young tableaux with walls. We give a rigorous definition for our exploration. Let λ\lambda be a partition of nn, denoted λn\lambda\vdash n. Its Young diagram is also denoted λ\lambda. See, e.g., [12, Section 1.7] for detailed definitions. In this paper, we focus on λ=(m,m)\lambda=(m,m), so the Young diagram of λ\lambda is a 2×m2\times m rectangle with |λ|=2m|\lambda|=2m cells. Edges between two neighboring cells of λ\lambda will be referred to as walls. A Young building \mathcal{B} of shape λ\lambda is a pair =(λ,W)\mathcal{B}=(\lambda,W) where WW is a subset of the walls of λ\lambda. The walls in WW are depicted as bold red edges. A Young tableau with walls over \mathcal{B} is a filling of the cells of λ\lambda by labels 1,2,,n1,2,\dots,n, such that each label appears exactly once and the labels are increasing along each row and column, (conditions for standard Young tableaux) but two labels separated by a wall need not be increasing. Denote by YT(λ,W)YT(\lambda,W) the set of all such tableaux. It reduces to the set of standard Young tableaux of shape λ\lambda when WW is empty. See Figure 1 for an example of Young tableaux with walls of shape λ=(6,6)\lambda=(6,6).

Refer to caption
Figure 1. A Young tableaux of shape λ=(6,6)\lambda=(6,6) with walls.

Let =(λ,W)\mathcal{B}=(\lambda,W) be a rectangular building. That is, λ\lambda is of rectangular shape. Define n\mathcal{B}^{n} to be the Young building formed by connecting nn blocks of \mathcal{B} horizontally. Then Young tableaux over n\mathcal{B}^{n} are called Young tableaux with periodic walls. Note that in this definition, WW is allowed to contain the rightmost edges of λ\lambda. Figure 2 illustrates a Young building 2\mathcal{B}_{2} on the left and a Young tableau with walls in YT(23)YT(\mathcal{B}_{2}^{3}) on the right.

Refer to caption
Figure 2. A Young tableaux with periodic walls of a block 2\mathcal{B}_{2} of shape (2,2)(2,2).

In [2], Banderier and Wallner counted Young tableaux with periodic walls by “the density method”. They considered a classification of 2×22\times 2 periodic shapes (a total of 26=642^{6}=64 different models) and obtained counting formulas for 6060 models. In particular, they obtained the counting formula

f¯2(n):=|YT(2n)|=22n+1Cat(n)Cat(2n+1),\overline{f}_{2}(n):=|YT(\mathcal{B}_{2}^{n})|=2^{2n+1}\texttt{Cat}(n)-\texttt{Cat}(2n+1),

where Cat(n)\texttt{Cat}(n) is the well-known Catalan number with the Catalan generating function

C(x)=n0Cat(n)xn=114x2x=n01n+1(2nn)xn.C(x)=\sum_{n\geq 0}\texttt{Cat}(n)x^{n}=\frac{1-\sqrt{1-4x}}{2x}=\sum_{n\geq 0}\frac{1}{n+1}\binom{2n}{n}x^{n}. (1)

Note that f¯2(n)\overline{f}_{2}(n) is the sequence [A079489] in OEIS [11].

More generally, let mm be a positive integer and m\mathcal{B}_{m} be the Young building shape λ=(m,m)\lambda=(m,m) with horizontal walls in all columns except the first column. See Figure 4 (on the left) for two examples of elements in YT(32)YT(\mathcal{B}_{3}^{2}). We are interested in the counting formula of f¯m(n)=|YT(mn)|\overline{f}_{m}(n)=|YT(\mathcal{B}_{m}^{n})|.

Computer experiment suggests that f¯m(n)\overline{f}_{m}(n) for m=3,4,5,6m=3,4,5,6 are respectively the sequences [A213403], [A213404], [A213405] and [A213406] in OEIS [11]. These sequences have nice generating functions, and can be summarized as the following result.

Theorem 1.1.

Suppose mm is a positive integer. Let f¯m(n)\overline{f}_{m}(n) denote the number of Young tableaux with periodic walls over mn\mathcal{B}_{m}^{n}, with convention f¯m(0)=1\overline{f}_{m}(0)=1. Then we have

F¯m(x)=n0f¯m(n)xn=k=1mC(ξkx1m)=exp(n1(2mn1mn1)xnn),\overline{F}_{m}(x)=\sum_{n\geq 0}\overline{f}_{m}(n)x^{n}=\prod_{k=1}^{m}C(\xi^{k}x^{\frac{1}{m}})=\exp\left(\sum_{n\geq 1}\binom{2mn-1}{mn-1}\frac{x^{n}}{n}\right),

where ξ=e2πim\xi=e^{\frac{2\pi i}{m}} and C(x)C(x) is the Catalan generating function.

The motivation of this paper is to give a proof of this theorem.

To this end, we find several combinatorial interpretations of f¯m(n)\overline{f}_{m}(n), including a lattice path interpretation. This allows us to prove Theorem 1.1 using lattice path counting theory (e.g., [14, 4, 8]). Indeed, our first result is the following.

Theorem 1.2.

Let k,k,\ell\in\mathbb{P}, nn\in\mathbb{N} and 1r1\leq r\leq\ell. Let fr(n)f_{r}(n) be the number of lattice paths from (0,0)(0,0) to (nr,kn)(\ell n-r,kn) that never go above the path (NkE)n1NkEr(N^{k}E^{\ell})^{n-1}N^{k}E^{\ell-r}, and let w1,,ww_{1},...,w_{\ell} be the unique solutions of the equation (w1)xwk+=0(w-1)^{\ell}-xw^{k+\ell}=0 that are fractional power series. Then the generating function Fr(x)=n0fr(n)xnF_{r}(x)=\sum_{n\geq 0}f_{r}(n)x^{n} (with the convention fr(0)=1f_{r}(0)=1) is given by

Fr(x)=i=0r1(1)i(r+ii)er+1+i(w1,w2,,w),F_{r}(x)=\sum_{i=0}^{r-1}(-1)^{i}\binom{\ell-r+i}{i}e_{\ell-r+1+i}(w_{1},w_{2},...,w_{\ell}),

where em(w1,,w)e_{m}(w_{1},...,w_{\ell}) is the mm-th elementary symmetric function introduced in subsection 3.1.

Specifically, when r=r=\ell, this result was obtained by de Mier and Noy ([10]) in 2005.

For the two cases of r=r=\ell and r=1r=1, we obtain the following exponential formulas.

Theorem 1.3.

Let q(n)q(n) be the number of lattice paths from (0,0)(0,0) to (n,kn)(\ell n,kn) that never go above the path (NkE)n(N^{k}E^{\ell})^{n}. Let Q(x)=n0q(n)xnQ(x)=\sum_{n\geq 0}q(n)x^{n}. Then we have

Q(x)=exp(n1(kn+nn)xnn).Q(x)=\exp\left(\sum_{n\geq 1}\binom{kn+\ell n}{\ell n}\frac{x^{n}}{n}\right).
Theorem 1.4.

Following the notation in Theorem 1.2. Let f1(n)f_{1}(n) be the number of lattice paths from (0,0)(0,0) to (n1,kn)(\ell n-1,kn) that never go above the path (NkE)n1NkE1(N^{k}E^{\ell})^{n-1}N^{k}E^{\ell-1}. Let F1(x)=n0f1(n)xnF_{1}(x)=\sum_{n\geq 0}f_{1}(n)x^{n}. Then we have

F1(x)=exp(n1(kn+n1n1)xnn).F_{1}(x)=\exp\left(\sum_{n\geq 1}\binom{kn+\ell n-1}{\ell n-1}\frac{x^{n}}{n}\right).

The paper is organized as follows. In Section 2, we consider Young tableaux with horizontal walls over a Young building \mathcal{B}, especially for \mathcal{B} of shape (m,m)(m,m), a two-rowed partition. By using the Yamanouchi word [13, Propsition 7.10.3(d)], such Young tableaux are encoded by yy-sequences, lattice paths, and reverse partitions. This allows us to give a determinant formula for YT()YT(\mathcal{B}) using lattice path counting theory. In particular, f¯m(n)\overline{f}_{m}(n) has a simple determinant formula in Proposition 2.5. Section 3 is devoted to the proof of our main result, Theorem 1.2. In subsection 3.1 we introduce some basic knowledge about symmetric functions and combinatorial sum identities that will be used in our proof. In subsection 3.2, we first introduce the functional equation system in [10] obtained from Matroid Theory, then solve the system and complete the proof of Theorem 1.2. Section 4 focuses on the cases r=1r=1 and r=r=\ell. We prove Theorems 1.3 and 1.4 on their exponential formulas, and complete the proof of Theorem 1.1. In Section 5, we discuss two other possible approaches to Theorem 1.2. This leads to two byproducts in Theorem 5.1 and Theorem 5.4.

2. Rectangular Young Tableaux with Horizontal Walls

Let =(λ,W)\mathcal{B}=(\lambda,W) be a rectangular Young building with horizontal walls. That is, λ=(m,m,,m)\lambda=(m,m,\dots,m), and WW contains only horizontal walls. We use the Yamanouchi word [13, Propsition 7.10.3(d)] to transform Young tableaux with walls over \mathcal{B} into several combinatorial objects, including lattice paths. Though most of the ideas work for general λ\lambda, we focus on the two-row case, i.e., λ=(m,m)\lambda=(m,m), for its simplicity and its close relation with fm(n)f_{m}(n) in Theorem 1.2.

For a Young tableau TT with walls over \mathcal{B} depicted in Figure 3, its labels satisfy x1<x2<x_{1}<x_{2}<\cdots and y1<y2<y_{1}<y_{2}<\cdots. Note that we did not mark walls in Figure 3. Since \mathcal{B} contains only horizontal walls, we may encode WW as a subset SS of [m]:={1,,m}[m]:=\{1,\dots,m\}. Thus we have the extra conditions xi<yix_{i}<y_{i} for all iSi\not\in S.

y1y2y3ymx1x2x3xm\begin{array}[]{|c|c|c|c|c|}\hline\cr y_{1}&y_{2}&y_{3}&\cdots&y_{m}\\ \hline\cr x_{1}&x_{2}&x_{3}&\cdots&x_{m}\\ \hline\cr\end{array}

Figure 3.

Now we introduce several different encodings of TYT((m,m),S)T\in YT((m,m),S).

  1. (1)

    The yy-sequence of TT is y(T)=(y1,y2,,ym)y(T)=(y_{1},y_{2},\dots,y_{m}). The xx-sequence of TT is x(T)=(x1,x2,,xm)x(T)=(x_{1},x_{2},\dots,x_{m}). Then {x1,,xm}=[2m]{y1,,ym}\{x_{1},\dots,x_{m}\}=[2m]\setminus\{y_{1},\dots,y_{m}\}.

  2. (2)

    The Yamanouchi word of TYT(λ,S)T\in YT(\lambda,S) is defined by w(T)=w1w2w2mw(T)=w_{1}w_{2}\cdots w_{2m} with wi=χ(iin first row)w_{i}=\chi(i\ \text{in first row}) where χ(true)=1\chi(true)=1 and χ(false)=0\chi(false)=0.

  3. (3)

    The lattice path of TT is defined by p(T)=w(T)|0N,1Ep(T)=w(T)|_{0\to N,1\to E}, which is the NENE-sequence of a lattice path, with N=(0,1)N=(0,1) and E=(1,0)E=(1,0) standing for north step and east step, respectively.

  4. (4)

    The reverse partition of TT is defined to be the partition μ(T)=(μ1,μ2,,μm)\mu(T)=(\mu_{1},\mu_{2},\dots,\mu_{m}) lying above p(T)p(T). More precisely, μi0\mu_{i}\geq 0 is the number of cells in the (mi+1)(m-i+1)-th row and to the left of p(T)p(T). In some contexts, it is also referred to as the co-area sequence.

For example, if TT is the figure in the bottom left corner of Figure 4 with S={2,3,5,6}S=\{2,3,5,6\}, then the Yamanouchi word of TT is 011100011100011100011100; the yy-sequence of TT is y(T)=(2,3,4,8,9,10)y(T)=(2,3,4,8,9,10); the xx-sequence of TT is x(T)=(1,5,6,7,11,12)x(T)=(1,5,6,7,11,12); the path of TT is p(T)=NE3N3E3N2p(T)=NE^{3}N^{3}E^{3}N^{2}; and the reverse partition of TT is μ(T)=(0,3,3,3,6,6)\mu(T)=(0,3,3,3,6,6). It should be evident that μi=xii\mu_{i}=x_{i}-i.

Refer to caption
Figure 4. An example of bijection for 32\mathcal{B}_{3}^{2} in Proposition 2.5.

The following lemma is immediate.

Lemma 2.1.

For any T,TT,T^{\prime} in YT((m,m),S)YT((m,m),S) as above, the following are equivalent.

  1. (1)

    μ(T)μ(T)\mu(T)\leq\mu(T^{\prime}) componentwise, i.e., μiμi\mu_{i}\leq\mu_{i}^{\prime} for all ii if μ(T)=(μ1,μ2,,μm)\mu(T)=(\mu_{1},\mu_{2},\dots,\mu_{m}) and μ(T)=(μ1,μ2,,μm)\mu(T^{\prime})=(\mu^{\prime}_{1},\mu^{\prime}_{2},\dots,\mu^{\prime}_{m}).

  2. (2)

    The path p(T)p(T) lies above the path p(T)p(T^{\prime});

  3. (3)

    The yy-sequence y(T)y(T) is larger than the yy-sequence y(T)y(T^{\prime}) componentwise. We also denote y(T)y(T)y(T)\leq y(T^{\prime})

We denote by TTT\leq T^{\prime} if one of the above conditions holds (hence all the conditions hold) true.

The proof of the lemma is straightforward and we omit it. Moreover, \leq is a partial order on YT((m,m),[m])YT((m,m),[m]) and hence on its restriction YT((m,m),S)YT((m,m),S) for all SS. Indeed, it is the partial order induced by the containment partial order on their reverse partitions.

Two extreme cases of SS are worth mentioning: i) When SS is empty, TYT((m,m),)T\in YT((m,m),\emptyset) is a standard Young tableau, and its (classical) Yamanouchi word w(T)w(T) corresponds to a Catalan path, i.e., a lattice path that never goes below the diagonal. The Yamanouchi map φ:Tp(T)\varphi:T\mapsto p(T) establishes a bijection from standard Young tableaux of shape (m,m)(m,m) to Catalan paths of size mm. ii) When S=[m]S=[m], one can verify that the map φ\varphi establishes a bijection from YT((m,m),[m])YT((m,m),[m]) to the set of all lattice paths from (0,0)(0,0) to (m,m)(m,m). These two cases invite us to find a good interpretation of the image of YT((m,m),S)YT((m,m),S) under φ\varphi for arbitrary subset SS.

To this end, we need the following lemma.

Lemma 2.2.

Suppose SS is a subset of [m][m]. Then any TYT((m,m),S)T\in YT((m,m),S) satisfies yminy(T)ymax(S)y_{\min}\leq y(T)\leq y_{\max}(S), where ymin=(m+1,m+2,,2m)y_{\min}=(m+1,m+2,\dots,2m) and ymax(S)=(Y1,,Ym)y_{\max}(S)=(Y_{1},\dots,Y_{m}) is determined by the following rule: If iSi\not\in S then Yi=2iY_{i}=2i; otherwise iSi\in S and Yi=i0+iY_{i}=i_{0}+i where i0i_{0} is the largest <i\ell<i satisfying S\ell\not\in S (if no such \ell exists, then we set i0=0i_{0}=0).

Proof.

Assume y(T)=(y1,,ym)y(T)=(y_{1},\dots,y_{m}). Since y1<y2<<ym2my_{1}<y_{2}<\cdots<y_{m}\leq 2m always hold, ymin=(m+1,m+2,,2m)y(T)y_{\min}=(m+1,m+2,\dots,2m)\leq y(T) for any SS.

To see that y(T)ymax(S)y(T)\leq y_{\max}(S), we first check that (Y1,,Ym)(Y_{1},\dots,Y_{m}) is a yy-sequence of certain TYT((m,m),S)T^{\prime}\in YT((m,m),S). This is straightforward. Next we claim that yiYiy_{i}\geq Y_{i} for all ii so that the proof is completed. We have to distinguish the two cases: i) If iSi\not\in S, then y1<<yiy_{1}<\cdots<y_{i} and x1<<xi<yix_{1}<\cdots<x_{i}<y_{i}. Thus yi2i=Yiy_{i}\geq 2i=Y_{i}; ii) If iSi\in S, then y1<<yiy_{1}<\cdots<y_{i} and x1<<xi0<yi0x_{1}<\cdots<x_{i_{0}}<y_{i_{0}}. Thus yii+i0=Yiy_{i}\geq i+i_{0}=Y_{i} as desired. ∎

Proposition 2.3.

Suppose SS is a subset of [m][m]. Let T0,TM(S)T^{0},T^{M}(S) be the tableau in YT((m,m),S)YT((m,m),S) with yy-sequence yminy_{\min} and ymax(S)y_{\max}(S) as in Lemma 2.2, respectively. Then the following sets have the same cardinality.

  1. (1)

    Tableaux TT in YT((m,m),S)YT((m,m),S) that satisfy T0TTM(S)T^{0}\leq T\leq T^{M}(S);

  2. (2)

    Lattice paths from (0,0)(0,0) to (m,m)(m,m) that stay above p(TM(S))p(T^{M}(S));

  3. (3)

    Partitions contained in μ(TM(S))\mu(T^{M}(S)).

Proof.

The bijection between sets (2)(2) and (3)(3) is obvious, so it suffices to show that Tp(T)T\mapsto p(T) is the desired bijection from set (1)(1) to set (2)(2). For a given tableaux TT in set (1)(1), the yy-sequence of TT is y(T)=(y1,y2,,ym)y(T)=(y_{1},y_{2},...,y_{m}). By Proposition 2.2, we have y(T)ymax(S)y(T)\leq y_{\max}(S), i.e., yiYiy_{i}\geq Y_{i}. By Lemma 2.1, the path p(T)p(T) lies above the path p(TM(S))p(T^{M}(S)). So p(T)p(T) is well defined between sets (1)(1) and (2)(2). Moreover, Tp(T)T\mapsto p(T) is an injection.

Given a path PP that stays above p(TM(S))p(T^{M}(S)), we need to find a desired TT satisfying p(T)=Pp(T)=P. Assuming that the yy-sequence of TM(S)T^{M}(S) is (Y1,Y2,,Ym)(Y_{1},Y_{2},...,Y_{m}), and the xx-sequence of TM(S)T^{M}(S) is (X1,X2,,Xm)(X_{1},X_{2},...,X_{m}). If iSi\notin S, then Xi<YiX_{i}<Y_{i}. Let’s construct a TT such that p(T)=Pp(T)=P. If the rr-th step of PP is an EE step, then the entry rr appears in the first row of TT. We mark it as (y1,y2,,ym)(y_{1},y_{2},...,y_{m}) in ascending order. So we have yiYiy_{i}\geq Y_{i} for all ii. If the rr-th step of PP is a NN step, then the entry rr appears in the second row of TT. We mark it as (x1,x2,,xm)(x_{1},x_{2},...,x_{m}) in ascending order. So we have xiXix_{i}\leq X_{i} for all ii. If iSi\notin S, then xiXi<Yiyix_{i}\leq X_{i}<Y_{i}\leq y_{i}. So TT in YT((m,m),S)YT((m,m),S) that satisfy T0TTM(S)T^{0}\leq T\leq T^{M}(S). Furthermore, p(T)=Pp(T)=P and p(T)p(T) is a bijection. This completes the proof. ∎

See Figure 4 for an example. The first step of p(T)p(T) is always NN since the x1x_{1} in mn\mathcal{B}_{m}^{n} has to be equal to 11.

For the cardinality of YT((m,m),S)YT((m,m),S), there is a determinant formula by the following result of Kreweras [9] in lattice path counting theory.

Lemma 2.4 ([9]).

Let P1=(a1,a2,,an)P_{1}=(a_{1},a_{2},...,a_{n}) and P2=(b1,b2,,bn)P_{2}=(b_{1},b_{2},...,b_{n}) be two fixed reverse partitions, where a1a2ana_{1}\leq a_{2}\leq\cdots\leq a_{n} and b1b2bnb_{1}\leq b_{2}\leq\cdots\leq b_{n}. Suppose P2P1P_{2}\leq P_{1}, i.e., biaib_{i}\leq a_{i} for 1in1\leq i\leq n. The number of partitions lying between P2P_{2} and P1P_{1} is

det((aibj+1ji+1))1i,jn.\det\left(\binom{a_{i}-b_{j}+1}{j-i+1}\right)_{1\leq i,j\leq n}.

For Lemma 2.4, the weighted counting of the partitions by area has been studied. A qq-analogous determinant formula was first obtained by Handa and Mohanty [7], and later proved by Gessel and Loehr [5] using an elegant involution argument.

Now we can give a determinant formula for f¯m(n)\overline{f}_{m}(n).

Proposition 2.5.

Follow the notation in Section 1. The cardinality f¯m(n)=#YT(mn)\overline{f}_{m}(n)=\#YT(\mathcal{B}_{m}^{n}) is the number of lattice paths from (0,0)(0,0) to (mn,mn)(mn,mn) that never go below the path N(EmNm)n1EmNm1N(E^{m}N^{m})^{n-1}E^{m}N^{m-1}. Moreover, we have

f¯m(n)=det((ai+1ji+1))1i,jnm1,\overline{f}_{m}(n)=\det\left(\binom{a_{i}+1}{j-i+1}\right)_{1\leq i,j\leq nm-1},

where

ai=hm,(h1)m+1ihm, 1hn1a_{i}=hm,\ \ (h-1)m+1\leq i\leq hm,\ 1\leq h\leq n-1

and

ai=mn,(n1)m+1inm1.a_{i}=mn,\ \ (n-1)m+1\leq i\leq nm-1.
Proof.

The first part follows by Proposition 2.3 with respect to S={jm+i2im,0jn1}S=\{jm+i\mid 2\leq i\leq m,0\leq j\leq n-1\}, where one can verify that

p(T0)=NmnEmn,μ(T0)=(0,0,,0)nm,p(T^{0})=N^{mn}E^{mn},\ \ \mu(T^{0})=\underbrace{(0,0,...,0)}_{nm},

and that

p(TM(S))=N(EmNm)n1EmNm1,p(T^{M}(S))=N(E^{m}N^{m})^{n-1}E^{m}N^{m-1},
μ(TM(S))=(0,m,,mm,2m,,2mm,,(n1)m,,(n1)mm,mn,,mnm1).\mu(T^{M}(S))=(0,\underbrace{m,...,m}_{m},\underbrace{2m,...,2m}_{m},...,\underbrace{(n-1)m,...,(n-1)m}_{m},\underbrace{mn,...,mn}_{m-1}).

The second part then follows by the first part and Lemma 2.4, in which we ignore the first element 0. ∎

We are interested in finding a nice formula for the generating function of f¯m(n)\overline{f}_{m}(n). Proposition 2.5 does not seem to help. Next we give a recursive formula as follows.

Proposition 2.6.

Follow the notation as above. Suppose m2m\geq 2 and n1n\geq 1. Then we have

f¯m(n)=(2mn1mn)i=1n1t=1mgt(i)(2m(ni)+mt1m(ni)1).\overline{f}_{m}(n)=\binom{2mn-1}{mn}-\sum_{i=1}^{n-1}\sum_{t=1}^{m}g_{t}(i)\binom{2m(n-i)+m-t-1}{m(n-i)-1}.

where,

gt(i)=(2mim+t1mi)j=1i1r=1mgr(j)(2m(ij)+tr1m(ij)1)g_{t}(i)=\binom{2mi-m+t-1}{mi}-\sum_{j=1}^{i-1}\sum_{r=1}^{m}g_{r}(j)\binom{2m(i-j)+t-r-1}{m(i-j)-1}

and the initial value are

gt(1)=(m+t1m), 1tm.g_{t}(1)=\binom{m+t-1}{m},\ \ 1\leq t\leq m.

Note that gm(i)=f¯m(i)g_{m}(i)=\overline{f}_{m}(i).

Proof.

We use the lattice path interpretation. The theorem is an application of the inclusion-exclusion principle. We only illustrate the idea by a simple example, since the detailed proof is conceptually simple but tedious to present.

Consider the counting of lattice paths for the case m=3,n=2m=3,n=2 in Figure 4. Firstly, we move (0,0)(0,0) and (6,6)(6,6) to (1,1)(-1,1) (point MM) and (5,7)(5,7) (point FF) respectively. Then we need to count lattice paths from MM to FF that never touch the lower red boundary NE3N3E3N2NE^{3}N^{3}E^{3}N^{2}. For this, we subtract from (6+66)\binom{6+6}{6} the number of bad lattice paths, i.e, those touch the boundary. We classify the bad paths by their first touching of the boundary, at the points AA, BB, CC and DD in Figure 4. This gives

f¯3(2)=(6+66)1×(5+65)1×(2+52)(3+13)(2+42)(3+23)(2+32)=281.\overline{f}_{3}(2)=\binom{6+6}{6}-1\times\binom{5+6}{5}-1\times\binom{2+5}{2}-\binom{3+1}{3}\binom{2+4}{2}-\binom{3+2}{3}\binom{2+3}{2}=281.

The subtracted terms are obtained in a similar way. Let us only explain the last term corresponding to bad paths first touching DD. The first term counts the number of ways to go from MM to DD without touching the boundary. This corresponds to going from (0,0)(0,0) to CC without going below the boundary, and is hence counted by gi(t)g_{i}(t), but becomes simply (3+23)\binom{3+2}{3} in the displayed picture; The second term counts the number of ways to go from DD to FF arbitrarily, which is clearly (2+32)\binom{2+3}{2}. ∎

In principle, Proposition 2.6 gives rise to a system of equations on the generating functions Gt(x)G_{t}(x) of gt(n)g_{t}(n) for t=1,2,,mt=1,2,\dots,m. Then one can solve for Gt(x)G_{t}(x) for all tt and hence F¯m(x)\overline{F}_{m}(x). Indeed, this idea allows us to work out F¯m(x)\overline{F}_{m}(x) for m=3,4m=3,4. Then in OEIS, we find that F¯3(x)\overline{F}_{3}(x) and F¯4(x)\overline{F}_{4}(x) correspond to [A213403] and [A213404] respectively. We further find that a similar system has been solved in lattice path counting theory, which will be discussed in the next section.

3. Lattice Path Enumeration

Before obtaining the main results of this section, we need to introduce some definitions and conclusions about symmetric functions.

3.1. Symmetric Functions

We need some basic results on symmetric functions. Most of them can be found in [13].

Let X=(x1,x2,)X=(x_{1},x_{2},...) be a set of indeterminates. A homogeneous symmetric function of degree nn (nn is a non-negative integer) over a commutative ring RR (with identity) is a formal power series

SF(X)=νcνXν,SF(X)=\sum_{\nu}c_{\nu}X^{\nu},

where ν\nu ranges over all weak compositions ν=(ν1,ν2,)\nu=(\nu_{1},\nu_{2},...) of nn, cνRc_{\nu}\in R, Xν=x1ν1x2ν2X^{\nu}=x_{1}^{\nu_{1}}x_{2}^{\nu_{2}}\cdots, and cν=cμc_{\nu}=c_{\mu} if μ\mu is obtained from ν\nu by permuting the entries.

Let Λn\Lambda^{n} be the set of all homogeneous symmetric functions of degree nn over RR. Then Λn\Lambda^{n} is a RR-module. If RR is the rational number field \mathbb{Q}, then Λn\Lambda^{n} is a vector space. Let Λ=Λ0Λ1\Lambda=\Lambda^{0}\oplus\Lambda^{1}\oplus\cdots be the vector space direct sum.

Let μ=(μ1,μ2,)\mu=(\mu_{1},\mu_{2},...)_{\geq} be a partition of nn, denoted μn\mu\vdash n. The length of μ\mu, denoted l(μ)l(\mu), is the number of nonzero entries of μi\mu_{i}. There are five important basis of Λn\Lambda^{n} indexed by partitions, namely:

  1. (1)

    The monomial symmetric functions: mμ=mμ(X)=νXνm_{\mu}=m_{\mu}(X)=\sum_{\nu}X^{\nu}, where ν\nu ranges over all rearrangements of the given partition μ\mu and m=1m_{\emptyset}=1.

  2. (2)

    The elementary symmetric functions: eμ=eμ1eμ2e_{\mu}=e_{\mu_{1}}e_{\mu_{2}}\cdots, where en=m1ne_{n}=m_{1^{n}} and e0=1e_{0}=1.

  3. (3)

    The complete symmetric functions: hμ=hμ1hμ2h_{\mu}=h_{\mu_{1}}h_{\mu_{2}}\cdots, where hn=κnmκh_{n}=\sum_{\kappa\vdash n}m_{\kappa} and h0=1h_{0}=1.

  4. (4)

    The power sum symmetric functions: pμ=pμ1pμ2p_{\mu}=p_{\mu_{1}}p_{\mu_{2}}\cdots, where pn=mnp_{n}=m_{n} and p0=1p_{0}=1.

  5. (5)

    The Schur functions: sμ=ν|μ|Kμνmνs_{\mu}=\sum_{\nu\vdash|\mu|}K_{\mu\nu}m_{\nu}, where KμνK_{\mu\nu} is the Kostka number.

We only use symmetric functions on a finite number of variables, say x1,,xnx_{1},\dots,x_{n}. One can treat xi=0x_{i}=0 for i>ni>n. Let Λn\Lambda_{n} be the set of all polynomials f(X)f(X) in [x1,,xn]\mathbb{Q}[x_{1},...,x_{n}] that are invariant under any permutation of the variables. Then f(X)f(X) is just a symmetric function on X=(x1,,xn,0,0,)X=(x_{1},...,x_{n},0,0,...).

The following is the classical definition of Schur functions in the variables (x1,,xn)(x_{1},...,x_{n}). See, e.g., [13, Theorem 7.15.1].

Lemma 3.1 ([13]).

Let μ=(μ1,μ2,)\mu=(\mu_{1},\mu_{2},...) be a partition of length l(μ)nl(\mu)\leq n. Let δ=(n1,n2,,0)\delta=(n-1,n-2,...,0). Then we have

sμ(x1,,xn)=aμ+δaδ,s_{\mu}(x_{1},...,x_{n})=\frac{a_{\mu+\delta}}{a_{\delta}},

where aμ+δ=det(xiμj+nj)i,j=1na_{\mu+\delta}=\det(x_{i}^{\mu_{j}+n-j})_{i,j=1}^{n} and aδ=det(xinj)i,j=1na_{\delta}=\det(x_{i}^{n-j})_{i,j=1}^{n}.

We denote the Vandermonde determinant by

detV(x1,,xn):=(1)n(n1)2aδ=1j<in(xixj).\det V(x_{1},...,x_{n}):=(-1)^{\frac{n(n-1)}{2}}a_{\delta}=\prod_{1\leq j<i\leq n}(x_{i}-x_{j}).

We also use the plethystic notation for symmetric functions to simplify the proof. A good reference to plethystic notation is [6]. Let E(x1,x2,)E(x_{1},x_{2},...) be a formal series of rational functions in the parameters x1,x2,x_{1},x_{2},.... We define the plethystic substitution of EE into pkp_{k}, denoted pk[E]p_{k}[E], by pk[E]=E(x1k,x2k,)p_{k}[E]=E(x_{1}^{k},x_{2}^{k},...). This is to distinguish pk[E]p_{k}[E] from the ordinary kk-th power sum in a set of variables EE.

Lemma 3.2 ([6]).

Let mm\in\mathbb{N}. We have

em[x1+x2++xn]=i1<i2<<imxi1xi2xim,\displaystyle e_{m}[x_{1}+x_{2}+\cdots+x_{n}]=\sum_{i_{1}<i_{2}<\cdots<i_{m}}x_{i_{1}}x_{i_{2}}\cdots x_{i_{m}},
em[EF]=i=0mei[E]emi[F],em[X]=(1)mhm[X].\displaystyle e_{m}[E-F]=\sum_{i=0}^{m}e_{i}[E]e_{m-i}[-F],\ \ \ \ e_{m}[-X]=(-1)^{m}h_{m}[X].
Lemma 3.3 (Section 7.6, [13]).

Suppose n1n\geq 1. The elementary symmetric functions and the complete symmetric functions have the following relationship:

i=0n(1)ieihni=0.\sum_{i=0}^{n}(-1)^{i}e_{i}h_{n-i}=0.

Now let’s give a result involving the Vandermonde determinant.

Lemma 3.4.

Let r1r\geq 1 and rr\in\mathbb{N}. Then

det(1w1w12w1r1w2w22w2r1ww2wr)=(1)(1)2sμ(w1,,w)detV(w1,,w)er(w1,,w),\displaystyle\det\left(\begin{array}[]{ccccc}1&w_{1}&\cdots&w_{1}^{\ell-2}&w_{1}^{-r}\\ 1&w_{2}&\cdots&w_{2}^{\ell-2}&w_{2}^{-r}\\ \vdots&\vdots&\ddots&\vdots&\vdots\\ 1&w_{\ell}&\cdots&w_{\ell}^{\ell-2}&w_{\ell}^{-r}\\ \end{array}\right)=\frac{(-1)^{(\ell-1)^{2}}s_{\mu}(w_{1},...,w_{\ell})\det V(w_{1},...,w_{\ell})}{e_{\ell}^{r}(w_{1},...,w_{\ell})},

where μ=(r1,r1,,r1)\mu=(r-1,r-1,...,r-1) with length (μ)=1\ell(\mu)=\ell-1.

Proof.

By direct computation, the left hand side becomes

(w1rwr)1det(w1rw1r+1w1+r21w2rw2r+1w2+r21wrwr+1w+r21)\displaystyle(w_{1}^{r}\cdots w_{\ell}^{r})^{-1}\det\left(\begin{array}[]{ccccc}w_{1}^{r}&w_{1}^{r+1}&\cdots&w_{1}^{\ell+r-2}&1\\ w_{2}^{r}&w_{2}^{r+1}&\cdots&w_{2}^{\ell+r-2}&1\\ \vdots&\vdots&\ddots&\vdots&\vdots\\ w_{\ell}^{r}&w_{\ell}^{r+1}&\cdots&w_{\ell}^{\ell+r-2}&1\\ \end{array}\right)
=\displaystyle= (1)(1)(2)2(w1rwr)1a(+r2,+r3,,r,0)(w1,,w)\displaystyle(-1)^{\frac{(\ell-1)(\ell-2)}{2}}(w_{1}^{r}\cdots w_{\ell}^{r})^{-1}a_{(\ell+r-2,\ell+r-3,...,r,0)}(w_{1},...,w_{\ell})
(by Lemma 3.1)=\displaystyle(\text{by Lemma \ref{SchurDeterm}})\quad= (1)(1)(2)2(w1rwr)1sμ(w1,,w)a(1,2,,1,0)(w1,,w)\displaystyle(-1)^{\frac{(\ell-1)(\ell-2)}{2}}(w_{1}^{r}\cdots w_{\ell}^{r})^{-1}s_{\mu}(w_{1},...,w_{\ell})\cdot a_{(\ell-1,\ell-2,...,1,0)}(w_{1},...,w_{\ell})
=\displaystyle= (1)(1)2sμ(w1,,w)detV(w1,,w)er(w1,,w).\displaystyle\frac{(-1)^{(\ell-1)^{2}}s_{\mu}(w_{1},...,w_{\ell})\det V(w_{1},...,w_{\ell})}{e_{\ell}^{r}(w_{1},...,w_{\ell})}.

This completes the proof. ∎

Lemma 3.5.

Let l1l\geq 1 and i0i\geq 0. Let λ=(i,i,,i)\lambda=(i,i,...,i) with length (λ)=1\ell(\lambda)=\ell-1. We have

hi(w11,w21,,w1)ei(w1,w2,,w)=sλ(w1,w2,,w).h_{i}(w_{1}^{-1},w_{2}^{-1},...,w_{\ell}^{-1})\cdot e_{\ell}^{i}(w_{1},w_{2},...,w_{\ell})=s_{\lambda}(w_{1},w_{2},...,w_{\ell}).
Proof.

Let δ=(1,2,,0)\delta=(\ell-1,\ell-2,...,0). By the classical definition of Schur functions in Lemma 3.1 with respect to λ=(i,i,,i,0)\lambda=(i,i,...,i,0), we have

sλ(w1,w2,,w)=aλ+δ(w1,w2,,w)aδ(w1,w2,,w)\displaystyle s_{\lambda}(w_{1},w_{2},...,w_{\ell})=\frac{a_{\lambda+\delta}(w_{1},w_{2},...,w_{\ell})}{a_{\delta}(w_{1},w_{2},...,w_{\ell})}
=\displaystyle= det(w1i+1w1i+2w1i+11w2i+1w2i+2w2i+11wi+1wi+2wi+11)(det(w11w12w11w21w22w21w1w2w1))1\displaystyle\det\left(\begin{array}[]{ccccc}w_{1}^{i+\ell-1}&w_{1}^{i+\ell-2}&\cdots&w_{1}^{i+1}&1\\ w_{2}^{i+\ell-1}&w_{2}^{i+\ell-2}&\cdots&w_{2}^{i+1}&1\\ \vdots&\vdots&\ddots&\vdots&\vdots\\ w_{\ell}^{i+\ell-1}&w_{\ell}^{i+\ell-2}&\cdots&w_{\ell}^{i+1}&1\\ \end{array}\right)\left(\det\left(\begin{array}[]{ccccc}w_{1}^{\ell-1}&w_{1}^{\ell-2}&\cdots&w_{1}&1\\ w_{2}^{\ell-1}&w_{2}^{\ell-2}&\cdots&w_{2}&1\\ \vdots&\vdots&\ddots&\vdots&\vdots\\ w_{\ell}^{\ell-1}&w_{\ell}^{\ell-2}&\cdots&w_{\ell}&1\\ \end{array}\right)\right)^{-1}
=\displaystyle= (w1w2w)i+1(w1w2w)1det(1w12w1i+11w22w2i+11w2wi+1)(det(1w12w111w22w211w2w1))1\displaystyle\frac{(w_{1}w_{2}\cdots w_{\ell})^{i+\ell-1}}{(w_{1}w_{2}\cdots w_{\ell})^{\ell-1}}\det\left(\begin{array}[]{cccc}1&\cdots&w_{1}^{2-\ell}&w_{1}^{-i-\ell+1}\\ 1&\cdots&w_{2}^{2-\ell}&w_{2}^{-i-\ell+1}\\ \vdots&\ddots&\vdots&\vdots\\ 1&\cdots&w_{\ell}^{2-\ell}&w_{\ell}^{-i-\ell+1}\\ \end{array}\right)\left(\det\left(\begin{array}[]{cccc}1&\cdots&w_{1}^{2-\ell}&w_{1}^{1-\ell}\\ 1&\cdots&w_{2}^{2-\ell}&w_{2}^{1-\ell}\\ \vdots&\ddots&\vdots&\vdots\\ 1&\cdots&w_{\ell}^{2-\ell}&w_{\ell}^{1-\ell}\\ \end{array}\right)\right)^{-1}
=\displaystyle= (w1w2w)idet((w11)i+1w11(w21)i+1w21(w1)i+1w1)(det((w11)1w11(w21)1w21(w1)1w1))1\displaystyle(w_{1}w_{2}\cdots w_{\ell})^{i}\det\left(\begin{array}[]{cccc}(w_{1}^{-1})^{i+\ell-1}&\cdots&w_{1}^{-\ell}&1\\ (w_{2}^{-1})^{i+\ell-1}&\cdots&w_{2}^{-\ell}&1\\ \vdots&\ddots&\vdots&\vdots\\ (w_{\ell}^{-1})^{i+\ell-1}&\cdots&w_{\ell}^{-\ell}&1\\ \end{array}\right)\left(\det\left(\begin{array}[]{cccc}(w_{1}^{-1})^{\ell-1}&\cdots&w_{1}^{-\ell}&1\\ (w_{2}^{-1})^{\ell-1}&\cdots&w_{2}^{-\ell}&1\\ \vdots&\ddots&\vdots&\vdots\\ (w_{\ell}^{-1})^{\ell-1}&\cdots&w_{\ell}^{-\ell}&1\\ \end{array}\right)\right)^{-1}
=\displaystyle= (w1w2w)isi(w11,w21,,w1)\displaystyle(w_{1}w_{2}\cdots w_{\ell})^{i}\cdot s_{i}(w_{1}^{-1},w_{2}^{-1},...,w_{\ell}^{-1})
=\displaystyle= ei(w1,w2,,w)hi(w11,w21,,w1).\displaystyle e_{\ell}^{i}(w_{1},w_{2},...,w_{\ell})h_{i}(w_{1}^{-1},w_{2}^{-1},...,w_{\ell}^{-1}).

The last “==” is due to the fact that si=his_{i}=h_{i}. ∎

We will use several variations of the well-known Vandermonde’s identity

i=0n(ai)(bni)=(a+bn).\sum_{i=0}^{n}\binom{a}{i}\binom{b}{n-i}=\binom{a+b}{n}.

Vandermonde’s identity can be easily proved by equating coefficients of xnx^{n} on both sides of the equation (1+x)a(1+x)b=(1+x)a+b(1+x)^{a}\cdot(1+x)^{b}=(1+x)^{a+b}. This trick is sufficient for our purpose.

Lemma 3.6.

Let jj\leq\ell and j+1rj+1\leq r. We have

m=0r1j(1)j+r+m+1(jr1mj)=(1)rj1(j1rj1).\sum_{m=0}^{r-1-j}(-1)^{j+r+m+1}\binom{\ell-j}{r-1-m-j}=(-1)^{r-j-1}\binom{\ell-j-1}{r-j-1}.
Lemma 3.7.

Let jj\leq\ell, i+jr1i+j\leq r-1 and 0\ell\geq 0. We have

m=ir1j(1)r+m+i+j(jr1mj)(m+li+)=(1)i1(r1i+j).\sum_{m=i}^{r-1-j}(-1)^{r+m+i+j}\binom{\ell-j}{r-1-m-j}\binom{m+l}{i+\ell}=(-1)^{i-1}\binom{r-1}{i+j}.
Lemma 3.8.

Let jj\leq\ell, j+1rj+1\leq r and 0\ell\geq 0. We have

m=0r1j(1)j+r+m(jr1mj)(m+)=(r1j).\sum_{m=0}^{r-1-j}(-1)^{j+r+m}\binom{\ell-j}{r-1-m-j}\binom{m+\ell}{\ell}=-\binom{r-1}{j}.
Sketched proofs of Lemmas 3.6, 3.7 and 3.8.

Lemma 3.6 follows by equating the coefficients of xrj1x^{r-j-1} on both sides of the identity

k=0j(jk)xkk0(x)k=(1+x)j(1+x)=(1+x)j1=k=0j1(j1k)xk.\displaystyle\sum_{k=0}^{\ell-j}\binom{\ell-j}{k}x^{k}\cdot\sum_{k\geq 0}(-x)^{k}=\frac{(1+x)^{\ell-j}}{(1+x)}=(1+x)^{\ell-j-1}=\sum_{k=0}^{\ell-j-1}\binom{\ell-j-1}{k}x^{k}.

Lemma 3.7 follows by equating the coefficients of xr1ijx^{r-1-i-j} on both sides of the identity

k0(i++ki+j)(x)kk=0j(jk)(x)k=(1+x)j(1+x)i++1=1(1+x)i+j+1=k0(i+j+ki+)(x)k.\displaystyle\sum_{k\geq 0}\binom{i+\ell+k}{i+j}(-x)^{k}\cdot\sum_{k=0}^{\ell-j}\binom{\ell-j}{k}(-x)^{k}=\frac{(1+x)^{\ell-j}}{(1+x)^{i+\ell+1}}=\frac{1}{(1+x)^{i+j+1}}=\sum_{k\geq 0}\binom{i+j+k}{i+\ell}(-x)^{k}.

Lemma 3.8 follows in a similar way, and we omit the proof. ∎

3.2. Lattice Paths

In this subsection, we focus on the counting of lattice paths from (0,0)(0,0) to (m,n)(m,n) that never go above a given path PP by using steps N=(0,1)N=(0,1) and E=(1,0)E=(1,0). For example, let’s flip the paths discussed in Section 2. Then f¯m(n)\overline{f}_{m}(n) counts the number of lattice paths from (0,0)(0,0) to (mn1,mn)(mn-1,mn) that never go above the path (NmEm)n1NmEm1(N^{m}E^{m})^{n-1}N^{m}E^{m-1}. We use the convention f¯m(0)=1\overline{f}_{m}(0)=1. When m=1m=1, n0f¯1(n)xn=C(x)\sum_{n\geq 0}\overline{f}_{1}(n)x^{n}=C(x) is the Catalan generating function.

Bonin, de Mier and Noy ([3]) made a connection between lattice paths and matroids. They defined the Tutte polynomial of a lattice path PP as follows.

t(P;z,y)=σ:not go above Pzi(σ)ye(σ),t(P;z,y)=\sum_{\sigma:\text{not go above }P}z^{i(\sigma)}y^{e(\sigma)},

where the sum ranges over all lattice paths σ\sigma that never go above PP, i(σ)i(\sigma) is the number of common NN steps of σ\sigma and PP, and e(σ)e(\sigma) is the number of common EE steps of σ\sigma and PP before the first NN step. In particular, t(P;1,1)t(P;1,1) is the number of all paths that never go above PP.

We need the following result, which was first obtained from Matroid Theory.

Lemma 3.9 ([3]).

Let PNPN be the path obtained from PP by appending a NN step, and PEPE be the path obtained similarly. Then we have

t(PN;z,y)=zt(P;z,y);\displaystyle t(PN;z,y)=z\cdot t(P;z,y);
t(PE;z,y)=zz1t(P;z,y)+(yzz1)t(P;1,y).\displaystyle t(PE;z,y)=\frac{z}{z-1}\cdot t(P;z,y)+\left(y-\frac{z}{z-1}\right)\cdot t(P;1,y).

Readers are invited to find a combinatorial proof of the above lemma.

Let q(n)q(n) be the number of lattice paths from (0,0)(0,0) to (n,kn)(\ell n,kn) that never go above the path (NkE)n(N^{k}E^{\ell})^{n}, where k,k,\ell\in\mathbb{P} and nn\in\mathbb{N}. Using Lemma 3.9, de Mier and Noy obtained a nice expression of the generating function Q(x)=n0q(n)xnQ(x)=\sum_{n\geq 0}q(n)x^{n}. See [10, Theorem 1] or Corollary 3.11 below. For completeness, we briefly describe their proof process as follows.

Let Pn=(NkE)nP_{n}=(N^{k}E^{\ell})^{n} and An=An(z,y)=t(Pn;z,y)A_{n}=A_{n}(z,y)=t(P_{n};z,y), where A0=1A_{0}=1. Define the operator Φ\Phi by

ΦA(z,y)=zz1A(z,y)+(yzz1)A(1,y).\Phi A(z,y)=\frac{z}{z-1}A(z,y)+\left(y-\frac{z}{z-1}\right)A(1,y).

Then Lemma 3.9 gives An+1=Φ(zkAn)A_{n+1}=\Phi^{\ell}(z^{k}A_{n}). For each nn\in\mathbb{N} and 1i1\leq i\leq\ell, define the polynomials

Bi,n=Φi(zkAn(z,y)),Ci,n=Bi,n(1,y),B_{i,n}=\Phi^{i}\left(z^{k}A_{n}(z,y)\right),\ \ C_{i,n}=B_{i,n}(1,y),

where we set C0,n(y)=An(1,y)C_{0,n}(y)=A_{n}(1,y). Thus we have B,n=An+1B_{\ell,n}=A_{n+1} and C0,n(1)=An(1,1)C_{0,n}(1)=A_{n}(1,1). Denote their generating functions by

A=A(x)=n0Anxn,Ci=Ci(x)=n0Ci,nxn, 0i.A=A(x)=\sum_{n\geq 0}A_{n}x^{n},\ \ \ C_{i}=C_{i}(x)=\sum_{n\geq 0}C_{i,n}x^{n},\ \ \ 0\leq i\leq\ell.

Based on the above definitions, we have

B1,n=zz1zkAn+(yzz1)C0,n,\displaystyle B_{1,n}=\frac{z}{z-1}z^{k}A_{n}+\left(y-\frac{z}{z-1}\right)C_{0,n},
B2,n=zz1B1,n+(yzz1)C1,n,\displaystyle B_{2,n}=\frac{z}{z-1}B_{1,n}+\left(y-\frac{z}{z-1}\right)C_{1,n},
\displaystyle... (2)
B,n=zz1B1,n+(yzz1)C1,n,\displaystyle B_{\ell,n}=\frac{z}{z-1}B_{\ell-1,n}+\left(y-\frac{z}{z-1}\right)C_{\ell-1,n},
An+1=B,n.\displaystyle A_{n+1}=B_{\ell,n}.

In the above equations, we start from the last equation and repeatedly replace Bi,nB_{i,n} from the previous equation. We have

n0An+1xn=A1x=zk+(z1)A+(yzyz)i=1zi1(z1)iCi.\sum_{n\geq 0}A_{n+1}x^{n}=\frac{A-1}{x}=\frac{z^{k+\ell}}{(z-1)^{\ell}}A+(yz-y-z)\sum_{i=1}^{\ell}\frac{z^{i-1}}{(z-1)^{i}}C_{\ell-i}.

Simplifying and setting y=1y=1 give

A((z1)xzk+)=(z1)xi=1zi+1(z1)iCi.A\left((z-1)^{\ell}-xz^{k+\ell}\right)=(z-1)^{\ell}-x\sum_{i=1}^{\ell}z^{i+1}(z-1)^{\ell-i}C_{\ell-i}. (3)

This functional equation can be solved by the well-known kernel method.

The factor ((z1)xzk+)\left((z-1)^{\ell}-xz^{k+\ell}\right) on the left hand side is called the kennel. As a polynomial in zz, it has k+k+\ell roots. By Puiseus’s theorem, [13, Chapter 6], these roots can be treated as elements in the field of fractional Laurent series

fra((x))={nNan,Mxn/M:an,M,N,M}.\mathbb{C}^{\text{fra}}((x))=\Big{\{}\sum_{n\geq N}a_{n,M}x^{n/M}:a_{n,M}\in\mathbb{C},N\in\mathbb{Z},\ M\in\mathbb{P}\Big{\}}.

Moreover, by [13, Proposition 6.1.8], exactly \ell roots of the kernel are fractional power series. Denote them by w1(x),w2(x)w_{1}(x),w_{2}(x),,w(x)...,w_{\ell}(x). Then wi(0)=1w_{i}(0)=1 for 1i1\leq i\leq\ell.

The substitution z=wjz=w_{j} in Equation (3) is valid for 1j1\leq j\leq\ell. This gives a system of \ell linear equations in CjC_{j}:

i=1wji1(wj1)ixCi=(wj1), 1j.\sum_{i=1}^{\ell}w_{j}^{i-1}(w_{j}-1)^{\ell-i}xC_{\ell-i}=(w_{j}-1)^{\ell},\ \ 1\leq j\leq\ell. (4)

Because Q(x)=C0=n0An(1,1)xnQ(x)=C_{0}=\sum_{n\geq 0}A_{n}(1,1)x^{n}, de Mier and Noy only obtained a nice expression for C0C_{0} by using Lagrange’s interpolation formula and an identity in [13, Exercise 7.4]. We find nice expressions of CrC_{r} for all rr.

Our intermediate result is the following.

Theorem 3.10.

Let \ell\in\mathbb{P} and 1r1\leq r\leq\ell. Let w¯i=wi1wi\overline{w}_{i}=\frac{w_{i}-1}{w_{i}} for 1i1\leq i\leq\ell. Then we have

xCr=m=0r1er1m(w1¯,,w¯)(i=0m(1)r+m+i(m+i+)sμ(w1,,w)ei+(1)r+m+1e),\displaystyle xC_{\ell-r}=\sum_{m=0}^{r-1}e_{r-1-m}(\overline{w_{1}},\cdots,\overline{w_{\ell}})\left(\sum_{i=0}^{m}(-1)^{r+m+i}\binom{m+\ell}{i+\ell}\frac{s_{\mu}(w_{1},...,w_{\ell})}{e_{\ell}^{i}}+(-1)^{r+m+1}e_{\ell}\right), (5)

where e=e(w1,,w)e_{\ell}=e_{\ell}(w_{1},...,w_{\ell}) and μ=(i,i,,i)\mu=(i,i,...,i) with length (μ)=1\ell(\mu)=\ell-1.

Proof.

By the system (4), we have

i=01(wiwj1)ixCi1=wj1, 1j.\sum_{i=0}^{\ell-1}\left(\frac{w_{i}}{w_{j}-1}\right)^{i}xC_{\ell-i-1}=w_{j}-1,\ \ 1\leq j\leq\ell. (6)

The left side of equations (6) can be regarded as the result of evaluating the polynomial T(X)=i=01(xCi1)XiT(X)=\sum_{i=0}^{\ell-1}(xC_{\ell-i-1})X^{i} of degree 1\ell-1 at X=Xj=wjwj1X=X_{j}=\frac{w_{j}}{w_{j}-1} for 1j1\leq j\leq\ell. That is, T(Xj)=wj1T(X_{j})=w_{j}-1 for 1j1\leq j\leq\ell. Then using Lagrange’s interpolation formulas, we have

T(X)=j=1ijXXiXjXiT(Xj).T(X)=\sum_{j=1}^{\ell}\prod_{i\neq j}\frac{X-X_{i}}{X_{j}-X_{i}}T(X_{j}).

This is the same as in [10] up to here.

Now consider the coefficient of Xr1X^{r-1} in T(X)T(X). We obtain

xCr\displaystyle xC_{\ell-r} =j=1ij1XjXi((1)r1k1<k2<<kr1k1,,kr1jX1X2XXk1Xkr1Xj)T(Xj)\displaystyle=\sum_{j=1}^{\ell}\prod_{i\neq j}\frac{1}{X_{j}-X_{i}}\left((-1)^{\ell-r}\sum_{1\leq k_{1}<k_{2}<\cdots<k_{r-1}\leq\ell\atop k_{1},...,k_{r-1}\neq j}\frac{X_{1}X_{2}\cdots X_{\ell}}{X_{k_{1}}\cdots X_{k_{r-1}}X_{j}}\right)T(X_{j})
=(1)rw1w2w(w11)(w1)j=1ij(wj1)(wi1)wiwj\displaystyle=(-1)^{\ell-r}\frac{w_{1}w_{2}\cdots w_{\ell}}{(w_{1}-1)\cdots(w_{\ell}-1)}\sum_{j=1}^{\ell}\prod_{i\neq j}\frac{(w_{j}-1)(w_{i}-1)}{w_{i}-w_{j}}\cdot
(1k1<k2<<kr1k1,,kr1j(wk11)(wkr11)(wj1)wk1wk2wkr1wj)(wj1)\displaystyle\ \ \cdot\left(\sum_{1\leq k_{1}<k_{2}<\cdots<k_{r-1}\leq\ell\atop k_{1},...,k_{r-1}\neq j}\frac{(w_{k_{1}}-1)\cdots(w_{k_{r-1}}-1)(w_{j}-1)}{w_{k_{1}}w_{k_{2}}\cdots w_{k_{r-1}}w_{j}}\right)(w_{j}-1)
=(1)rw1wj=1ij1wjwi(1k1<<kr1k1,,kr1j(wk11)(wkr11)wk1wk2wkr1wj)(wj1)\displaystyle=(-1)^{\ell-r}w_{1}\cdots w_{\ell}\sum_{j=1}^{\ell}\prod_{i\neq j}\frac{-1}{w_{j}-w_{i}}\cdot\left(\sum_{1\leq k_{1}<\cdots<k_{r-1}\leq\ell\atop k_{1},...,k_{r-1}\neq j}\frac{(w_{k_{1}}-1)\cdots(w_{k_{r-1}}-1)}{w_{k_{1}}w_{k_{2}}\cdots w_{k_{r-1}}w_{j}}\right)(w_{j}-1)^{\ell}
=(1)r+1w1wj=1ij1wjwi(1k1<<kr1k1,,kr1j(wk11)(wkr11)wk1wk2wkr1)(wj1)wj.\displaystyle=(-1)^{r+1}w_{1}\cdots w_{\ell}\sum_{j=1}^{\ell}\prod_{i\neq j}\frac{1}{w_{j}-w_{i}}\cdot\left(\sum_{1\leq k_{1}<\cdots<k_{r-1}\leq\ell\atop k_{1},...,k_{r-1}\neq j}\frac{(w_{k_{1}}-1)\cdots(w_{k_{r-1}}-1)}{w_{k_{1}}w_{k_{2}}\cdots w_{k_{r-1}}}\right)\frac{(w_{j}-1)^{\ell}}{w_{j}}.

Denote by

g(wj)=(1k1<<kr1k1,,kr1j(wk11)(wkr11)wk1wk2wkr1)(wj1)wj.g(w_{j})=\left(\sum_{1\leq k_{1}<\cdots<k_{r-1}\leq\ell\atop k_{1},...,k_{r-1}\neq j}\frac{(w_{k_{1}}-1)\cdots(w_{k_{r-1}}-1)}{w_{k_{1}}w_{k_{2}}\cdots w_{k_{r-1}}}\right)\frac{(w_{j}-1)^{\ell}}{w_{j}}.

Then by definition of the Vandermonde determinant, we have

xCr\displaystyle xC_{\ell-r} =(1)r+1w1wj=1ij1wjwig(wj)\displaystyle=(-1)^{r+1}w_{1}\cdots w_{\ell}\sum_{j=1}^{\ell}\prod_{i\neq j}\frac{1}{w_{j}-w_{i}}\cdot g(w_{j})
=(1)r+1w1wj=1(1)jg(wj)detV(w1,,wwj)detV(w1,,w),\displaystyle=(-1)^{r+1}w_{1}\cdots w_{\ell}\sum_{j=1}^{\ell}\frac{(-1)^{\ell-j}g(w_{j})\det V(w_{1},...,w_{\ell}\setminus w_{j})}{\det V(w_{1},...,w_{\ell})},

where detV(w1,,wwj)\det V(w_{1},...,w_{\ell}\setminus w_{j}) represents the Vandermonde determinant of the variables
w1,,wj1,wj+1,,ww_{1},...,w_{j-1},w_{j+1},...,w_{\ell}. Furthermore, we have

G=j=1(1)+jg(wj)detV(w1,,wwj)=det(1w1w12g(w1)1w2w22g(w1)1w3w32g(w3)1ww2g(w)).\displaystyle G=\sum_{j=1}^{\ell}(-1)^{\ell+j}g(w_{j})\det V(w_{1},...,w_{\ell}\setminus w_{j})=\det\left(\begin{array}[]{ccccc}1&w_{1}&\cdots&w_{1}^{\ell-2}&g(w_{1})\\ 1&w_{2}&\cdots&w_{2}^{\ell-2}&g(w_{1})\\ 1&w_{3}&\cdots&w_{3}^{\ell-2}&g(w_{3})\\ \vdots&\vdots&\ddots&\vdots&\vdots\\ 1&w_{\ell}&\cdots&w_{\ell}^{\ell-2}&g(w_{\ell})\\ \end{array}\right).

Observe that GG becomes i) 0 if g(w)g(w) is 1,w,,w21,w,\dots,w^{\ell-2}; ii) detV(w1,,w)\det V(w_{1},...,w_{\ell}) if g(w)=w1g(w)=w^{\ell-1}; iii) (1)(1)2sμdetV(w1,,w)ei+1(w1,,w)\frac{(-1)^{(\ell-1)^{2}}s_{\mu}\det V(w_{1},...,w_{\ell})}{e_{\ell}^{i+1}(w_{1},...,w_{\ell})} (by Lemma 3.4) if g(w)=wi1g(w)=w^{-i-1} for i0i\geq 0. Thus we shall write g(wj)g(w_{j}) as a polynomial in wjw_{j} with coefficients symmetric in w1,,ww_{1},\dots,w_{\ell}.

To this end, we use plethestic notation by treating w¯i=wi1wi\overline{w}_{i}=\frac{w_{i}-1}{w_{i}} for 1i1\leq i\leq\ell as variables. Then wi=w¯i1w¯iw_{i}=\frac{\overline{w}_{i}-1}{\overline{w}_{i}} is not a variable. Let w¯=w1¯+w2¯++w¯\overline{w}=\overline{w_{1}}+\overline{w_{2}}+\cdots+\overline{w_{\ell}}. By Lemma 3.2, we have

g(wj)\displaystyle g(w_{j}) =er1[w¯w¯j](wj1)wj=m=0r1er1m[w¯]em[wj¯](wj1)wj\displaystyle=e_{r-1}[\overline{w}-\overline{w}_{j}]\cdot\frac{(w_{j}-1)^{\ell}}{w_{j}}=\sum_{m=0}^{r-1}e_{r-1-m}[\overline{w}]\cdot e_{m}[-\overline{w_{j}}]\cdot\frac{(w_{j}-1)^{\ell}}{w_{j}}
=m=0r1er1m[w¯](1)m(wj1)m+wjm+1=m=0r1(1)mer1m[w¯]i=0m+(1)i(m+i)wji1.\displaystyle=\sum_{m=0}^{r-1}e_{r-1-m}[\overline{w}]\cdot(-1)^{m}\cdot\frac{(w_{j}-1)^{m+\ell}}{w_{j}^{m+1}}=\sum_{m=0}^{r-1}(-1)^{m}e_{r-1-m}[\overline{w}]\sum_{i=0}^{m+\ell}(-1)^{i}\binom{m+\ell}{i}w_{j}^{\ell-i-1}.

Therefore, we obtain

G\displaystyle G =m=0r1(1)mer1m[w¯](i=0m(1)+i+(1)2(m+i+)sμdetV(w1,,w)ei+1+detV(w1,,w)).\displaystyle=\sum_{m=0}^{r-1}(-1)^{m}e_{r-1-m}[\overline{w}]\left(\sum_{i=0}^{m}(-1)^{\ell+i+(\ell-1)^{2}}\binom{m+\ell}{i+\ell}\frac{s_{\mu}\det V(w_{1},...,w_{\ell})}{e_{\ell}^{i+1}}+\det V(w_{1},...,w_{\ell})\right).

Now we have

xCr\displaystyle xC_{\ell-r} =(1)r+1w1w(m=0r1(1)mer1m[w¯](i=0m(1)i+1(m+i+)sμei+1+1))\displaystyle=(-1)^{r+1}w_{1}\cdots w_{\ell}\left(\sum_{m=0}^{r-1}(-1)^{m}e_{r-1-m}[\overline{w}]\left(\sum_{i=0}^{m}(-1)^{i+1}\binom{m+\ell}{i+\ell}\frac{s_{\mu}}{e_{\ell}^{i+1}}+1\right)\right)
=m=0r1er1m[w¯](i=0m(1)r+m+i(m+i+)sμei+(1)r+m+1e).\displaystyle=\sum_{m=0}^{r-1}e_{r-1-m}[\overline{w}]\left(\sum_{i=0}^{m}(-1)^{r+m+i}\binom{m+\ell}{i+\ell}\frac{s_{\mu}}{e_{\ell}^{i}}+(-1)^{r+m+1}e_{\ell}\right).

This completes the proof. ∎

Now we present our proof of Theorem 1.2.

Proof of Theorem 1.2.

For 1r1\leq r\leq\ell, let fr(n)f_{r}(n) be the number of lattice paths from (0,0)(0,0) to (nr,kn)(\ell n-r,kn) that never go above the path (NkE)n1NkEr(N^{k}E^{\ell})^{n-1}N^{k}E^{\ell-r}. We consider the generating function Fr(x)=n0fr(n)xnF_{r}(x)=\sum_{n\geq 0}f_{r}(n)x^{n} with the convention fr(0)=1f_{r}(0)=1. Then it is easy to see that Fr(x)=xCr+1F_{r}(x)=xC_{\ell-r}+1 for 1r1\leq r\leq\ell.

To simplify Equation (5), we need to write er1m[w¯]e_{r-1-m}[\overline{w}] in terms of ei=ei(w1,,w)e_{i}=e_{i}(w_{1},\dots,w_{\ell}). Consider the generating function of ei[w¯]e_{i}[\overline{w}] as follows.

i=1(1+tw¯i)\displaystyle\prod_{i=1}^{\ell}(1+t\overline{w}_{i}) =i=1(1+t(1wi1))=i=1(1+ttwi1)\displaystyle=\prod_{i=1}^{\ell}\left(1+t(1-w_{i}^{-1})\right)=\prod_{i=1}^{\ell}(1+t-t\cdot w_{i}^{-1})
=(1+t)i=1(1t1+twi1)\displaystyle=(1+t)^{\ell}\prod_{i=1}^{\ell}\left(1-\frac{t}{1+t}w_{i}^{-1}\right)
=(1+t)j0ej(w11,w21,w1)(1)j(t1+t)j\displaystyle=(1+t)^{\ell}\sum_{j\geq 0}e_{j}(w_{1}^{-1},w_{2}^{-1},...w_{\ell}^{-1})(-1)^{j}\left(\frac{t}{1+t}\right)^{j}
=j0ej(w11,w21,,w1)(1)jtj(1+t)j.\displaystyle=\sum_{j\geq 0}e_{j}(w_{1}^{-1},w_{2}^{-1},...,w_{\ell}^{-1})(-1)^{j}t^{j}(1+t)^{\ell-j}.

Taking the coefficient of trm1t^{r-m-1} on both sides of the above equation, we obtain

er1m[w¯]\displaystyle e_{r-1-m}[\overline{w}] =[trm1j]j=0rm1(1)jej(w11,w21,,w1)(1+t)j\displaystyle=[t^{r-m-1-j}]\sum_{j=0}^{r-m-1}(-1)^{j}e_{j}(w_{1}^{-1},w_{2}^{-1},...,w_{\ell}^{-1})(1+t)^{\ell-j} (7)
=j=0rm1(1)j(jrm1j)ej(w11,w21,,w1)\displaystyle=\sum_{j=0}^{r-m-1}(-1)^{j}\binom{\ell-j}{r-m-1-j}e_{j}(w_{1}^{-1},w_{2}^{-1},...,w_{\ell}^{-1})
=e1j=0rm1(1)j(jrm1j)ej(w1,w2,,w).\displaystyle=e_{\ell}^{-1}\sum_{j=0}^{r-m-1}(-1)^{j}\binom{\ell-j}{r-m-1-j}e_{\ell-j}(w_{1},w_{2},...,w_{\ell}).

By Equation (5), we first obtain

m=0r1er1m[w¯](1)r+m+1e=\displaystyle\sum_{m=0}^{r-1}e_{r-1-m}[\overline{w}](-1)^{r+m+1}e_{\ell}= m=0r1j=0r1m(1)r+m+1+j(jr1mj)ej(w1,w2,,w)\displaystyle\sum_{m=0}^{r-1}\sum_{j=0}^{r-1-m}(-1)^{r+m+1+j}\binom{\ell-j}{r-1-m-j}e_{\ell-j}(w_{1},w_{2},...,w_{\ell})
=\displaystyle= j=0r1m=0r1j(1)j+r+m+1(jr1mj)ej(w1,w2,,w)\displaystyle\sum_{j=0}^{r-1}\sum_{m=0}^{r-1-j}(-1)^{j+r+m+1}\binom{\ell-j}{r-1-m-j}e_{\ell-j}(w_{1},w_{2},...,w_{\ell})
(By Lemma 3.6.)=\displaystyle(\text{By Lemma }\ref{CTbinomTT1}.)\ \ \ = j=0r1(1)rj1(j1rj1)ej(w1,w2,,w)\displaystyle\sum_{j=0}^{r-1}(-1)^{r-j-1}\binom{\ell-j-1}{r-j-1}e_{\ell-j}(w_{1},w_{2},...,w_{\ell})
=\displaystyle= i=0r1(1)i(r+ii)er+1+i(w1,w2,,w).\displaystyle\sum_{i=0}^{r-1}(-1)^{i}\binom{\ell-r+i}{i}e_{\ell-r+1+i}(w_{1},w_{2},...,w_{\ell}). (8)

By Fr(x)=xCr+1F_{r}(x)=xC_{\ell-r}+1 and observing Equation (5), we assert that

1+m=0r1er1m[w¯](i=0m(1)r+m+i(m+i+)sμei)=0.\displaystyle 1+\sum_{m=0}^{r-1}e_{r-1-m}[\overline{w}]\left(\sum_{i=0}^{m}(-1)^{r+m+i}\binom{m+\ell}{i+\ell}\frac{s_{\mu}}{e_{\ell}^{i}}\right)=0. (9)

i.e.,

1+m=0r1(j=0rm1(1)j(jrm1j)ej(w11,,w1))(i=0m(1)r+m+i(m+i+)sμei)=0.\displaystyle 1+\sum_{m=0}^{r-1}\left(\sum_{j=0}^{r-m-1}(-1)^{j}\binom{\ell-j}{r-m-1-j}e_{j}(w_{1}^{-1},...,w_{\ell}^{-1})\right)\left(\sum_{i=0}^{m}(-1)^{r+m+i}\binom{m+\ell}{i+\ell}\frac{s_{\mu}}{e_{\ell}^{i}}\right)=0.

We first consider the case i=0i=0 on the left side of the above equation. We have

1+m=0r1(j=0rm1(1)j(jrm1j)ej(w11,,w1))((1)r+m(m+))\displaystyle 1+\sum_{m=0}^{r-1}\left(\sum_{j=0}^{r-m-1}(-1)^{j}\binom{\ell-j}{r-m-1-j}e_{j}(w_{1}^{-1},...,w_{\ell}^{-1})\right)\left((-1)^{r+m}\binom{m+\ell}{\ell}\right)
=\displaystyle= 1+j=0r1m=0r1j(1)r+m+j(jr1mj)(m+)ej(w11,,w1)\displaystyle 1+\sum_{j=0}^{r-1}\sum_{m=0}^{r-1-j}(-1)^{r+m+j}\binom{\ell-j}{r-1-m-j}\binom{m+\ell}{\ell}e_{j}(w_{1}^{-1},...,w_{\ell}^{-1})
=\displaystyle= 1+j=0r1(r1j)ej(w11,,w1)(By Lemma 3.8.)\displaystyle 1+\sum_{j=0}^{r-1}-\binom{r-1}{j}e_{j}(w_{1}^{-1},...,w_{\ell}^{-1})\ \ \ (\text{By Lemma }\ref{CTbinomTT2}.)
=\displaystyle= j=1r1(r1j)ej(w11,,w1).\displaystyle-\sum_{j=1}^{r-1}\binom{r-1}{j}e_{j}(w_{1}^{-1},...,w_{\ell}^{-1}).

Now we consider the case i>0i>0. By Equation (9), we have

m=1r1j=0rm1i=1m(1)r+m+i+j(jrm1j)(m+i+)ej(w11,,w1)sμei\displaystyle\sum_{m=1}^{r-1}\sum_{j=0}^{r-m-1}\sum_{i=1}^{m}(-1)^{r+m+i+j}\binom{\ell-j}{r-m-1-j}\binom{m+\ell}{i+\ell}e_{j}(w_{1}^{-1},...,w_{\ell}^{-1})\frac{s_{\mu}}{e_{\ell}^{i}}
=\displaystyle= i=1r1j=0r1im=ir1j(1)r+m+i+j(jrm1j)(m+i+)ej(w11,,w1)sμei\displaystyle\sum_{i=1}^{r-1}\sum_{j=0}^{r-1-i}\sum_{m=i}^{r-1-j}(-1)^{r+m+i+j}\binom{\ell-j}{r-m-1-j}\binom{m+\ell}{i+\ell}e_{j}(w_{1}^{-1},...,w_{\ell}^{-1})\frac{s_{\mu}}{e_{\ell}^{i}}
=\displaystyle= i=1r1j=0r1i(1)i1(r1j+i)ej(w11,,w1)sμei.(By Lemma 3.7.)\displaystyle\sum_{i=1}^{r-1}\sum_{j=0}^{r-1-i}(-1)^{i-1}\binom{r-1}{j+i}e_{j}(w_{1}^{-1},...,w_{\ell}^{-1})\frac{s_{\mu}}{e_{\ell}^{i}}.\ \ \ (\text{By Lemma }\ref{CTbinomTT3}.)
=\displaystyle= i=1r1j=0r1i(1)i1(r1j+i)ej(w11,,w1)hi(w11,,w1)(By Lemma 3.5.)\displaystyle\sum_{i=1}^{r-1}\sum_{j=0}^{r-1-i}(-1)^{i-1}\binom{r-1}{j+i}e_{j}(w_{1}^{-1},...,w_{\ell}^{-1})h_{i}(w_{1}^{-1},...,w_{\ell}^{-1})\ \ \ (\text{By Lemma }\ref{HESlambda}.)
=\displaystyle= (r11)e1(w11,,w1)++(r1r1)er1(w11,,w1)(By Lemma 3.3.)\displaystyle\binom{r-1}{1}e_{1}(w_{1}^{-1},...,w_{\ell}^{-1})+\cdots+\binom{r-1}{r-1}e_{r-1}(w_{1}^{-1},...,w_{\ell}^{-1})\ \ \ (\text{By Lemma }\ref{EtoHorHtoE}.)
=\displaystyle= j=1r1(r1j)ej(w11,,w1).\displaystyle\sum_{j=1}^{r-1}\binom{r-1}{j}e_{j}(w_{1}^{-1},...,w_{\ell}^{-1}).

Therefore, Equation (9) is correct. Furthermore, combining Equations (5), (3.2), and (9), we have

Fr(x)=xCr+1=i=0r1(1)i(r+ii)er+1+i(w1,w2,,w).F_{r}(x)=xC_{\ell-r}+1=\sum_{i=0}^{r-1}(-1)^{i}\binom{\ell-r+i}{i}e_{\ell-r+1+i}(w_{1},w_{2},...,w_{\ell}).

This completes the proof. ∎

The following results were first obtained by de Mier and Noy.

Corollary 3.11 ([10]).

Following the notation in Theorem 1.2. Let q(n)q(n) be the number of lattice paths from (0,0)(0,0) to (n,kn)(\ell n,kn) that never go above the path (NkE)n(N^{k}E^{\ell})^{n}. Then the generating function Q(x)=n0q(n)xnQ(x)=\sum_{n\geq 0}q(n)x^{n} is given by

Q(x)=1x(1w1)(1w).Q(x)=\frac{-1}{x}(1-w_{1})\cdots(1-w_{\ell}).
Proof.

By the lattice path interpretation of C0C_{0}, we have Q(x)=C0Q(x)=C_{0}. By Theorem 1.2, we have

Q(x)\displaystyle Q(x) =1x(F(x)1)=1x(i=01(1)iei+11)\displaystyle=\frac{1}{x}(F_{\ell}(x)-1)=\frac{1}{x}\left(\sum_{i=0}^{\ell-1}(-1)^{i}e_{i+1}-1\right)
=1x(i=0(1)i1ei)=1x(1w1)(1w2)(1w).\displaystyle=\frac{1}{x}\left(\sum_{i=0}^{\ell}(-1)^{i-1}e_{i}\right)=\frac{-1}{x}(1-w_{1})(1-w_{2})\cdots(1-w_{\ell}).

This completes the proof. ∎

Corollary 3.12.

Following the notation in Theorem 1.2. If r=1r=1, then we have

F1(x)=w1w2w.F_{1}(x)=w_{1}w_{2}\cdots w_{\ell}.

If r=2r=2, then we have

F2(x)=(1+j=11wj)j=1wj.F_{2}(x)=\left(1-\ell+\sum_{j=1}^{\ell}\frac{1}{w_{j}}\right)\cdot\prod_{j=1}^{\ell}w_{j}.

In particular, if r=1r=1 and k=k=\ell, then

F1(x)=j=1C(ξjx1).F_{1}(x)=\prod_{j=1}^{\ell}C(\xi^{j}x^{\frac{1}{\ell}}).

If r=2r=2 and k=k=\ell, then

F2(x)=(1+j=11C(ξjx1))j=1C(ξjx1),F_{2}(x)=\left(1-\ell+\sum_{j=1}^{\ell}\frac{1}{C(\xi^{j}x^{\frac{1}{\ell}})}\right)\cdot\prod_{j=1}^{\ell}C(\xi^{j}x^{\frac{1}{\ell}}),

where C(x)C(x) is the Catalan function and ξ\xi is a primitive \ell-th root of unity ξ=e2πi\xi=e^{\frac{2\pi i}{\ell}}.

4. Exponential Formulas for Q(x)Q(x) and F1(x)F_{1}(x)

In this section, we mainly give two exponential formulas in Theorems 4.1 and 4.3, using ideas from the doctoral thesis of the second author, especially [15, Chapter 1, 1-5]. The thesis also includes some basic knowledge of Puiseux’s theorem.

Throughout this section, we always use the following factorization

(w1)xwk+=x(ww1)(ww)(ww+1)(wwk+).\displaystyle(w-1)^{\ell}-xw^{k+\ell}=-x(w-w_{1})\cdots(w-w_{\ell})(w-w_{\ell+1})\cdots(w-w_{k+\ell}). (10)

By Puiseux’s theorem, we may assume wiw_{i} are all fractional Laurent series in fra((x))\mathbb{C}^{\text{fra}}((x)) and w1,,ww_{1},...,w_{\ell} is the unique fractional power series. Moreover, wi|x=0=1w_{i}|_{x=0}=1 for 1i1\leq i\leq\ell.

Now we state our first result.

Theorem 4.1.

Follow the notation in Corollary 3.11. Let q(n)q(n) be the number of lattice paths from (0,0)(0,0) to (n,kn)(\ell n,kn) that never go above the path (NkE)n(N^{k}E^{\ell})^{n}. Then we have

Q(x)=n0q(n)xn=exp(n1(kn+nn)xnn).Q(x)=\sum_{n\geq 0}q(n)x^{n}=\exp\left(\sum_{n\geq 1}\binom{kn+\ell n}{\ell n}\frac{x^{n}}{n}\right).
Proof.

Put w=u+1w=u+1 in (10). We obtain the factorization

ux(u+1)k+=x(uu1)(uu2)(uuk+),\displaystyle u^{\ell}-x(u+1)^{k+\ell}=-x(u-u_{1})(u-u_{2})\cdots(u-u_{k+\ell}), (11)

where ui=wi1, 1ik+u_{i}=w_{i}-1,\ 1\leq i\leq k+\ell are fractional Laurent series, and u1,,uu_{1},...,u_{\ell} are the only fractional power series. Moreover, ui|x=0=0u_{i}|_{x=0}=0 for 1i1\leq i\leq\ell.

Rewrite (11) as

1x(u+1)k+u=A(x)(1u1u)(1uu)(1uu+1)(1uuk+).\displaystyle 1-\frac{x(u+1)^{k+\ell}}{u^{\ell}}=A(x)\Big{(}1-\frac{u_{1}}{u}\Big{)}\cdots\Big{(}1-\frac{u_{\ell}}{u}\Big{)}\Big{(}1-\frac{u}{u_{\ell+1}}\Big{)}\cdots\Big{(}1-\frac{u}{u_{k+\ell}}\Big{)}. (12)

Then by comparing the lowest power terms of uu on both sides of the above equation, we obtain

xu=A(x)(u1u)(uu).-\frac{x}{u^{\ell}}=A(x)\cdot(-\frac{u_{1}}{u})\cdots(-\frac{u_{\ell}}{u}).

Therefore by Corollary 3.11, we have

(A(x))1=(1)+1xu1u=(1)+1x(w11)(w1)=Q(x).(A(x))^{-1}=\frac{(-1)^{\ell+1}}{x}u_{1}\cdots u_{\ell}=\frac{(-1)^{\ell+1}}{x}(w_{1}-1)\cdots(w_{\ell}-1)=Q(x).

Now we extract (A(x))1(A(x))^{-1} through taking logarithms and working in ((u))fra((x))\mathbb{C}((u))^{\text{fra}}((x)), i.e., the field of fractional Laurent series in xx with coefficients Laurent series in uu. We have

ln(1x(u+1)k+u)1\displaystyle\ln\left(1-\frac{x(u+1)^{k+\ell}}{u^{\ell}}\right)^{-1} =ln(A(x))1+i=1ln(1uiu)1+j=+1+kln(1uui)1\displaystyle=\ln(A(x))^{-1}+\sum_{i=1}^{\ell}\ln\left(1-\frac{u_{i}}{u}\right)^{-1}+\sum_{j=\ell+1}^{\ell+k}\ln\left(1-\frac{u}{u_{i}}\right)^{-1}
=ln(A(x))1+i=1n1uinnun+j=+1+kn1unnujn\displaystyle=\ln(A(x))^{-1}+\sum_{i=1}^{\ell}\sum_{n\geq 1}\frac{u_{i}^{n}}{nu^{n}}+\sum_{j=\ell+1}^{\ell+k}\sum_{n\geq 1}\frac{u^{n}}{nu_{j}^{n}}
=ln(A(x))1+n11nuni=1uin+n1unnj=+1+k1ujn.\displaystyle=\ln(A(x))^{-1}+\sum_{n\geq 1}\frac{1}{nu^{n}}\sum_{i=1}^{\ell}u_{i}^{n}+\sum_{n\geq 1}\frac{u^{n}}{n}\sum_{j=\ell+1}^{\ell+k}\frac{1}{u_{j}^{n}}. (13)

On the right hand side of Equation (13), the second term only contains negative power terms of uu, and the third term only contains positive power terms of uu. Thus the first term ln(A(x))1\ln(A(x))^{-1} is exactly the constant term, denoted CTu\mathop{\mathrm{CT}}_{u}, and we have

ln(A(x))1\displaystyle\ln(A(x))^{-1} =CTuln(1x(u+1)k+u)1=CTun11n((1+u)k+xu)n\displaystyle=\mathop{\mathrm{CT}}_{u}\ln\left(1-\frac{x(u+1)^{k+\ell}}{u^{\ell}}\right)^{-1}=\mathop{\mathrm{CT}}_{u}\sum_{n\geq 1}\frac{1}{n}\left(\frac{(1+u)^{k+\ell}\cdot x}{u^{\ell}}\right)^{n}
=CTun1(1+u)kn+nnunxn=n1(kn+nn)xnn.\displaystyle=\mathop{\mathrm{CT}}_{u}\sum_{n\geq 1}\frac{(1+u)^{kn+\ell n}}{nu^{\ell n}}\cdot x^{n}=\sum_{n\geq 1}\binom{kn+\ell n}{\ell n}\frac{x^{n}}{n}.

Therefore,

Q(x)=(A(x))1=exp(n1(kn+nn)xnn).Q(x)=(A(x))^{-1}=\exp\left(\sum_{n\geq 1}\binom{kn+\ell n}{\ell n}\frac{x^{n}}{n}\right).

This completes the proof. ∎

By observing the second term on the right hand side of Equation (13), we found that

m11mumi=1uim=m1pm(u1,,u)mum,\sum_{m\geq 1}\frac{1}{mu^{m}}\sum_{i=1}^{\ell}u_{i}^{m}=\sum_{m\geq 1}\frac{p_{m}(u_{1},...,u_{\ell})}{m}u^{-m},

where pmp_{m} is the power sum symmetric function. Thus we can also extract an interesting formula for pm(u1,,u)p_{m}(u_{1},...,u_{\ell}) as follows. m1pm(u1,,u)m^{-1}p_{m}(u_{1},...,u_{\ell}) is the coefficient of umu^{-m} in Equation (13). Therefore we have

m1pm(u1,,u)=[um]ln(1x(u+1)k+u)1=[um]n1(1+u)kn+nxnnun=n1(kn+nnm)xnn,\displaystyle m^{-1}p_{m}(u_{1},...,u_{\ell})=[u^{-m}]\ln\left(1-\frac{x(u+1)^{k+\ell}}{u^{\ell}}\right)^{-1}=[u^{-m}]\sum_{n\geq 1}\frac{(1+u)^{kn+\ell n}x^{n}}{nu^{\ell n}}=\sum_{n\geq 1}\binom{kn+\ell n}{\ell n-m}\frac{x^{n}}{n},

where (kn+nnm)=0\binom{kn+\ell n}{\ell n-m}=0 for n<m\ell n<m. We summarize the above results as the following proposition.

Proposition 4.2.

Let m1m\geq 1. Let w1,,ww_{1},...,w_{\ell} be the unique solutions of the equation (w1)xwk+=0(w-1)^{\ell}-xw^{k+\ell}=0 that are fractional power series. We have

pm(w11,,w1)=mn1(kn+nnm)xnn,p_{m}(w_{1}-1,...,w_{\ell}-1)=m\cdot\sum_{n\geq 1}\binom{kn+\ell n}{\ell n-m}\frac{x^{n}}{n},

where (kn+nnm)=0\binom{kn+\ell n}{\ell n-m}=0 for n<m\ell n<m.

Theorem 4.3.

Follow the notation in Theorem 1.2. Let f1(n)f_{1}(n) be the number of lattice paths from (0,0)(0,0) to (n1,kn)(\ell n-1,kn) that never go above the path (NkE)n1NkE1(N^{k}E^{\ell})^{n-1}N^{k}E^{\ell-1}. Then we have

F1(x)=n0f1(n)xn=exp(n1(kn+n1n1)xnn).F_{1}(x)=\sum_{n\geq 0}f_{1}(n)x^{n}=\exp\left(\sum_{n\geq 1}\binom{kn+\ell n-1}{\ell n-1}\frac{x^{n}}{n}\right).
Proof.

Rewrite Equation (10) as

(w1)xwk+=B(x)i=1(wwi)j=+1+k(1wwj),(w-1)^{\ell}-xw^{k+\ell}=B(x)\cdot\prod_{i=1}^{\ell}(w-w_{i})\prod_{j=\ell+1}^{\ell+k}\left(1-\frac{w}{w_{j}}\right),

where

B(x)=(1)k+1xw+1w+k.B(x)=(-1)^{k+1}x\cdot w_{\ell+1}\cdots w_{\ell+k}.

Thus we obtain

1xwk+(w1)=B(x)i=1wwiw1j=+1+k(1wwj).1-\frac{xw^{k+\ell}}{(w-1)^{\ell}}=B(x)\cdot\prod_{i=1}^{\ell}\frac{w-w_{i}}{w-1}\prod_{j=\ell+1}^{\ell+k}\left(1-\frac{w}{w_{j}}\right).

Now making the substitution ww1w\to w^{-1} and simplifying gives

1xwk(1w)=B(x)i=11wwi1wj=+1+k(11wwj).\displaystyle 1-\frac{xw^{-k}}{(1-w)^{\ell}}=B(x)\cdot\prod_{i=1}^{\ell}\frac{1-ww_{i}}{1-w}\prod_{j=\ell+1}^{\ell+k}\left(1-\frac{1}{ww_{j}}\right). (14)

By Equation (14), we have

wk(1w)x=B(x)i=1(1wiw)j=+1+k(w1wj).w^{k}(1-w)^{\ell}-x=B(x)\cdot\prod_{i=1}^{\ell}(1-w_{i}w)\prod_{j=\ell+1}^{\ell+k}\left(w-\frac{1}{w_{j}}\right).

Then by comparing the coefficients of wk+w^{k+\ell} on both sides of the above equation, we obtain

(1)wk+=B(x)(1)w+ki=1wi.(-1)^{\ell}w^{k+\ell}=B(x)\cdot(-1)^{\ell}w^{\ell+k}\prod_{i=1}^{\ell}w_{i}.

Therefore we have

(B(x))1=w1w2w=F1(x),(B(x))^{-1}=w_{1}w_{2}\cdots w_{\ell}=F_{1}(x),

where the second equality follows by Corollary 3.12.

Again, by taking logarithms in (14), we obtain

ln(1xwk(1w))1\displaystyle\ln\left(1-\frac{xw^{-k}}{(1-w)^{\ell}}\right)^{-1} =ln(B(x))1+i=1ln(1wwi1w)1+j=+1+kln(11wwj)1\displaystyle=\ln(B(x))^{-1}+\sum_{i=1}^{\ell}\ln\left(\frac{1-ww_{i}}{1-w}\right)^{-1}+\sum_{j=\ell+1}^{\ell+k}\ln\left(1-\frac{1}{ww_{j}}\right)^{-1}
=ln(B(x))1+i=1ln(1w(wi1)1w)1+j=+1+kln(11wwj)1\displaystyle=\ln(B(x))^{-1}+\sum_{i=1}^{\ell}\ln\left(1-\frac{w(w_{i}-1)}{1-w}\right)^{-1}+\sum_{j=\ell+1}^{\ell+k}\ln\left(1-\frac{1}{ww_{j}}\right)^{-1}
=ln(B(x))1+i=1n11n((wi1)w1w)n+j=+1+kn11n(wwj)n.\displaystyle=\ln(B(x))^{-1}+\sum_{i=1}^{\ell}\sum_{n\geq 1}\frac{1}{n}\left(\frac{(w_{i}-1)w}{1-w}\right)^{n}+\sum_{j=\ell+1}^{\ell+k}\sum_{n\geq 1}\frac{1}{n(ww_{j})^{n}}. (15)

On the right hand side of Equation (15), the second term only contains positive power terms of ww, and the third term only contains negative power terms of ww. Thus the first term is just the constant term

ln(B(x))1=CTwln(1xwk(1w))1=n1xnnCTw1wkn(1w)n=n1(kn+n1kn)xnn.\displaystyle\ln(B(x))^{-1}=\mathop{\mathrm{CT}}\limits_{w}\ln\left(1-\frac{xw^{-k}}{(1-w)^{\ell}}\right)^{-1}=\sum_{n\geq 1}\frac{x^{n}}{n}\mathop{\mathrm{CT}}\limits_{w}\frac{1}{w^{kn}(1-w)^{\ell n}}=\sum_{n\geq 1}\binom{kn+\ell n-1}{kn}\frac{x^{n}}{n}.

Thus we have

F1(x)=(B(x))1=exp(n1(kn+n1n1)xnn).F_{1}(x)=(B(x))^{-1}=\exp\left(\sum_{n\geq 1}\binom{kn+\ell n-1}{\ell n-1}\frac{x^{n}}{n}\right).

This completes the proof. ∎

Proof of Theorem 1.1.

The theorem is followed by Proposition 2.3, Theorem 1.2 and Theorem 1.4. ∎

Similar to Proposition 4.2, we can obtain the following result.

Proposition 4.4.

Let m1m\geq 1. Let w1,,ww_{1},...,w_{\ell} be the unique solutions of the equation (w1)xwk+=0(w-1)^{\ell}-xw^{k+\ell}=0 that are fractional power series. Let w+1,,w+kw_{\ell+1},...,w_{\ell+k} be the other kk solutions in the field of fractional Laurent series fra((x))\mathbb{C}^{\text{fra}}((x)). We have

pm(w+11,,w+k1)=mn1(n+knm1knm)xnn,p_{m}(w_{\ell+1}^{-1},...,w_{\ell+k}^{-1})=m\cdot\sum_{n\geq 1}\binom{\ell n+kn-m-1}{kn-m}\frac{x^{n}}{n},

where (kn+nm1knm)=0\binom{kn+\ell n-m-1}{kn-m}=0 for kn<mkn<m or kn+n<m+1kn+\ell n<m+1.

5. Two Byproducts

Our proof of Theorem 1.2 seems lengthy, so we tried to solve the system (4) for CtC_{t}, 0t0\leq t\leq\ell in other ways. This leads to two byproducts, namely, Corollary 5.3 and Theorem 5.4. Simple proofs of the byproducts may give rise simple proofs of Theorem 1.2.

5.1. Recursive Operation

In [10], de Mier and Noy obtained the full Tutte polynomials as follows:

n0An(z,y)xn=(zw1)(zw)(xzk+(z1))(y+w1yw1)(y+wyw).\sum_{n\geq 0}A_{n}(z,y)x^{n}=\frac{-(z-w_{1})\cdots(z-w_{\ell})}{(xz^{k+\ell}-(z-1)^{\ell})(y+w_{1}-yw_{1})\cdots(y+w_{\ell}-yw_{\ell})}.

We only need the specialization at y=1y=1. Let

A¯(z):=n0An(z,1)xn=(zw1)(zw)xzk+(z1).\displaystyle\mathrm{\overline{A}}(z):=\sum_{n\geq 0}A_{n}(z,1)x^{n}=\frac{-(z-w_{1})\cdots(z-w_{\ell})}{xz^{k+\ell}-(z-1)^{\ell}}. (16)
Theorem 5.1.

Following the notation in Theorem 1.2. Let 1r11\leq r\leq\ell-1. Then we have

Fr(x)=1+i=1rS(r,i)xB¯1(i1)(1),F_{r}(x)=1+\sum_{i=1}^{\ell-r}S(\ell-r,i)x\mathrm{\overline{B}}_{1}^{(i-1)}(1),

where

B¯1(z)=(k+1)zkA¯(z)+zk+1A¯(z),A¯(z)=(zw1)(zw)xzk+(z1),\mathrm{\overline{B}}_{1}(z)=(k+1)z^{k}\mathrm{\overline{A}}(z)+z^{k+1}\mathrm{\overline{A}}^{\prime}(z),\ \ \ \mathrm{\overline{A}}(z)=\frac{-(z-w_{1})\cdots(z-w_{\ell})}{xz^{k+\ell}-(z-1)^{\ell}},

the S(r,i)S(\ell-r,i) are Stirling numbers of the second kind and B¯1(i1)(z)\mathrm{\overline{B}}^{(i-1)}_{1}(z) is a (i1)(i-1)-th degree differential operation for B¯1(z)\mathrm{\overline{B}}_{1}(z).

Proof.

Follow the notation in Subsection 3.2. Let

B¯i(z)=n0B1,nxn, 1i.\mathrm{\overline{B}}_{i}(z)=\sum_{n\geq 0}B_{1,n}x^{n},\ \ \ 1\leq i\leq\ell.

We simultaneously calculate the generating functions for both sides of Equation (2) at y=1y=1. This gives the following equations.

B¯1(z)=zk+1A¯(z)C0z1,B¯2(z)=zB¯1(z)C1z1,,B¯(z)=zB¯1(z)C1z1.\displaystyle\mathrm{\overline{B}}_{1}(z)=\frac{z^{k+1}\mathrm{\overline{A}}(z)-C_{0}}{z-1},\ \ \mathrm{\overline{B}}_{2}(z)=\frac{z\mathrm{\overline{B}}_{1}(z)-C_{1}}{z-1},\ \ \cdots,\ \ \mathrm{\overline{B}}_{\ell}(z)=\frac{z\mathrm{\overline{B}}_{\ell-1}(z)-C_{\ell-1}}{z-1}.

By L’Hospital’s rule, we have

B¯1(z)=(k+1)zkA¯(z)+zk+1A¯(z),B¯i(z)=B¯i1(z)+zB¯i1(z), 2i.\mathrm{\overline{B}}_{1}(z)=(k+1)z^{k}\mathrm{\overline{A}}(z)+z^{k+1}\mathrm{\overline{A}}^{\prime}(z),\ \ \ \mathrm{\overline{B}}_{i}(z)=\mathrm{\overline{B}}_{i-1}(z)+z\mathrm{\overline{B}}^{\prime}_{i-1}(z),\ \ 2\leq i\leq\ell.

By the recursion for the Stirling numbers of the second kind S(t,i)S(t,i) (see [11, A008277] or [12, Chapter 1]):

S(t,i)=iS(t1,i)+S(t1,i1),S(1,1)=1,S(t,i)=0,(i>t),S(t,i)=i\cdot S(t-1,i)+S(t-1,i-1),\ \ S(1,1)=1,\ \ S(t,i)=0,(i>t),

we have

B¯t(z)=i=1tS(t,i)zi1B¯1(i1)(z), 1t,\mathrm{\overline{B}}_{t}(z)=\sum_{i=1}^{t}S(t,i)z^{i-1}\mathrm{\overline{B}}^{(i-1)}_{1}(z),\ \ 1\leq t\leq\ell,

where B¯1(i1)(z)\mathrm{\overline{B}}^{(i-1)}_{1}(z) is a (i1)(i-1)-th derivative of B¯1(z)\mathrm{\overline{B}}_{1}(z) with respect to zz. Furthermore, by Ct=n0Ct,nxn=B¯t(1)C_{t}=\sum_{n\geq 0}C_{t,n}x^{n}=\mathrm{\overline{B}}_{t}(1), 1t11\leq t\leq\ell-1, we have

Ct=i=1tS(t,i)B¯1(i1)(1), 1t1.C_{t}=\sum_{i=1}^{t}S(t,i)\mathrm{\overline{B}}_{1}^{(i-1)}(1),\ \ 1\leq t\leq\ell-1.

Note that C0=A¯(1)C_{0}=\mathrm{\overline{A}}(1).

The proof is then completed by Fr(x)=xCr+1F_{r}(x)=xC_{\ell-r}+1 for 1r1\leq r\leq\ell. ∎

Corollary 5.2.

Follow the notation in Theorem 5.1. We have

F1(x)=1+(1i=111wi)i=1(1wi).F_{\ell-1}(x)=1+\left(\ell-1-\sum_{i=1}^{\ell}\frac{1}{1-w_{i}}\right)\prod_{i=1}^{\ell}(1-w_{i}).
Proof.

By Theorem 5.1, we have

A¯(1)=1xj=1(1wj),A¯(1)=1xj=1(1wj)(i=111wik).\mathrm{\overline{A}}(1)=\frac{-1}{x}\prod_{j=1}^{\ell}(1-w_{j}),\ \ \ \ \ \mathrm{\overline{A}}^{\prime}(1)=-\frac{1}{x}\prod_{j=1}^{\ell}(1-w_{j})\left(\sum_{i=1}^{\ell}\frac{1}{1-w_{i}}-\ell-k\right).

Moreover,

B¯1(1)=(k+1)A¯(1)+A¯(1)=1x(1i=111wi)j=1(1wj).\mathrm{\overline{B}}_{1}(1)=(k+1)\mathrm{\overline{A}}(1)+\mathrm{\overline{A}}^{\prime}(1)=\frac{1}{x}\left(\ell-1-\sum_{i=1}^{\ell}\frac{1}{1-w_{i}}\right)\prod_{j=1}^{\ell}(1-w_{j}).

Furthermore, by Theorem 5.1, we have F1=1+xB¯1(1)F_{\ell-1}=1+x\mathrm{\overline{B}}_{1}(1). This completes the proof. ∎

Theorem 1.2 implies that

F1(x)\displaystyle F_{\ell-1}(x) =i=02(1)i(1+ii)ei+2(w1,,w)\displaystyle=\sum_{i=0}^{\ell-2}(-1)^{i}\binom{1+i}{i}e_{i+2}(w_{1},...,w_{\ell})
=e2(w1,,w)2e3(w1,,w)++(1)2(1)e(w1,,w).\displaystyle=e_{2}(w_{1},...,w_{\ell})-2e_{3}(w_{1},...,w_{\ell})+\cdots+(-1)^{\ell-2}(\ell-1)e_{\ell}(w_{1},...,w_{\ell}).

This result can also be explained by Corollary 5.2 as follows.

F1(x)\displaystyle F_{\ell-1}(x) =1+(1w1)(1w)(1i=111wi)\displaystyle=1+(1-w_{1})\cdots(1-w_{\ell})\left(\ell-1-\sum_{i=1}^{\ell}\frac{1}{1-w_{i}}\right)
=1+(1)(e0e1+e2++(1)e)(j=1(1wj)1w1++j=1(1wj)1w)\displaystyle=1+(\ell-1)(e_{0}-e_{1}+e_{2}+\cdots+(-1)^{\ell}e_{\ell})-\left(\frac{\prod_{j=1}^{\ell}(1-w_{j})}{1-w_{1}}+\cdots+\frac{\prod_{j=1}^{\ell}(1-w_{j})}{1-w_{\ell}}\right)
=1+(1)(e0e1++(1)e)(e0(l1)e1++(1)1e1+0e)\displaystyle=1+(\ell-1)(e_{0}-e_{1}+\cdots+(-1)^{\ell}e_{\ell})-(\ell e_{0}-(l-1)e_{1}+\cdots+(-1)^{\ell-1}e_{\ell-1}+0\cdot e_{\ell})
=e22e3++(1)2(1)e,\displaystyle=e_{2}-2e_{3}+\cdots+(-1)^{\ell-2}(\ell-1)e_{\ell},

where ei=ei(w1,,w)e_{i}=e_{i}(w_{1},...,w_{\ell}) for 0i0\leq i\leq\ell.

Corollary 5.3.

Follow the notation in Theorem 5.1. Then we have

i=1rS(r,i)x((k+1)zkA¯j(z)+zk+1A¯j(1)(z))(i1)|z=1\displaystyle\sum_{i=1}^{\ell-r}S(\ell-r,i)x\Big{(}(k+1)z^{k}\mathrm{\overline{A}}_{j}(z)+z^{k+1}\mathrm{\overline{A}}_{j}^{(1)}(z)\Big{)}^{(i-1)}\Big{|}_{z=1}
=\displaystyle= {1ifj=0;0if 1jr;(1)j+r1(j1j+r1)ifr+1j,\displaystyle\left\{\begin{array}[]{ll}-1&\ \text{if}\ \ j=0;\\ 0&\ \text{if}\ \ 1\leq j\leq\ell-r;\\ (-1)^{j+r-1-\ell}\binom{j-1}{j+r-1-\ell}&\ \text{if}\ \ \ell-r+1\leq j\leq\ell,\end{array}\right.

where A¯j(z)=(1)j+1zjxzk+(z1)\mathrm{\overline{A}}_{j}(z)=\frac{(-1)^{j+1}z^{\ell-j}}{xz^{k+\ell}-(z-1)^{\ell}} and the S(r,i)S(\ell-r,i) are Stirling numbers of the second kind.

Proof.

By Equation (16), we have

A¯(z)=j=0A¯j(z)ej(w1,,w).\mathrm{\overline{A}}(z)=\sum_{j=0}^{\ell}\mathrm{\overline{A}}_{j}(z)e_{j}(w_{1},...,w_{\ell}).

Therefore

B¯1(z)=(k+1)zkA¯(z)+zk+1A¯(z)=j=0((k+1)zkA¯j(z)+zk+1A¯j(z))ej(w1,,w).\mathrm{\overline{B}}_{1}(z)=(k+1)z^{k}\mathrm{\overline{A}}(z)+z^{k+1}\mathrm{\overline{A}}^{\prime}(z)=\sum_{j=0}^{\ell}((k+1)z^{k}\mathrm{\overline{A}}_{j}(z)+z^{k+1}\mathrm{\overline{A}}_{j}^{\prime}(z))e_{j}(w_{1},...,w_{\ell}).

By Theorem 5.1, we have

Fr(x)=1+j=0(i=1rS(r,i)x((k+1)zkA¯j(z)+zk+1A¯j(1)(z))(i1)|z=1)ej(w1,,w).F_{r}(x)=1+\sum_{j=0}^{\ell}\left(\sum_{i=1}^{\ell-r}S(\ell-r,i)x\Big{(}(k+1)z^{k}\mathrm{\overline{A}}_{j}(z)+z^{k+1}\mathrm{\overline{A}}_{j}^{(1)}(z)\Big{)}^{(i-1)}\Big{|}_{z=1}\right)e_{j}(w_{1},...,w_{\ell}).

The corollary is followed by Theorem 1.2. ∎

5.2. A Symmetric Function Formula

It is natural to solve the system (4) for CrC_{\ell-r} using Kramer’s rule. Rewrite (4) as

i=1(wjwj1)i1xCi=(wj1), 1j.\displaystyle\sum_{i=1}^{\ell}\left(\frac{w_{j}}{w_{j}-1}\right)^{i-1}xC_{\ell-i}=(w_{j}-1),\ \ 1\leq j\leq\ell.

Let w^i=wiwi1\widehat{w}_{i}=\frac{w_{i}}{w_{i}-1}, i.e., wi=w^iw^i1w_{i}=\frac{\widehat{w}_{i}}{\widehat{w}_{i}-1} for 1i1\leq i\leq\ell. By Kramer’s rule, we have xCr=DrDxC_{\ell-r}=\frac{D_{r}}{D}, where

D=det(1w^1w^111w^2w^211w^w^1)=detV(w^1,w^2,,w^).\displaystyle D=\det\left(\begin{array}[]{cccc}1&\widehat{w}_{1}&\cdots&\widehat{w}_{1}^{\ell-1}\\ 1&\widehat{w}_{2}&\cdots&\widehat{w}_{2}^{\ell-1}\\ \vdots&\vdots&\ddots&\vdots\\ 1&\widehat{w}_{\ell}&\cdots&\widehat{w}_{\ell}^{\ell-1}\end{array}\right)=\det V(\widehat{w}_{1},\widehat{w}_{2},...,\widehat{w}_{\ell}).

and

Dr=det(1w^1w^1r2w11w^1rw^111w^2w^2r2w21w^2rw^211w^w^r2w1w^rw^1).\displaystyle D_{r}=\det\left(\begin{array}[]{cccccccc}1&\widehat{w}_{1}&\cdots&\widehat{w}_{1}^{r-2}&w_{1}-1&\widehat{w}_{1}^{r}&\cdots&\widehat{w}_{1}^{\ell-1}\\ 1&\widehat{w}_{2}&\cdots&\widehat{w}_{2}^{r-2}&w_{2}-1&\widehat{w}_{2}^{r}&\cdots&\widehat{w}_{2}^{\ell-1}\\ \vdots&\vdots&\ddots&\vdots&\vdots&\vdots&\ddots&\vdots\\ 1&\widehat{w}_{\ell}&\cdots&\widehat{w}_{\ell}^{r-2}&w_{\ell}-1&\widehat{w}_{\ell}^{r}&\cdots&\widehat{w}_{\ell}^{\ell-1}\end{array}\right).

Now rewrite

wi1=w^i1w^i1=m0(w^i)m, 1i,w_{i}-1=-\frac{\widehat{w}_{i}}{1-\widehat{w}_{i}}-1=-\sum_{m\geq 0}(\widehat{w}_{i})^{m},\ \ \ 1\leq i\leq\ell,

and put in the above equation for DrD_{r} and expand by linearity. When 0m0\leq m\leq\ell, the corresponding term is 0, except that the term corresponding to m=r1m=r-1 is detV(w^1,w^2,,w^)-\det V(\widehat{w}_{1},\widehat{w}_{2},...,\widehat{w}_{\ell}). The remaining terms correspond to mm\geq\ell. By Lemma 3.1 and some simple calculations, we have

Dr=detV(w^1,,w^)+(1)r+1detV(w^1,,w^)mls(m+1,1r)(w^1,,w^),D_{r}=-\det V(\widehat{w}_{1},...,\widehat{w}_{\ell})+(-1)^{\ell-r+1}\det V(\widehat{w}_{1},...,\widehat{w}_{\ell})\sum_{m\geq l}s_{(m-\ell+1,1^{\ell-r})}(\widehat{w}_{1},...,\widehat{w}_{\ell}),

where (m+1,1r)=(m+1,1,1,,1)(m-\ell+1,1^{\ell-r})=(m-\ell+1,1,1,...,1) with length r+1\ell-r+1. Therefore, we have

Fr(x)=1+xCr=1+DrD=(1)r+1m1s(m,1r)(w^1,,w^).F_{r}(x)=1+xC_{\ell-r}=1+\frac{D_{r}}{D}=(-1)^{\ell-r+1}\sum_{m\geq 1}s_{(m,1^{\ell-r})}(\widehat{w}_{1},...,\widehat{w}_{\ell}). (17)

By Theorem 1.2, we complete the proof of Theorem 5.4 as follows.

Theorem 5.4.

Let ,r\ell,r\in\mathbb{P} and r\ell\geq r. Let w^i=wiwi1\widehat{w}_{i}=\frac{w_{i}}{w_{i}-1}, 1i1\leq i\leq\ell. We have

m1s(m,1r)(w^1,,w^)=i=0r1(1)r+i+1(r+ii)er+i+1(w1,,w).\sum_{m\geq 1}s_{(m,1^{\ell-r})}(\widehat{w}_{1},...,\widehat{w}_{\ell})=\sum_{i=0}^{r-1}(-1)^{\ell-r+i+1}\binom{\ell-r+i}{i}e_{\ell-r+i+1}(w_{1},...,w_{\ell}).
Example 5.5.

Suppose =r\ell=r. Theorem 1.2 implies that

F(x)=i=01(1)iei+1(w1,,w).F_{\ell}(x)=\sum_{i=0}^{\ell-1}(-1)^{i}e_{i+1}(w_{1},...,w_{\ell}). (18)

This formula can be explained by Equation (17) as follows.

F(x)=m1sm(w^1,,w^).F_{\ell}(x)=-\sum_{m\geq 1}s_{m}(\widehat{w}_{1},...,\widehat{w}_{\ell}).

Because sm(x1,x2,)=hm(x1,x2,)s_{m}(x_{1},x_{2},...)=h_{m}(x_{1},x_{2},...) and m0hm(x1,x2,)tm=1i(1xit)\sum_{m\geq 0}h_{m}(x_{1},x_{2},...)t^{m}=\frac{1}{\prod_{i}(1-x_{i}t)} (see, e.g., [13, Theorem 7.6.1]), we have

F(x)\displaystyle F_{\ell}(x) =m1hm(w^1,,w^)=1m0hm(w^1,,w^)\displaystyle=-\sum_{m\geq 1}h_{m}(\widehat{w}_{1},...,\widehat{w}_{\ell})=1-\sum_{m\geq 0}h_{m}(\widehat{w}_{1},...,\widehat{w}_{\ell})
=11(1w^1)(1w^2)(1w^)=1(1w1)(1w2)(1w)\displaystyle=1-\frac{1}{(1-\widehat{w}_{1})(1-\widehat{w}_{2})\cdots(1-\widehat{w}_{\ell})}=1-(1-w_{1})(1-w_{2})\cdots(1-w_{\ell})
=e1(w1,,w)e2(w1,,w)++(1)1e(w1,,w).\displaystyle=e_{1}(w_{1},...,w_{\ell})-e_{2}(w_{1},...,w_{\ell})+\cdots+(-1)^{\ell-1}e_{\ell}(w_{1},...,w_{\ell}).

6. Concluding Remark

Let SmS_{m} (2m62\leq m\leq 6) be the sequences [A079489], [A213403], [A213404], [A213405] and [A213406] in OEIS [11]. We have proved in Theorem 1.1 that SmS_{m} is in fact our f¯m(n)\overline{f}_{m}(n). Now we can enrich the content of these sequences. We summarize as follows:

  1. (1)

    The sequences SmS_{m} are the number of rectangular Young tableaux mn\mathcal{B}_{m}^{n} with periodic walls of a block m\mathcal{B}_{m}.

  2. (2)

    The sequences SmS_{m} are the number of lattice paths from (0,0)(0,0) to (mn,mn)(mn,mn) that never go below the path N(EmNm)n1EmNm1N(E^{m}N^{m})^{n-1}E^{m}N^{m-1}. (See Proposition 2.5.)

  3. (3)

    A recursive formula for sequences SmS_{m} is shown in Proposition 2.6.

  4. (4)

    A determinant formula for sequences SmS_{m} is given in Proposition 2.5.

  5. (5)

    The generating functions of sequences SmS_{m} are given in Corollary 1.1. These are also the explanation of these sequences in OEIS.

  6. (6)

    The generating functions of sequences SmS_{m} are given by the exponential formulas in Theorem 4.3 with k=l=mk=l=m.

In this paper, we only enumerated Young tableaux with periodic walls over m\mathcal{B}_{m}. The idea works for all Young building blocks with horizontal walls. Our next project is to study these generating functions.

Acknowledgements: We are grateful to Menghao Qu for many useful suggestions. This work is partially supported by the National Natural Science Foundation of China [12071311].

References

  • [1] C. Banderier, P. Marchal, and M. Wallner, Rectangular Young tableaux with local decreases and the density method for uniform random generation, GASCom 2018, CEUR Workshop Proceedings. Vol.2113 (2018), 60–68.
  • [2] C. Banderier, and M. Wallner, Young tableaux with periodic walls: counting with the density method, Proceedings of the 33rd Conference on FPSAC. 85B (2021), Article#47.
  • [3] J. Bonin, A. de Mier, and M. Noy, Lattice path matroids: enumerative aspects and Tutte polynomials, J. Combin. Theory Ser. A. 104 (2003), 63–94.
  • [4] R. J. Chapman, T. Chow, A. Khetan, D. P. Moulton, and R. J. Waters, Simple formulas for lattice paths avoiding certain periodic staircase boundaries, J. Combin. Theory Ser. A. 116 (1)(2009), 205–214.
  • [5] Ira M. Gessel, and M. Loehr, Note on enumeration of partitions contained in a given shape, Linear Algebra Appl. 432 (2010), 583–585.
  • [6] J. Haglund, The qq,tt-Catalan Number and the Space of Diagonal Harmonics: with an appendix on the combinatorics of Macdonald polynomials, AMS University Lecture Series, 2008.
  • [7] B. R. Handa, and S. G. Mohanty, On qq-binomial coefficients and some statistical applications, SIAM J. Math. Anal. 11 (1980), 1027–1035.
  • [8] J. Irving, and A. Rattan, The number of lattice paths below a cyclically shifting boundary, J. Combin. Theory Ser. A. 116 (2009), 499–514.
  • [9] G. Kreweras, Sur une classe de problémes de denombrement liés au treillis des partitions des entiers, Cahiers du BURO. 6 (1965), 5–105.
  • [10] A. de Mier, and M. Noy, A solution to the tennis ball problem, Theoret. Comput. Sci. 346 (2005), 254–264.
  • [11] N. J. A. Sloane, The on-line encyclopedia of integer sequences, published electronically at http://oeis.org.
  • [12] R. P. Stanley, Enumerative Combinatorics (volume 1), Cambridge Studies in Advanced Mathematics, vol. 49, Cambridge University Press, 2012.
  • [13] R. P. Stanley, Enumerative Combinatorics (volume 2), Cambridge Studies in Advanced Mathematics, vol. 62, Cambridge University Press, 1999.
  • [14] U. Tamm, Lattice paths not touching a given boundary, J. Statist. Plann. Inference. 105 (2002), 433–448.
  • [15] G. Xin, The ring of Malcev-Neumann series and the residue theorem, Ph.D. thesis, Brandeis University (2004).