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

On Promotion and Quasi-Tangled Labelings of Posets

Eliot Hodges Department of Mathematics, Harvard University, Cambridge, MA 02138 eliothodges@college.harvard.edu
Abstract.

In 2022, Defant and Kravitz introduced extended promotion (denoted \partial), a map that acts on the set of labelings of a poset. Extended promotion is a generalization of Schützenberger’s promotion operator, a well-studied map that permutes the set of linear extensions of a poset. It is known that if LL is a labeling of an nn-element poset PP, then n1(L)\partial^{n-1}(L) is a linear extension. This allows us to regard \partial as a sorting operator on the set of all labelings of PP, where we think of the linear extensions of PP as the labelings which have been sorted. The labelings requiring n1n-1 applications of \partial to be sorted are called tangled; the labelings requiring n2n-2 applications are called quasi-tangled. In addition to computing the sizes of the fibers of promotion for rooted tree posets, we count the quasi-tangled labelings of a relatively large class of posets called inflated rooted trees with deflated leaves. Given an nn-element poset with a unique minimal element with the property that the minimal element has exactly one parent, it follows from the aforementioned enumeration that this poset has 2(n1)!(n2)!2(n-1)!-(n-2)! quasi-tangled labelings. Using similar methods, we outline an algorithmic approach to enumerating the labelings requiring nk1n-k-1 applications to be sorted for any fixed k{1,,n2}k\in\{1,\ldots,n-2\}. We also make partial progress towards proving a conjecture of Defant and Kravitz on the maximum possible number of tangled labelings of an nn-element poset.

1. Introduction

1.1. Background

Let PP be an nn-element poset, whose order relation we denote by <P<_{P}. A labeling of PP is a bijection L:P[n]L:P\to[n] (where [n]={1,,n}[n]=\{1,\ldots,n\}). A labeling LL is called a linear extension if it preserves the order on PP, i.e. if for all pairs x,yPx,y\in P with x<Pyx<_{P}y we have L(x)<L(y)L(x)<L(y). Let Λ(P)\Lambda(P) be the set of all labelings of PP; let (P)Λ(P)\mathcal{L}(P)\subset\Lambda(P) be the subset consisting of all linear extensions.

In [Sch63, Sch72, Sch76], Schützenberger introduced an intriguing bijection on (P)\mathcal{L}(P) called promotion. Promotion has connections with various topics in algebraic combinatorics and representation theory, as seen in [EG87, Hua20, PPR09, Rho10, Sta09].

In [DK22], Defant and Kravitz extended the promotion map to an operator :Λ(P)Λ(P)\partial:\Lambda(P)\to\Lambda(P), not necessarily invertible, that is defined on all of Λ(P)\Lambda(P). When the poset is a chain, extended promotion is dynamically equivalent to the bubble-sort map studied in [Knu98]. Promotion can also be described in terms of Bender-Knuth involutions (first introduced by Haiman [Hai92] as well as Malvenuto and Reutenauer [MR94]); in [DK22], Defant and Kravitz extended these Bender-Knuth involutions to arrive at an equivalent “toggle” definition of extended promotion. The following results of [DK22] are crucial properties of extended promotion:

  1. (1)

    When restricted to (P)\mathcal{L}(P), \partial agrees with Schützenberger’s promotion operator.

  2. (2)

    If LL is a labeling of an nn-element poset PP, then n1(L)(P)\partial^{n-1}(L)\in\mathcal{L}(P).

Thus, (extended) promotion111Henceforth, “promotion” always refers to \partial rather than its restriction to (P)\mathcal{L}(P). may be regarded as a sorting operator, where linear extensions are considered “sorted.” Property (2) shows that promotion sorts every labeling after at most n1n-1 applications.

We define the sorting time of a labeling LL to be the smallest kk\in\mathbb{N} such that k(L)(P)\partial^{k}(L)\in\mathcal{L}(P). Defant and Kravitz mainly studied tangled labelings—those labelings with sorting time n1n-1. In particular, they enumerated these tangled labelings for a large class of posets called inflated rooted forests. They also studied sortable labelings—those labelings LL such that (L)(P)\partial(L)\in\mathcal{L}(P)—and enumerated these sortable labelings for arbitrary posets.

1.2. Outline and Summary of Main Results

In Section 2, we present the main definitions and background results needed for the rest of the paper. In Section 3, we study the cardinality of 1(L)\partial^{-1}(L) for an arbitrary labeling LL. Our 3.3 gives a formula for |1(L)||\partial^{-1}(L)| when PP is a rooted tree poset. We also study the degree of noninvertibility of promotion and give a sharp lower bound for this (3.7). Section 4 contains an explicit enumeration of the labelings with sorting time n2n-2 for a large class of posets called inflated rooted trees with deflated leaves (4.6). A corollary of this result is that an nn-element poset with a unique minimal element with the property that the minimal element has exactly one parent has 2(n1)!(n2)!2(n-1)!-(n-2)! quasi-tangled labelings. In Section 5, we present an algorithmic approach to enumerating the labelings of a rooted tree poset with sorting time nk1n-k-1 for fixed k{1,,n2}k\in\{1,\ldots,n-2\}, and in Section 6, we make partial progress (6.4) on the following conjecture:

Conjecture 1.1 ([DK22], Conjecture 5.1).

If PP is an nn-element poset, then PP has at most (n1)!(n-1)! tangled labelings.

Finally, in Section 7, we present several open problems and further directions of inquiry.

1.3. Extended Promotion

Let PP be an nn-element poset, and let LL be a labeling of PP. For xPx\in P not maximal, the L-successor of xx is the element greater than xx with minimal label. Now, let v1=L1(1)v_{1}=L^{-1}(1). Let v2v_{2} be the LL-successor of v1v_{1}; let v3v_{3} the LL-successor of v2v_{2}, and so on until we get an element vmv_{m} that is maximal. The resulting chain v1<Pv2<P<Pvmv_{1}<_{P}v_{2}<_{P}\cdots<_{P}v_{m} is called the promotion chain of LL. Now, define (L)\partial(L) to be the labeling

(L)(x)={L(x)1ifx{v1,,vm};L(vi+1)1ifx=vifori{1,,m1};nifx=vm.\partial(L)(x)=\begin{cases}L(x)-1&\mathrm{if\ }x\not\in\{v_{1},\ldots,v_{m}\};\\ L(v_{i+1})-1&\mathrm{if\ }x=v_{i}\ \mathrm{for\ }i\in\{1,\ldots,m-1\};\\ n&\mathrm{if\ }x=v_{m}.\end{cases}

In other words, promotion may be thought of as decreasing each label by 1 (working modulo nn so that 0=n0=n) and then cycling the promotion chain downwards one step. The following proposition captures a fundamental sorting property of promotion.

Proposition 1.2 ([DK22], Proposition 2.7).

If PP is an nn-element poset, then n1(Λ(P))=(P)\partial^{n-1}(\Lambda(P))=\mathcal{L}(P).

2. Preliminaries, Frozen Elements, and Some Special Classes of Posets

2.1. Preliminary Definitions

A lower order ideal of a poset PP is a subset QPQ\subset P such that for every xQx\in Q and yPy\in P with y<Pxy<_{P}x we have yQy\in Q. Similarly, an upper order ideal of a poset PP is a subset QPQ\subset P such that for all xQx\in Q and yPy\in P with y>Pxy>_{P}x we have yQy\in Q. It is often useful to note that QQ is a lower order ideal of PP if and only if PQP\setminus Q is an upper order ideal of PP. For x,yPx,y\in P, we say that yy covers xx and write xyx\lessdot y if x<Pyx<_{P}y and {zP|x<Pz<Py}=\{z\in P\;|\;x<_{P}z<_{P}y\}=\emptyset. In this case, we say that yy is a parent of xx and that xx is a child of yy.

The Hasse diagram of a poset PP is a graphical illustration of its covering relations. Each element of PP is represented by a vertex, and if x<Pyx<_{P}y, then the vertex corresponding to xx is drawn below that corresponding to yy; there exists an edge between these vertices if and only if xPyx\lessdot_{P}y. We say a poset is connected if its Hasse diagram is connected when regarded as a graph; the connected components of PP are the subposets induced by the connected components of the Hasse diagram of PP.

Suppose PP is an nn-element poset, and let f:Pf:P\to\mathbb{Z} be an injective function. Then the standardization of ff, denoted st(f)\mathrm{st}(f), is the labeling L:P[n]L:P\to[n] such that for all x,yPx,y\in P, L(x)<L(y)L(x)<L(y) if and only if f(x)<f(y)f(x)<f(y). Note that this labeling is unique. Equivalently, if g:f(P)[n]g:f(P)\to[n] is an order-preserving bijection, then st(f)=gf\mathrm{st}(f)=g\circ f.

1.2 motivates the following definitions:

Definition 2.1.

Let LL be a labeling of a poset PP. The sorting time of PP is the minimum number kk\in\mathbb{N} such that k(L)(P)\partial^{k}(L)\in\mathcal{L}(P).

Definition 2.2.

For an nn-element poset PP, a labeling is called tangled if it has sorting time n1n-1. A labeling is called quasi-tangled if it has sorting time n2n-2. We also say a labeling LL is kk-promotion-sortable (or just kk-sortable) if k(L)(P)\partial^{k}(L)\in\mathcal{L}(P). We call 1-sortable labelings sortable.

We let the set of all tangled labelings of PP be denoted by 𝒯(P)\mathcal{T}(P); we denote the set of all sortable labelings by Σ(P)\Sigma(P). If PP is a poset, LL a labeling of PP, and γ\gamma\in\mathbb{N}, we also will frequently use the shorthand LγL_{\gamma} to denote γ(L)\partial^{\gamma}(L). Note that L0=LL_{0}=L.

2.2. Frozen Elements and Some Useful Results About Promotion

Definition 2.3.

Let LL be a labeling of an nn-element poset PP. Define aa to be the largest nonnegative integer less than or equal to nn such that for each j{na+1,,n}j\in\{n-a+1,\ldots,n\}, the set {xP|jL(x)n}\{x\in P\;|\;j\leq L(x)\leq n\} forms an upper order ideal in PP. An element xPx\in P is said to be frozen with respect to LL if na+1L(x)nn-a+1\leq L(x)\leq n.

Note that LL is a linear extension if and only if all elements of PP are frozen. Also, we remark that it is possible for a labeling of an nn-element poset to have no frozen elements, namely, when L1(n)L^{-1}(n) is not maximal.

Example 2.4.
76513247542613\partial6731542\partial
Figure 1. The above illustrates how extended promotion works. In the figure, the promotion chain is colored in blue, while the frozen elements are denoted using square-shaped nodes.

Defant and Kravitz proved the following useful results about frozen elements.

Lemma 2.5 ([DK22], Lemma 2.5).

Let LL be a labeling of an nn-element poset PP. If F0F_{0} denotes the set of frozen elements with respect to LL and F1F_{1} denotes the set of frozen elements with respect to L1=(L)L_{1}=\partial(L), then F0F_{0} is properly contained in F1F_{1} unless L(P)L\in\mathcal{L}(P).

Lemma 2.6 ([DK22], Lemma 2.6).

With notation as above, for every 0γn0\leq\gamma\leq n, the elements Lγ1(nγ+1),,Lγ1(n)L_{\gamma}^{-1}(n-\gamma+1),\ldots,L_{\gamma}^{-1}(n) are frozen with respect to LγL_{\gamma}.

Note that the γ=n1\gamma=n-1 case of Lemma 2.6 immediately implies 1.2.

Lemma 2.7.

Let PP be an nn-element poset. For γ\gamma\in\mathbb{N} and any x{2,,n}x\in\{2,\ldots,n\}, we have that Lγ1(x)PLγ+11(x1)L_{\gamma}^{-1}(x)\geq_{P}L_{\gamma+1}^{-1}(x-1); equality holds if and only if Lγ1(x)L_{\gamma}^{-1}(x) is not in the promotion chain of LγL_{\gamma}.

Proof.

Suppose first that Lγ1(x)L_{\gamma}^{-1}(x) is not in the γ\gammath promotion chain. Then Lγ+11(x1)=Lγ1(x)L_{\gamma+1}^{-1}(x-1)=L_{\gamma}^{-1}(x), and we are done. Otherwise, since x>1x>1, there exists an element a<PLγ1(x)a<_{P}L_{\gamma}^{-1}(x) such that aa is in the promotion chain and Lγ1(x)L_{\gamma}^{-1}(x) is the LγL_{\gamma}-successor of aa. Hence, a=Lγ+11(x1)<PLγ1(x)a=L_{\gamma+1}^{-1}(x-1)<_{P}L_{\gamma}^{-1}(x), as desired. ∎

The following results will also be helpful later:

Lemma 2.8.

Let xx and yy be two elements of PP with y<Pxy<_{P}x. Fix some γ{1,,n2}\gamma\in\{1,\ldots,n-2\}, and suppose that Lγ(y)>Lγ(x)L_{\gamma}(y)>L_{\gamma}(x). Then Lγ(y)=Lγ1(y)1L_{\gamma}(y)=L_{\gamma-1}(y)-1. Moreover, we have that Lγ1(y)>Lγ1(x)L_{\gamma-1}(y)>L_{\gamma-1}(x).

Proof.

Begin by letting a=Lγ(y)a=L_{\gamma}(y) and b=Lγ(x)b=L_{\gamma}(x), and note that a>ba>b. Observe that the first part of the lemma holds if and only if yy is not in the (γ1)(\gamma-1)st promotion chain. Assume for a contradiction that yy is in the promotion chain of Lγ1L_{\gamma-1}. By Lemma 2.7, Lγ11(b+1)PLγ1(b)=x>PyL_{\gamma-1}^{-1}(b+1)\geq_{P}L_{\gamma}^{-1}(b)=x>_{P}y, so Lγ11(a+1)L_{\gamma-1}^{-1}(a+1) cannot be the Lγ1L_{\gamma-1}-successor of yy, since a>ba>b. This is a contradiction. The second part of the lemma is simple: Lγ1(x)b+1<a+1=Lγ1(y)L_{\gamma-1}(x)\leq b+1<a+1=L_{\gamma-1}(y). ∎

Theorem 2.9 ([DK22], Theorem 2.10).

For 0kn20\leq k\leq n-2, an nn-element poset PP has a labeling with sorting time nk1n-k-1 if and only if it has a lower order ideal of size k+2k+2 that is not an antichain.

2.3. Rooted Tree and Forest Posets

Definition 2.10.

A rooted forest poset is a poset in which each element is covered by at most one other element. A rooted tree poset is a connected rooted forest poset. Given a rooted tree poset, we say a subset QQ of PP is a subchain (respectively, subtree) if the poset induced by QQ is a chain (respectively, tree). A rooted star poset is a rooted tree poset where the root covers every leaf.

Note here that a rooted tree poset is a poset whose Hasse diagram is a rooted tree where the root is the unique maximal element. A rooted forest poset is a poset whose connected components are rooted tree posets.

3. Promotion Fibers

Given a sorting procedure, it is a fundamental problem to enumerate the objects requiring only one iteration to become sorted. For promotion, Defant and Kravitz enumerated the sortable labelings of a poset in [DK22]. It is then natural to turn our attention to the labelings that can be sorted in two applications of promotion. Note that this problem can be reduced to enumerating the preimages (under \partial) of the sortable labelings.

Definition 3.1.

Let LL be a labeling of a poset PP. We say an element xPx\in P is LL-golden if for all y>Pxy>_{P}x, we have L(y)>L(x)L(y)>L(x). A chain or tree is called LL-golden if all of its elements are LL-golden. If PP is a rooted tree poset, an LL-golden subchain or subtree TT is called maximal if adding any other elements to TT makes it no longer an LL-golden subchain or subtree.

Given a rooted tree poset PP, its highest branch vertex is the largest element xx that covers more than one element. Let PP be a rooted tree poset with root rr and highest branch vertex bb. Note that the set {x|bPxPr}\{x\;|\;b\leq_{P}x\leq_{P}r\} forms a chain. Thus, given a labeling LL of PP, there is a unique maximal LL-golden chain of elements greater than or equal to the highest branch vertex.

Proposition 3.2 ([DK22], Proposition 4.1).

Let PP be an nn-element poset. A labeling LL of PP is in the image of \partial if and only if L1(n)L^{-1}(n) is a maximal element of PP. If Lim()L\in\operatorname{\mathrm{im}}(\partial), then |1(L)||\partial^{-1}(L)| is equal to the number of LL-golden chains of PP containing L1(n)L^{-1}(n).

Let TT be a rooted tree poset with root rr. Given xTx\in T, let x=p1p2pω(x)=rx=p_{1}\lessdot p_{2}\lessdot\cdots\lessdot p_{\omega(x)}=r be the path from xx to the root. In our notation, ω(x)\omega(x) gives the number of elements in this path.

Theorem 3.3.

Let PP be a rooted tree poset with nn elements, and fix a labeling LL of PP. Let c1<P<Pckc_{1}<_{P}\cdots<_{P}c_{k} be the maximal LL-golden chain of elements greater than or equal to the highest branch vertex. Define \mathcal{M} to be the set of all maximal LL-golden rooted subtrees of PP whose roots are less than the maximal branch vertex. If LL is a labeling in im()\operatorname{\mathrm{im}}(\partial) (i.e., if L1(n)L^{-1}(n) is maximal), then

|1(L)|=2k1+TxT2ω(x)+k2,|\partial^{-1}(L)|=2^{k-1}+\sum_{T\in\mathcal{M}}\sum_{x\in T}2^{\omega(x)+k-2},

where ω(x)\omega(x) is computed with respect to TT, not PP.

We start with a preparatory lemma before proving the theorem.

Lemma 3.4.

With notation as in 3.3, let M1(L)M\in\partial^{-1}(L), and let v1<P<Pvmv_{1}<_{P}\cdots<_{P}v_{m} be the promotion chain of MM. The promotion chain is an LL-golden chain containing L1(n)L^{-1}(n), and

{v1,,vm}TT{c1,,ck}.\{v_{1},\ldots,v_{m}\}\subset\bigcup_{T\in\mathcal{M}}T\cup\{c_{1},\ldots,c_{k}\}.
Proof.

Because PP is a rooted tree poset and L1(n)L^{-1}(n) is maximal, the promotion chain of MM is an LL-golden chain containing L1(n)L^{-1}(n). The lemma follows from the identity

TT{c1,,ck}=C𝒞C,\bigcup_{T\in\mathcal{M}}T\cup\{c_{1},\ldots,c_{k}\}=\bigcup_{C\in\mathcal{C}}C,

where 𝒞\mathcal{C} is the set of all LL-golden chains containing L1(n)L^{-1}(n). First, note that ck=L1(n)c_{k}=L^{-1}(n) by assumption. Now, let GG be some LL-golden chain containing L1(n)L^{-1}(n). Suppose that x1<P<Pxix_{1}<_{P}\cdots<_{P}x_{i} are the elements of GG less than the highest branch vertex bb. These form an LL-golden subtree in \mathcal{M}; therefore they are contained in some maximal LL-golden subtree of PP. The remaining elements of GG form some subset of {c1,,ck}\{c_{1},\ldots,c_{k}\}. This shows the inclusion “\supset”; the reverse inclusion is immediate from the definitions. ∎

Proof of 3.3.

We count the labelings M1(L)M\in\partial^{-1}(L). Again take 𝒞\mathcal{C} to be the set of LL-golden chains containing L1(n)L^{-1}(n). By 3.2, we have |1(L)|=|𝒞|.|\partial^{-1}(L)|=|\mathcal{C}|. Let 𝒢\mathcal{G} be the set of LL-golden elements of PP. Note that every element in 𝒞\mathcal{C} contains the root of PP. Let x𝒢x\in\mathcal{G}, and suppose xx is the starting (lowest) element of CC. We count the C𝒞C\in\mathcal{C} with lowest element xx. Every LL-golden element greater than xx is possibly an element in CC; there are ω(x)+k1\omega(x)+k-1 such elements (take ω(x)=0\omega(x)=0 when x{c1,,ck}x\in\{c_{1},\ldots,c_{k}\}). Thus, there are 2ω(x)+k22^{\omega(x)+k-2} possible LL-golden chains starting with xx. Hence,

|𝒞|=x𝒢2ω(x)+k2.|\mathcal{C}|=\sum_{x\in\mathcal{G}}2^{\omega(x)+k-2}.

The theorem then follows from Lemma 3.4 and from noting the following: when the promotion chain of MM consists solely of elements greater than or equal to the highest branch vertex, choosing the promotion chain of MM is just choosing some subset of {c1,,ck1}\{c_{1},\ldots,c_{k-1}\}. ∎

While the formula given in 3.3 is slightly more complicated than

|1(L)|=x𝒢2ω(x)+k2,|\partial^{-1}(L)|=\sum_{x\in\mathcal{G}}2^{\omega(x)+k-2},

the more expanded form is preferable as it allows us to compute |1(L)||\partial^{-1}(L)| easily when PP is a chain. Moreover, we also see immediately that 3.3 is a generalization of Corollary 2.4 of [BCFnt]. 3.3 also lets us enumerate the 2-sortable labelings of rooted tree posets:

Corollary 3.5.

Let PP be an nn-element rooted tree poset. Let Σ(P)=Σ(P)im()\Sigma(P)^{\prime}=\Sigma(P)\cap\operatorname{\mathrm{im}}(\partial) denote the set of sortable labelings of PP in which L1(n)L^{-1}(n) is maximal. Then the number of 2-promotion-sortable labelings of PP is

LΣ(P)|1(L)|=LΣ(P)(2k1+TxT2ω(x)+k2).\sum_{L\in\Sigma(P)^{\prime}}|\partial^{-1}(L)|=\sum_{L\in\Sigma(P)^{\prime}}\left(2^{k-1}+\sum_{T\in\mathcal{M}}\sum_{x\in T}2^{\omega(x)+k-2}\right).

3.3 also yields a similar formula for the kk-sortable labelings for any kk between 1 and n1n-1. However, such a formula would require being able to easily determine if a labeling is (k1)(k-1)-sortable, which is not so easy for k>2k>2, whereas determining if a labeling is sortable is simple.

3.3 also allows us to study the degree of noninvertibility of \partial. In their 2020 paper [DP20], Defant and Propp defined the degree of noninvertibility of a map f:XXf:X\to X to be

deg(f)=1|X|xX|f1(x)|2.\deg(f)=\frac{1}{|X|}\sum_{x\in X}|f^{-1}(x)|^{2}.

For any function f:XXf:X\to X, we have that 1deg(f)|X|1\leq\deg(f)\leq|X|, and the lower bound is achieved when ff is injective, whereas the upper bound is achieved when ff is a constant map. More generally, the degree of a kk-to-1 map is kk. The following corollary can be proven naïvely—by simply computing that \partial is nn-to-1 when PP is a rooted star—or by applying 3.3:

Corollary 3.6.

If PP is a rooted star poset with nn elements, then deg()=n\deg(\partial)=n.

Theorem 3.7.

Let PP be a rooted tree poset with nn elements. Then

ndeg(),n\leq\deg(\partial),

with equality if and only if PP is a rooted star poset.

Proof.

Recall that a labeling LL is in im()\operatorname{\mathrm{im}}(\partial) if and only if L1(n)L^{-1}(n) is maximal. Hence, |im()|=(n1)!|\operatorname{\mathrm{im}}(\partial)|=(n-1)!. The Cauchy-Schwarz inequality gives

(n!)2=(Lim()|1(L)|)2Lim()12Lim()|1(L)|2=|im()|n!deg(),(n!)^{2}=\left(\sum_{L\in\operatorname{\mathrm{im}}(\partial)}|\partial^{-1}(L)|\right)^{2}\leq\sum_{L\in\operatorname{\mathrm{im}}(\partial)}1^{2}\sum_{L\in\operatorname{\mathrm{im}}(\partial)}|\partial^{-1}(L)|^{2}=|\operatorname{\mathrm{im}}(\partial)|n!\deg(\partial),

and rearranging gives ndeg()n\leq\deg(\partial), as desired.

For the second statement, recall that equality in the Cauchy-Schwarz inequality holds if and only if the two vectors are parallel. Hence, equality holds if and only if |1(L)||\partial^{-1}(L)| is the same for all Lim()L\in\operatorname{\mathrm{im}}(\partial). This forces every Lim()L\in\operatorname{\mathrm{im}}(\partial) to be a linear extension: Each labeling in 1(L)\partial^{-1}(L) corresponds to a LL-golden chain containing the root of PP. When LL is linear, every chain containing the root is LL-golden. When LL is not linear, there exists some element xx that is not LL-golden. In particular, if rr is the root of PP, x<Prx<_{P}r is not an LL-golden chain, which reduces |1(L)||\partial^{-1}(L)|, compared to the count for linear extensions. Thus, every labeling in im()\operatorname{\mathrm{im}}(\partial) is linear, which forces PP to be a rooted star poset. ∎

4. Quasi-Tangled Labelings of Inflated Rooted Trees with Deflated Leaves

4.1. Inflated Rooted Trees

Definition 4.1.

Let QQ be a finite poset. An inflation of QQ is a poset PP along with a surjective map φ:PQ\varphi:P\to Q satisfying:

  1. (1)

    For all vQv\in Q, the set φ1(v)\varphi^{-1}(v) has a unique minimal element.

  2. (2)

    If x,yPx,y\in P are such that φ(x)φ(y)\varphi(x)\neq\varphi(y), then x<Pyx<_{P}y if and only if φ(x)<Qφ(y)\varphi(x)<_{Q}\varphi(y).

An inflated rooted tree poset is an inflation of a rooted tree poset. An inflated rooted tree poset with deflated leaves is an inflation of a rooted tree poset QQ such that for all leaves Q\ell\in Q we have |φ1()|=1|\varphi^{-1}(\ell)|=1.

PPQQφ\varphi
Figure 2. (P,φ)(P,\varphi) is an inflated rooted tree with deflated leaves, where PP inflates QQ. The colors illustrate the preimages of the elements of QQ.
Remark 4.2.

We say that a rooted tree poset is reduced if no vertex has exactly one child. Since the composition of inflations is also an inflation, we see that every inflated rooted tree poset is the inflation of a reduced rooted tree poset. When we refer to an inflation φ:PQ\varphi:P\to Q of a rooted tree poset with deflated leaves we will assume that the subposet of QQ obtained by removing its leaves is reduced.

4.2. Setup and Statement of the Main Theorem

In the following, we give a formula enumerating the quasi-tangled labelings of inflated rooted tree posets with deflated leaves. The following is shown in the proof of 2.9, but we state it as its own lemma here:

Lemma 4.3 ([DK22], Proof of Theorem 2.10).

Let PP be an nn-element poset and LL a labeling of PP such that Lnk(P)L_{n-k}\not\in\mathcal{L}(P). Then {Lnk1(1),,Lnk1(k)}\{L_{n-k}^{-1}(1),\ldots,L_{n-k}^{-1}(k)\} forms a lower order ideal of size kk that is not an antichain, and the restriction of LnkL_{n-k} to this set is not a linear extension.

Lemma 4.4.

If LL is a quasi-tangled labeling of an nn-element inflated rooted tree poset PP with deflated leaves, then one of the following holds:

  1. (1)

    L1(n1)L^{-1}(n-1) is minimal;

  2. (2)

    L1(n)L^{-1}(n) is minimal;

  3. (3)

    L1(n)L^{-1}(n) covers a minimal element.

In each case, the element in question is involved in an inversion after n3n-3 promotions.

Proof.

By Lemma 4.3, the set Y={Ln31(1),Ln31(2),Ln31(3)}Y=\{L_{n-3}^{-1}(1),L_{n-3}^{-1}(2),L_{n-3}^{-1}(3)\} forms a lower order ideal that is not an antichain. Moreover, Ln3L_{n-3} restricted to this set is not a linear extension.

Now, note that every non-antichain lower order ideal II with three elements occurring in an inflated rooted tree poset with deflated leaves is one of the following:

  1. (1)

    II is a chain with 3 elements;

  2. (2)

    II is the rooted tree poset on 3 elements with 2 leaves;

  3. (3)

    II is the disjoint union of a singleton and a chain of size 2.

Suppose YY is of type (1). If Ln31(3)L_{n-3}^{-1}(3) is maximal in II, then Ln31(2)Ln31(1)Ln31(3)L_{n-3}^{-1}(2)\lessdot L_{n-3}^{-1}(1)\lessdot L_{n-3}^{-1}(3), and repeatedly applying Lemma 2.8 tells us that L1(n1)L^{-1}(n-1) is minimal. If Ln31(3)L_{n-3}^{-1}(3) is not maximal, then it covers a minimal element or is minimal itself. We may again repeatedly apply Lemma 2.8 to see that L1(n)L^{-1}(n) either covers a minimal element or is minimal.

Suppose YY is of type (2). It follows that Ln31(3)L_{n-3}^{-1}(3) cannot be maximal, because then Ln3L_{n-3} would be a linear extension. So, Ln31(3)L_{n-3}^{-1}(3) is minimal. The element covering Ln31(3)L_{n-3}^{-1}(3) has a smaller label in Ln3L_{n-3}, so L1(n)L^{-1}(n) is minimal by Lemma 2.8.

Lastly, suppose YY is of type (3). If Ln31(3)L_{n-3}^{-1}(3) occupies its own component, then Ln31(2)Ln31(1)L_{n-3}^{-1}(2)\lessdot L_{n-3}^{-1}(1) and L1(n1)L^{-1}(n-1) is minimal. Otherwise, Ln31(3)L_{n-3}^{-1}(3) is covered by either Ln31(1)L_{n-3}^{-1}(1) or Ln31(2)L_{n-3}^{-1}(2), so L1(n)L^{-1}(n) must be minimal. ∎

4.6 enumerates the quasi-tangled labelings of inflated rooted trees with deflated leaves. In particular, note that this relatively large class of posets includes rooted tree posets. While this particular class of posets seems artificial, it will be important that the posets we are working with have the property that if one removes a minimal element or an element covering a minimal element from the poset, then the resulting poset remains an inflated rooted tree.

We remark here that the quasi-tangled labelings of a poset are those labelings LL such that Ln3L_{n-3} is not a linear extension but Ln2L_{n-2} is a linear extension, i.e., LL is not tangled. To count these labelings, we condition on the positions of L1(n1)L^{-1}(n-1) and L1(n)L^{-1}(n) and keep track of where Lγ1(n1γ)L_{\gamma}^{-1}(n-1-\gamma) and Lγ1(n2γ)L_{\gamma}^{-1}(n-2-\gamma) end up (after n3n-3 promotions these are the elements that are labeled 2 and 1, respectively). In particular, we consider a uniformly random labeling LL of PP with L1(n)L^{-1}(n) or L1(n1)L^{-1}(n-1) fixed and calculate the probability that, at each “branch vertex,” the elements in question “get pulled down” in the desired direction, i.e., in the direction such that Ln3L_{n-3} is not a linear extension. The proof of the theorem involves a lot of casework, which we prove beforehand in several technical lemmas.

Let (P,φ)(P,\varphi) be an inflation of a rooted tree poset QQ. For each leaf \ell of QQ, there exists a unique path in the Hasse diagram of QQ from \ell to the root. Let the elements of this path be called u,0,u,1,,u,ω()u_{\ell,0},u_{\ell,1},\ldots,u_{\ell,\omega(\ell)}, where u,0=u_{\ell,0}=\ell and for all 1jω()1\leq j\leq\omega(\ell), u,ju_{\ell,j} covers u,j1u_{\ell,j-1}. Then define

b,j=vQu,j1|φ1(v)|andc,j=v<Qu,j|φ1(v)|.b_{\ell,j}=\sum_{v\leq_{Q}u_{\ell,j-1}}|\varphi^{-1}(v)|\qquad\mathrm{and}\qquad c_{\ell,j}=\sum_{v<_{Q}u_{\ell,j}}|\varphi^{-1}(v)|.

Note that b,jc,j\frac{b_{\ell,j}}{c_{\ell,j}} is the fraction of elements below the minimal element of φ1(u,j)\varphi^{-1}(u_{\ell,j}) that “lie in the direction” of φ1()\varphi^{-1}(\ell). The following enumerates the number of tangled labelings of PP:

Theorem 4.5 ([DK22], Theorem 3.5).

If PP is an inflation of a rooted tree poset QQ, then the number of tangled labelings of PP is

(n1)!i=1sj=1ω(i)bi,j1ci,j1,(n-1)!\sum_{i=1}^{s}\prod_{j=1}^{\omega(i)}\frac{b_{i,j}-1}{c_{i,j}-1},

where ss is the number of leaves of QQ.

Now, define MM to be the set of minimal elements of PP. We define subsets R,S,TR,S,T of MM: Let M\ell\in M, and let xx cover \ell.

  1. (1)

    Put \ell in RR if \ell is the only child of xx and the parent of xx exists and has multiple children.

  2. (2)

    Put \ell in SS if xx has precisely two children;

  3. (3)

    Put \ell in TT if \ell is the only child of xx, the parent of xx exists, and xx is the only child of its parent.

Note here that RR, SS, and TT are disjoint but do not partition PP. However, it is not difficult to see that RR, SS, and TT partition the set of minimal elements of PP that, along with their parents, lie in non-antichain lower order ideals of size 3. See Figure 3.

\ellxxRR\ellxxSS\ellxxTT\ellxxM(RST)M\setminus(R\cup S\cup T)
Figure 3. An illustration of what the elements of R,R, S,S, and TT “look like.”
Theorem 4.6.

Let (P,φ)(P,\varphi) be an inflation of the rooted tree poset QQ with deflated leaves. Let MM, RR, SS, TT, and the u,ju_{\ell,j}’s, b,jb_{\ell,j}’s, and c,jc_{\ell,j}’s be as defined above. Then the number of quasi-tangled labelings of PP is

(n1)!(2Tj=2ω()b,j1c,j11n1Tj=2ω()b,j2c,j2+2Rj=2ω()b,j1c,j1+Sj=2ω()(b,j1)(b,j2)(c,j1)(c,j2)).(n-1)!\left(2\displaystyle\sum_{\ell\in T}\prod_{j=2}^{\omega(\ell)}\frac{b_{\ell,j}-1}{c_{\ell,j}-1}-\frac{1}{n-1}\sum_{\ell\in T}\prod_{j=2}^{\omega(\ell)}\frac{b_{\ell,j}-2}{c_{\ell,j}-2}+2\sum_{\ell\in R}\prod_{j=2}^{\omega(\ell)}\frac{b_{\ell,j}-1}{c_{\ell,j}-1}+\sum_{\ell\in S}\prod_{j=2}^{\omega(\ell)}\frac{(b_{\ell,j}-1)(b_{\ell,j}-2)}{(c_{\ell,j}-1)(c_{\ell,j}-2)}\right).

Corollary 4.7.

Let PP be a poset with nn elements. Suppose that PP has a unique minimal element and that this minimal element has exactly one parent. Then PP has 2(n1)!(n2)!2(n-1)!-(n-2)! quasi-tangled labelings.

4.3. Computing Probabilities

In this section, we compute several probabilities that will help us in the proof of 4.6. Importantly, in Lemma 4.11, we generalize Lemma 3.11 of [DK22], which tells us that the probability of a certain label ending up in some subtree of our inflated rooted tree poset after a certain number of promotions is proportional to the size of the subtree.

Lemma 4.8 ([DK22], Lemma 3.9).

Let PP be an NN-element poset, and let X={yP|y<Px}X=\{y\in P\;|\;y<_{P}x\} for some xPx\in P. Suppose every element of PP that is comparable with some element of XX is also comparable with xx. If LL and L~\tilde{L} are labelings of PP that agree on PXP\setminus X, then for every γ1\gamma\geq 1, the labelings LγL_{\gamma} and L~γ\tilde{L}_{\gamma} also agree on PXP\setminus X.

Lemma 4.9 ([DK22], Lemma 3.10).

Let PP, xx, and XX be defined as in Lemma 4.8. If LL is a labeling of PP and γ0\gamma\geq 0, then the set Lγ(X)L_{\gamma}(X) depends only on the set L(X)L(X) and the restriction LPXL_{P\setminus X}; it does not depend on the way in which the labels in L(X)L(X) are distributed among the elements of XX.

For the rest of this subsection, let PP, xx, and XX be defined as in Lemma 4.8. Suppose there is a partition of XX into disjoint subsets AA and BB such that no element of AA is comparable to an element of BB. Note that both AA and BB are lower order ideals of PP.

Definition 4.10.

Let LL be a labeling of PP. Suppose that k[N1]k\in[N-1] and m[N]m\in[N] are such that m+kNm+k\leq N, Lk1(m)XL_{k}^{-1}(m)\in X, and L1(m+k)XL^{-1}(m+k)\not\in X. We say that γ\gamma pulls down m+km+k if γ<k\gamma<k is the largest index such that Lγ1(m+kγ)XL_{\gamma}^{-1}(m+k-\gamma)\not\in X. We see immediately from the definition that γ\gamma is the unique such index pulling down m+km+k and that γ\gamma pulls down exactly one label.

With notation as above, note that in order to determine whether Lk1(m)AL_{k}^{-1}(m)\in A, it suffices to determine whether Lγ1(1)AL_{\gamma}^{-1}(1)\in A: because Lγ1(m+kγ)XL_{\gamma}^{-1}(m+k-\gamma)\not\in X and Lγ+11(m+kγ1)XL_{\gamma+1}^{-1}(m+k-\gamma-1)\in X, we have that Lγ1(m+kγ)L_{\gamma}^{-1}(m+k-\gamma) is in the γ\gammath promotion chain. Hence, Lγ+11(m+kγ1)AL_{\gamma+1}^{-1}(m+k-\gamma-1)\in A if and only if Lγ1(1)AL_{\gamma}^{-1}(1)\in A (recall AA and BB are disjoint lower order ideals).

To each such m+km+k, one can associate a decreasing sequence of indices γ0,γ1,,γr\gamma_{0},\gamma_{1},\ldots,\gamma_{r} whose values depend only on L|PXL|_{P\setminus X} as follows. Let γ0\gamma_{0} pull down m+km+k. By Lemma 4.9, the value γ0\gamma_{0} depends only on L|PXL|_{P\setminus X}. If L1(γ0+1)XL^{-1}(\gamma_{0}+1)\in X, we are done. Otherwise, let γ1\gamma_{1} pull down γ0+1\gamma_{0}+1, where we let γ0\gamma_{0} take the role of kk and 11 the role of mm (note that γ1<γ0\gamma_{1}<\gamma_{0}). If L1(γ1+1)XL^{-1}(\gamma_{1}+1)\in X, we are done; otherwise, let γ2\gamma_{2} pull down γ1+1\gamma_{1}+1. This process can be continued, where γi+1\gamma_{i+1} pulls down γi+1\gamma_{i}+1. Since 0γi+1<γi0\leq\gamma_{i+1}<\gamma_{i}, this process eventually terminates, yielding a decreasing sequence γ0,,γr\gamma_{0},\ldots,\gamma_{r}. By Lemma 4.9, we see that the values γ0,,γr\gamma_{0},\ldots,\gamma_{r} depend only on the set L(X)L(X) and L|PXL|_{P\setminus X}, not on the way in which the labels in L(X)L(X) are distributed.

Let a,b[N]a,b\in[N] and k[N1]k\in[N-1] be such that a+k,b+kNa+k,b+k\leq N, Lk1(a),Lk1(b)XL_{k}^{-1}(a),L_{k}^{-1}(b)\in X, and L1(a+k),L1(b+k)XL^{-1}(a+k),L^{-1}(b+k)\not\in X. Suppose aba\neq b. Let α0,,αr\alpha_{0},\ldots,\alpha_{r} and β0,,βs\beta_{0},\ldots,\beta_{s} be the sequences associated to a+ka+k and b+kb+k, respectively. We claim that {α0,,αr}{β0,,βs}=\{\alpha_{0},\ldots,\alpha_{r}\}\cap\{\beta_{0},\ldots,\beta_{s}\}=\emptyset. Assume the contrary. If αi=βj\alpha_{i}=\beta_{j} for some 0ir0\leq i\leq r and 0js0\leq j\leq s, then it follows from 4.10 that αr=βs\alpha_{r}=\beta_{s}.

Without loss of generality, suppose rsr\leq s. If r<sr<s, then by definition, αr=βs\alpha_{r}=\beta_{s} pulls down αr1+1=βs1+1\alpha_{r-1}+1=\beta_{s-1}+1, αr1=βs1\alpha_{r-1}=\beta_{s-1} pulls down αr2+1=βs2+1\alpha_{r-2}+1=\beta_{s-2}+1, etc., until α0=βsr\alpha_{0}=\beta_{s-r} pulls down a+k=βsr1+1a+k=\beta_{s-r-1}+1. It follows that a+k1=βsr1a+k-1=\beta_{s-r-1}. Note that a+k1ka+k-1\geq k, but βsr1β0<k\beta_{s-r-1}\leq\beta_{0}<k. This is a contradiction. The case where s>rs>r is identical. If r=sr=s, then α0=β0\alpha_{0}=\beta_{0}, and it follows that a+k=b+ka+k=b+k, contradicting our assumption that aba\neq b.

The following is a generalization of Lemma 3.11 in [DK22]. For d=1d=1, the lemmas are exactly the same. This lemma will be applied repeatedly in the proof of 4.6. Informally, it states that given a list of dd labels whose corresponding elements are in XX after kk promotions, the probability that all of these labels are in AA is proportional to (|A|)!/(|A|d1)!(|A|)!/(|A|-d-1)!.

Lemma 4.11 (Probability Lemma).

Let PP, xx, and XX be defined in Lemma 4.8, and let AA and BB be defined as above. Let k[N1]k\in[N-1], and let n1,,nd[N]n_{1},\ldots,n_{d}\in[N] be such that ni+kNn_{i}+k\leq N for all 1id1\leq i\leq d. Fix an injective map M:PX[N]M:P\setminus X\to[N] such that every labeling LL extending MM has the property that Lk1(n1),,Lk1(nd)XL_{k}^{-1}(n_{1}),\ldots,L_{k}^{-1}(n_{d})\in X. If such an LL is chosen uniformly at random among all such extensions of MM, then the probability that Lk1(n1),,Lk1(nd)AL_{k}^{-1}(n_{1}),\ldots,L_{k}^{-1}(n_{d})\in A is

|A|(|A|1)(|A|d)|X|(|X|1)(|X|d).\frac{|A|(|A|-1)\cdots(|A|-d)}{|X|(|X|-1)\cdots(|X|-d)}.
Proof.

Suppose that n1<n2<<ndn_{1}<n_{2}<\cdots<n_{d}, and let ni1<<nitn_{i_{1}}<\cdots<n_{i_{t}} be the subset of labels such that L1(ni+k)XL^{-1}(n_{i}+k)\not\in X. For all s{i1,,it}s\not\in\{i_{1},\ldots,i_{t}\}, because L1(ns+k)XL^{-1}(n_{s}+k)\in X, Lemma 2.7 gives that Lk1(ns)AL_{k}^{-1}(n_{s})\in A if and only if L1(ns+k)AL^{-1}(n_{s}+k)\in A.

Now, by our discussion above, to each nijn_{i_{j}} we may associate a decreasing sequence of γj\gamma^{j}’s given by γ0j>>γr(j)j\gamma^{j}_{0}>\cdots>\gamma^{j}_{r(j)}. Note that by Lemma 4.9, the set of γj\gamma^{j}’s depends only on MM, not on how the labels in L(X)L(X) are distributed. Recall that the sets {γ0j,,γr(j)j}\{\gamma^{j}_{0},\ldots,\gamma^{j}_{r(j)}\} are pairwise disjoint. Importantly, we have that the γr(j)j\gamma^{j}_{r(j)}’s are all distinct.

Fix some njn_{j} for j{i1,,it}j\in\{i_{1},\ldots,i_{t}\}. In this and the next paragraph, denote the associated sequence by γ0,,γr\gamma_{0},\ldots,\gamma_{r}. We claim that the probability that Lk1(nj)AL_{k}^{-1}(n_{j})\in A is equal to the probability that Lγr1(1)AL_{\gamma_{r}}^{-1}(1)\in A. To see why this is true, recall that γ0\gamma_{0} pulls down nj+kn_{j}+k. In the discussion above, we showed that Lk1(nj)AL_{k}^{-1}(n_{j})\in A if and only if Lγ01(1)AL_{\gamma_{0}}^{-1}(1)\in A. Now, γ1\gamma_{1} pulls down γ0+1\gamma_{0}+1, so Lγ01(1)AL_{\gamma_{0}}^{-1}(1)\in A if and only if Lγ11(1)AL_{\gamma_{1}}^{-1}(1)\in A. Clearly, we may continue in this manner until we see that Lk1(nj)AL_{k}^{-1}(n_{j})\in A if and only if Lγr1(1)AL_{\gamma_{r}}^{-1}(1)\in A.

By assumption, L1(γr+1)XL^{-1}(\gamma_{r}+1)\in X. Because AA and BB are disjoint lower order ideals, Lemma 2.7 implies that Lγr1(1)AL_{\gamma_{r}}^{-1}(1)\in A if and only if L1(γr+1)AL^{-1}(\gamma_{r}+1)\in A. Hence, the probability that Lk1(nj)AL_{k}^{-1}(n_{j})\in A is equal to the probability that L1(γr+1)AL^{-1}(\gamma_{r}+1)\in A.

Thus, for all ii, we have reduced calculating the probability that Lk1(ni)AL_{k}^{-1}(n_{i})\in A to calculating the probability that L1(ai)AL^{-1}(a_{i})\in A for some particular label ai[N]a_{i}\in[N]. Note that our assumptions on MM give L1(ai)XL^{-1}(a_{i})\in X for all ii. For each nijn_{i_{j}}, we have that aij=γr(j)j+1a_{i_{j}}=\gamma^{j}_{r(j)}+1. Recall that the γr(j)j\gamma_{r(j)}^{j}’s are all distinct; moreover note that for all s{i1,,it}s\not\in\{i_{1},\ldots,i_{t}\}, we have that γr(j)j+1ns+k\gamma^{j}_{r(j)}+1\neq n_{s}+k, since γr(j)j+1γ0j+1k<ns+k\gamma^{j}_{r(j)}+1\leq\gamma^{j}_{0}+1\leq k<n_{s}+k. For each s{i1,,it}s\not\in\{i_{1},\ldots,i_{t}\}, as=ns+ka_{s}=n_{s}+k. Thus, the aia_{i}’s are distinct. Since LL is chosen uniformly at random from the labelings extending MM, it follows that the probability Lk1(ni)AL_{k}^{-1}(n_{i})\in A for all ii is

|A|(|A|1)(|A|d)|X|(|X|1)(|X|d),\frac{|A|(|A|-1)\cdots(|A|-d)}{|X|(|X|-1)\cdots(|X|-d)},

as desired. ∎

4.4. Proof of the Main Theorem

Lemma 4.12.

Let PP be an nn-element poset, and let LL be a labeling of PP. Let x0Px_{0}\in P. Define P~=P{x0}\tilde{P}=P\setminus\{x_{0}\} and L~=st(L|P~)\tilde{L}=\mathrm{st}(L|_{\tilde{P}}). Suppose that x0x_{0} is not part of the promotion chain for any of the first γ\gamma promotions. Then st(Lγ|P~)=L~γ\mathrm{st}(L_{\gamma}|_{\tilde{P}})=\tilde{L}_{\gamma}.

Proof.

Recall that promotion depends only on the promotion chain, which in turn depends only on the relative order of the labels. Since x0x_{0} is never in the promotion chain for the first γ\gamma promotions, the promotion chains of L0,,LγL_{0},\ldots,L_{\gamma} and L~0,,L~γ\tilde{L}_{0},\ldots,\tilde{L}_{\gamma} are the same, as desired. ∎

Before proving 4.6, we define some notation. Let PP be as in 4.6, and let x0Px_{0}\in P either cover a unique minimal element or be minimal itself. If x0x_{0} is minimal, let =x0\ell=x_{0}; if it covers a minimal element, denote this minimal element by \ell. Define P~\tilde{P} and L~\tilde{L} as in Lemma 4.12, and let φ~=φ|P~\tilde{\varphi}=\varphi|_{\tilde{P}}. Note that ω()1\omega(\ell)\geq 1 unless QQ is a one-element poset; if QQ has only one element, then so does PP. A one-element poset has no quasi-tangled labelings, so henceforth we assume QQ has more than one element.

For j{2,,ω()}j\in\{2,\ldots,\omega(\ell)\}, let xjx_{j} be the minimal element of φ~1(u,j)\tilde{\varphi}^{-1}(u_{\ell,j}). Also define

Xj={yP~|y<P~xj}andAj=vQu,j1φ~1(v).X_{j}=\{y\in\tilde{P}\;|\;y<_{\tilde{P}}x_{j}\}\quad\mathrm{and}\quad A_{j}=\bigcup_{v\leq_{Q}u_{\ell,j-1}}\tilde{\varphi}^{-1}(v).

Let AjA_{j}^{\prime} be defined analogously but with φ\varphi instead of φ~\tilde{\varphi}. Recall that for j{2,,ω()}j\in\{2,\ldots,\omega(\ell)\},

b,j=vQu,j1|φ1(v)|andc,j=v<Qu,j|φ1(v)|,b_{\ell,j}=\sum_{v\leq_{Q}u_{\ell,j-1}}|\varphi^{-1}(v)|\qquad\mathrm{and}\qquad c_{\ell,j}=\sum_{v<_{Q}u_{\ell,j}}|\varphi^{-1}(v)|,

where u,0,u,1,,u,ω()u_{\ell,0},u_{\ell,1},\ldots,u_{\ell,\omega(\ell)} is the unique path in QQ from φ()\varphi(\ell) to the root.

In order to count the quasi-tangled labelings of PP, we condition on the label of x0x_{0} and count the labelings LL such that Ln3(P)L_{n-3}\not\in\mathcal{L}(P) and there exists yPy\in P such that y>Px0y>_{P}x_{0} and Ln3(y)<Ln3(x0)L_{n-3}(y)<L_{n-3}(x_{0}). For example, when x0x_{0} is minimal, we count the labelings LL such that L(x0)=n1L(x_{0})=n-1 and Ln31(1)>Px0=Ln31(2)L_{n-3}^{-1}(1)>_{P}x_{0}=L_{n-3}^{-1}(2). Note here that Ln31(1)>PLn31(2)L_{n-3}^{-1}(1)>_{P}L_{n-3}^{-1}(2) if and only if Ln31(1)A2L_{n-3}^{-1}(1)\in A_{2}^{\prime}. Our strategy is to fix the label of x0x_{0} and choose a labeling uniformly at random among the (n1)!(n-1)! such labelings of PP; observe that this induces the uniform distribution on the labelings of P~\tilde{P}. Given such a random labeling, we want to calculate the probability that certain labels end up in A2A_{2}^{\prime}.

We will show later that, in each case, calculating this probability can be reduced to calculating the probability that the labels in question end up in A2A_{2}. Thus, we make the following definitions: Let KK be some nonempty subset of {1,2}\{1,2\}. For x0x_{0} and \ell as defined above and j{2,,ω()}j\in\{2,\ldots,\omega(\ell)\}, let E,jE_{\ell,j} be the event that KL~n3(Aj)K\subset\tilde{L}_{n-3}(A_{j}). In other words, E,jE_{\ell,j} is the event that every label in KK ends up on the “correct side” of xjx_{j} after n3n-3 promotions. We would like to compute (E,2)\mathbb{P}(E_{\ell,2}) for L~\tilde{L}. To do so, we note that

(1) (E,2)=(E,ω())(E,ω()1|E,ω())(E,2|E,3)\mathbb{P}(E_{\ell,2})=\mathbb{P}(E_{\ell,\omega(\ell)})\mathbb{P}(E_{\ell,\omega(\ell)-1}\,|\,E_{\ell,\omega(\ell)})\cdots\mathbb{P}(E_{\ell,2}\,|\,E_{\ell,3})

and compute the multiplicands on the right-hand side of the equation above.

x0=x_{0}=\ellx3x_{3}x2x_{2}x1x_{1}u,3u_{\ell,3}u,2u_{\ell,2}u,1u_{\ell,1}u,0u_{\ell,0}PPQQ
Figure 4. An illustration of the notation defined above, where PP is the inflated rooted tree from Figure 2. The black and green boxes denote X3X_{3} and X2X_{2}, respectively, while the red and blue boxes denote A3A_{3} and A2A_{2}, respectively.
Lemma 4.13.

Fix r{1,2}r\in\{1,2\}. Let PP, x0x_{0}, \ell, and the AjA_{j}’s, AjA_{j}^{\prime}’s, and XjX_{j}’s be defined as above. If x0x_{0} is minimal, fix a{n1,n}a\in\{n-1,n\}. Otherwise fix a=na=n. Set L(x0)=aL(x_{0})=a. Then Ln31(r)A2L_{n-3}^{-1}(r)\in A_{2}^{\prime} if and only if L~n31(r)A2\tilde{L}_{n-3}^{-1}(r)\in A_{2}.

Proof.

Suppose x0x_{0} is minimal. Then Ln3(x0)=L(x0)n+3{2,3}L_{n-3}(x_{0})=L(x_{0})-n+3\in\{2,3\}, and x0x_{0} is never in the promotion chain for the first n3n-3 promotions. The lemma follows immediately from applying Lemma 4.12 to PP, LL, and x0x_{0}.

Suppose x0x_{0} covers a unique minimal element \ell and L(x0)=nL(x_{0})=n. Also assume that Ln31(r)A2L_{n-3}^{-1}(r)\in A_{2}^{\prime}. We claim that this implies x0x_{0} is not in the promotion chain for the first n3n-3 promotions. Suppose to the contrary that x0x_{0} is in the γ\gammath promotion chain for some 0γn40\leq\gamma\leq n-4. This forces Lγ()=1L_{\gamma}(\ell)=1 and implies that x0x_{0} is the LγL_{\gamma}-successor of \ell. Note that Ln31(r)A2L_{n-3}^{-1}(r)\in A_{2}^{\prime} implies that Lα1(r+n3α)L_{\alpha}^{-1}(r+n-3-\alpha) is comparable to x0x_{0} for all 0αn30\leq\alpha\leq n-3. In particular, Lγ1(r+n3γ)L_{\gamma}^{-1}(r+n-3-\gamma) must be above x0x_{0}, since x0x_{0} is above only \ell and Lγ()=1L_{\gamma}(\ell)=1. It follows that x0x_{0} cannot be the LγL_{\gamma}-successor of \ell, since r+n3γ<nγ=Lγ(x0)r+n-3-\gamma<n-\gamma=L_{\gamma}(x_{0}). This is a contradiction, so x0x_{0} is not in the promotion chains of L,,Ln4L,\ldots,L_{n-4}. Hence, we may apply Lemma 4.12, and it follows that L~n31(r)A2\tilde{L}_{n-3}^{-1}(r)\in A_{2}.

For the converse, assume that Ln31(r)A2L_{n-3}^{-1}(r)\not\in A_{2}^{\prime}. We have two cases: (1) x0x_{0} is not in the promotion chains of L,,Ln4L,\ldots,L_{n-4}; (2) x0x_{0} is in the promotion chain of LγL_{\gamma} for some γ{0,,n4}\gamma\in\{0,\ldots,n-4\}. For case (1), we simply apply Lemma 4.12 and are done.

For case (2), we note that for all 0αγ0\leq\alpha\leq\gamma, Lemma 4.12 implies that st(Lα|P~)=L~α\mathrm{st}(L_{\alpha}|_{\tilde{P}})=\tilde{L}_{\alpha}. In particular, we have st(Lγ|P~)=L~γ\mathrm{st}(L_{\gamma}|_{\tilde{P}})=\tilde{L}_{\gamma}. Since we are assuming that x0x_{0} is in the promotion chain of LγL_{\gamma}, it follows that Lγ()=1L_{\gamma}(\ell)=1 and that x0x_{0} is the LγL_{\gamma}-successor of \ell. Hence, with respect to LγL_{\gamma}, there are no elements of PP above x0x_{0} with label smaller than nγn-\gamma. In particular, Lγ1(r+n3γ)L_{\gamma}^{-1}(r+n-3-\gamma) is not comparable to x0x_{0} or \ell. Since st(Lγ|P~)=L~γ\mathrm{st}(L_{\gamma}|_{\tilde{P}})=\tilde{L}_{\gamma}, it follows that L~γ1(r+n3γ)\tilde{L}_{\gamma}^{-1}(r+n-3-\gamma) is not comparable to \ell in P~\tilde{P}. Therefore, L~n31(r)A2\tilde{L}_{n-3}^{-1}(r)\not\in A_{2}, as desired. ∎

The next step is using this machinery to compute the conditional probabilities (E,j|E,j+1)\mathbb{P}(E_{\ell,j}|E_{\ell,j+1}) as well as (E,ω())\mathbb{P}(E_{\ell,\omega(\ell)}).

Lemma 4.14.

Let PP, x0x_{0}, \ell, KK, and the E,jE_{\ell,j}’s, AjA_{j}’s, XjX_{j}’s, b,jb_{\ell,j}’s, and c,jc_{\ell,j}’s be defined as above. Let j{2,,ω()1}j\in\{2,\ldots,\omega(\ell)-1\}, and fix any injective map

Mj:P~Xj[n1]M_{j}:\tilde{P}\setminus X_{j}\to[n-1]

such that every labeling L~:P~[n1]\tilde{L}:\tilde{P}\to[n-1] extending MjM_{j} has the property that E,j+1E_{\ell,j+1} occurs. Consider the uniform distribution on such labelings L~\tilde{L}. Then

(E,j|E,j+1)=t=1|K||Aj|t+1|Xj|t+1=t=1|K|b,jtc,jt.\mathbb{P}(E_{\ell,j}\,|\,E_{\ell,j+1})=\prod_{t=1}^{|K|}\frac{|A_{j}|-t+1}{|X_{j}|-t+1}=\prod_{t=1}^{|K|}\frac{b_{\ell,j}-t}{c_{\ell,j}-t}.
Proof.

Recall that we may always assume QQ has more than one element and thus that ω()1\omega(\ell)\geq 1. Also, by Remark 4.2, we have that |A2|2|A_{2}|\geq 2.

By hypothesis, every labeling L~:P~[n1]\tilde{L}:\tilde{P}\to[n-1] extending MjM_{j} has the property that E,j+1E_{\ell,j+1} occurs. Recall that this implies KL~n3(Aj+1)K\subset\tilde{L}_{n-3}(A_{j+1}) and hence that KL~n3(Xj)K\subset\tilde{L}_{n-3}(X_{j}), since {L~n31(1),L~n31(2)}\{\tilde{L}_{n-3}^{-1}(1),\tilde{L}_{n-3}^{-1}(2)\} forms a lower order ideal of P~\tilde{P}. By Lemma 4.9, L~γ\tilde{L}_{\gamma} depends only on L~|P~Xj\tilde{L}|_{\tilde{P}\setminus X_{j}}. Hence, (E,j|E,j+1)\mathbb{P}(E_{\ell,j}\,|\,E_{\ell,j+1}) depends only on L~|P~Xj\tilde{L}|_{\tilde{P}\setminus X_{j}}. Apply the Probability Lemma (Lemma 4.11) with N=n1N=n-1, x=xjx=x_{j}, X=XjX=X_{j}, A=AjA=A_{j}, M=MjM=M_{j}, k=n3k=n-3, and {n1,,nd}=K\{n_{1},\ldots,n_{d}\}=K. This tells us that

(E,j|E,j+1)=t=1|K||Aj|t+1|Xj|t+1.\mathbb{P}(E_{\ell,j}\,|\,E_{\ell,j+1})=\prod_{t=1}^{|K|}\frac{|A_{j}|-t+1}{|X_{j}|-t+1}.

The lemma follows. ∎

Lemma 4.15.

With notation as in the previous lemma, fix any injective map

Mω():P~Xω()[n1],M_{\omega(\ell)}:\tilde{P}\setminus X_{\omega(\ell)}\to[n-1],

and consider the uniform distribution on the labelings L~:P~[n1]\tilde{L}:\tilde{P}\to[n-1] extending Mω()M_{\omega(\ell)}. Then

(E,ω())=t=1|K||Aω()|t+1Xω()t+1=t=1|K|b,ω()tc,ω()t.\mathbb{P}(E_{\ell,\omega(\ell)})=\prod_{t=1}^{|K|}\frac{|A_{\omega(\ell)}|-t+1}{X_{\omega(\ell)}-t+1}=\prod_{t=1}^{|K|}\frac{b_{\ell,\omega(\ell)}-t}{c_{\ell,\omega(\ell)}-t}.
Proof.

We split into cases based on whether or not PP has a unique minimal element. Suppose PP has a unique minimal element. Then Aj=XjA_{j}=X_{j}, and, consequentially, b,j=c,jb_{\ell,j}=c_{\ell,j}. Hence, it suffices to show that the probability in question is 1. Since {L~n31(1),L~n31(2)}\{\tilde{L}_{n-3}^{-1}(1),\tilde{L}_{n-3}^{-1}(2)\} forms a lower order ideal of size 2, it is not difficult to see that when PP has a unique minimal element, {L~n31(1),L~n31(2)}A2\{\tilde{L}_{n-3}^{-1}(1),\tilde{L}_{n-3}^{-1}(2)\}\subset A_{2}. Hence, KL~n3(A2)K\subset\tilde{L}_{n-3}(A_{2}). Because A2AjA_{2}\subset A_{j}, the probability in question is 1, as desired.

Suppose PP does not have a unique minimal element. The argument is identical to that in Lemma 4.14 as long as we show that for any such labeling L~\tilde{L}, KL~n3(Xω())K\subset\tilde{L}_{n-3}(X_{\omega(\ell)}). This simply follows from recalling that {L~n31(1),L~n31(2)}\{\tilde{L}_{n-3}^{-1}(1),\tilde{L}_{n-3}^{-1}(2)\} forms a lower order ideal of size 2, because |P~|=n1|\tilde{P}|=n-1. Since PP does not have a unique minimal element, xω()x_{\omega(\ell)} is greater than at least two elements, implying {L~n31(1),L~n31(2)}Xω()={yP~|y<P~xω()}\{\tilde{L}_{n-3}^{-1}(1),\tilde{L}_{n-3}^{-1}(2)\}\subset X_{\omega(\ell)}=\{y\in\tilde{P}\;|\;y<_{\tilde{P}}x_{\omega(\ell)}\}. ∎

The previous two lemmas only give us information about P~\tilde{P}. In the following, we use Lemma 4.13 to translate these results into information about PP.

Lemma 4.16.

Let φ:PQ\varphi:P\to Q be an inflation of a rooted tree poset with deflated leaves, and let nn be the number of elements in PP. Let x0x_{0} cover a unique minimal element or be minimal itself. If x0x_{0} is a minimal element, let =x0\ell=x_{0}; otherwise let \ell be the minimal element covered by x0x_{0}. If x0x_{0} is minimal, fix a{n1,n}a\in\{n-1,n\}. Otherwise fix a=na=n. Let the AjA_{j}’s, XjX_{j}’s, AjA_{j}^{\prime}’s, E,jE_{\ell,j}’s, b,jb_{\ell,j}’s, and c,jc_{\ell,j}’s be defined as above. Let KK be some nonempty subset of {1,2}\{1,2\}. Suppose L:P[n]L:P\to[n] is a labeling chosen uniformly at random among the (n1)!(n-1)! labelings with L(x0)=aL(x_{0})=a. Then the probability that every label in KK is in Ln3(A2)L_{n-3}(A_{2}^{\prime}) is

j=2ω()t=1|K|b,jtc,jt.\prod_{j=2}^{\omega(\ell)}\prod_{t=1}^{|K|}\frac{b_{\ell,j}-t}{c_{\ell,j}-t}.
Proof.

Note that LL induces the uniform distribution on labelings L~:P~[n1]\tilde{L}:\tilde{P}\to[n-1]. By Lemma 4.13, KLn3(A2)K\subset L_{n-3}(A_{2}^{\prime}) if and only if KL~n3(A2)K\subset\tilde{L}_{n-3}(A_{2}). Thus, we would like to calculate

(E,2)=(E,ω())(E,ω()1|E,ω())(E,2|E,3).\mathbb{P}(E_{\ell,2})=\mathbb{P}(E_{\ell,\omega(\ell)})\mathbb{P}(E_{\ell,\omega(\ell)-1}\,|\,E_{\ell,\omega(\ell)})\cdots\mathbb{P}(E_{\ell,2}\,|\,E_{\ell,3}).

The result follows from applying Lemma 4.15 and Lemma 4.14. ∎

The following three lemmas are applications of Lemma 4.16 to the configurations of interest for the proof of 4.6:

Lemma 4.17.

With notation as in 4.6, the number of labelings LL of PP with L1(n1)SL^{-1}(n-1)\not\in S, L1(n1)L^{-1}(n-1) minimal, and Ln31(2)<PLn31(1)L_{n-3}^{-1}(2)<_{P}L_{n-3}^{-1}(1) is

(n1)!(Tj=2ω()b,j1c,j1+Rj=2ω()b,j1c,j1).(n-1)!\left(\sum_{\ell\in T}\prod_{j=2}^{\omega(\ell)}\frac{b_{\ell,j}-1}{c_{\ell,j}-1}+\sum_{\ell\in R}\prod_{j=2}^{\omega(\ell)}\frac{b_{\ell,j}-1}{c_{\ell,j}-1}\right).
Proof.

Begin by noting that Ln31(2)=L1(n1)L_{n-3}^{-1}(2)=L^{-1}(n-1). Because L1(n1)L^{-1}(n-1) is minimal, Lemma 4.3 gives us that L1(n1)TL^{-1}(n-1)\in T, L1(n1)RL^{-1}(n-1)\in R, or L1(n1)SL^{-1}(n-1)\in S.

Case (1): Assume L1(n1)TL^{-1}(n-1)\in T. We would like to compute the probability that Ln31(1)>PL1(n1)=Ln31(2)L_{n-3}^{-1}(1)>_{P}L^{-1}(n-1)=L_{n-3}^{-1}(2). Note that this event occurs if and only if Ln31(1)A2L_{n-3}^{-1}(1)\in A_{2}^{\prime}. Applying Lemma 4.16 with x0==L1(n1)x_{0}=\ell=L^{-1}(n-1), we see that this probability is just

j=2ω()b,j1c,j1.\prod_{j=2}^{\omega(\ell)}\frac{b_{\ell,j}-1}{c_{\ell,j}-1}.

Summing over the minimal elements in TT, we get the summation corresponding to TT in the formula.

Case (2): An analogous argument works for L1(n1)RL^{-1}(n-1)\in R. The lemma follows. ∎

Lemma 4.18.

With notation as in 4.6, the number of labelings LL of PP with L1(n)L^{-1}(n) minimal and Ln31(3)<PLn31(1)L_{n-3}^{-1}(3)<_{P}L_{n-3}^{-1}(1) or Ln31(3)<PLn31(2)L_{n-3}^{-1}(3)<_{P}L_{n-3}^{-1}(2) is

(n1)!(2Tj=2ω()b,j1c,j1+2Rj=2ω()b,j1c,j1+Sj=2ω()(b,j1)(b,j2)(c,j1)(c,j2)Tj=2ω()(b,j1)(b,j2)(c,j1)(c,j2)).(n-1)!\left(2\displaystyle\sum_{\ell\in T}\prod_{j=2}^{\omega(\ell)}\frac{b_{\ell,j}-1}{c_{\ell,j}-1}+2\sum_{\ell\in R}\prod_{j=2}^{\omega(\ell)}\frac{b_{\ell,j}-1}{c_{\ell,j}-1}\\ +\sum_{\ell\in S}\prod_{j=2}^{\omega(\ell)}\frac{(b_{\ell,j}-1)(b_{\ell,j}-2)}{(c_{\ell,j}-1)(c_{\ell,j}-2)}-\sum_{\ell\in T}\prod_{j=2}^{\omega(\ell)}\frac{(b_{\ell,j}-1)(b_{\ell,j}-2)}{(c_{\ell,j}-1)(c_{\ell,j}-2)}\right).

Proof.

Recall that Lemma 4.3 implies that L1(n)L^{-1}(n) is in either RR, SS, or TT. When L1(n)L^{-1}(n) is in RR or TT, the process of counting the number of such labelings (where L1(n)L^{-1}(n) is minimal and Ln31(3)<PLn31(1)L_{n-3}^{-1}(3)<_{P}L_{n-3}^{-1}(1) or Ln31(3)<PLn31(2)L_{n-3}^{-1}(3)<_{P}L_{n-3}^{-1}(2)) is nearly identical to the one used in the proof of Lemma 4.17 (just apply Lemma 4.16). However, if L1(n)TL^{-1}(n)\in T, it is possible that Ln31(3)<PLn31(1)L_{n-3}^{-1}(3)<_{P}L_{n-3}^{-1}(1) and Ln31(3)<PLn31(2)L_{n-3}^{-1}(3)<_{P}L_{n-3}^{-1}(2); we are twice-counting such labelings. To count the labelings where L1(n)TL^{-1}(n)\in T, Ln31(3)<PLn31(1)L_{n-3}^{-1}(3)<_{P}L_{n-3}^{-1}(1), and Ln31(3)<PLn31(2)L_{n-3}^{-1}(3)<_{P}L_{n-3}^{-1}(2), we apply Lemma 4.16. Thus, for L1(n)RTL^{-1}(n)\in R\cup T, there are

(n1)!(2Tj=2ω()b,j1c,j1+2Rj=2ω()b,j1c,j1Tj=2ω()(b,j1)(b,j2)(c,j1)(c,j2))(n-1)!\left(2\sum_{\ell\in T}\prod_{j=2}^{\omega(\ell)}\frac{b_{\ell,j}-1}{c_{\ell,j}-1}+2\sum_{\ell\in R}\prod_{j=2}^{\omega(\ell)}\frac{b_{\ell,j}-1}{c_{\ell,j}-1}-\sum_{\ell\in T}\prod_{j=2}^{\omega(\ell)}\frac{(b_{\ell,j}-1)(b_{\ell,j}-2)}{(c_{\ell,j}-1)(c_{\ell,j}-2)}\right)

labelings where Ln31(3)<PLn31(1)L_{n-3}^{-1}(3)<_{P}L_{n-3}^{-1}(1) or Ln31(3)<PLn31(2)L_{n-3}^{-1}(3)<_{P}L_{n-3}^{-1}(2). The term being subtracted in the above expression is the number of labelings with L1(n)TL^{-1}(n)\in T, Ln31(3)<PLn31(1)L_{n-3}^{-1}(3)<_{P}L_{n-3}^{-1}(1), and Ln31(3)<PLn31(2)L_{n-3}^{-1}(3)<_{P}L_{n-3}^{-1}(2).

However, the process changes when L1(n)SL^{-1}(n)\in S. Let mm be the other element covered by the parent of L1(n)L^{-1}(n); let YY be the lower order ideal of size 3 consisting of L1(n)L^{-1}(n), mm, and their parent. Note that we must have Y={Ln31(1),Ln31(2),Ln31(3)}Y=\{L_{n-3}^{-1}(1),L_{n-3}^{-1}(2),L_{n-3}^{-1}(3)\}. Setting L1(n)=x0=L^{-1}(n)=x_{0}=\ell, with notation as in Lemma 4.16, we have that Y={Ln31(1),Ln31(2),Ln31(3)}Y=\{L_{n-3}^{-1}(1),L_{n-3}^{-1}(2),L_{n-3}^{-1}(3)\} if and only if Ln31(1)L_{n-3}^{-1}(1) and Ln31(2)L_{n-3}^{-1}(2) are in A2A_{2}^{\prime}. By Lemma 4.16, the probability that both Ln31(1)L_{n-3}^{-1}(1) and Ln31(2)L_{n-3}^{-1}(2) are in A2A_{2}^{\prime} is

j=2ω()(b,j1)(b,j2)(c,j1)(c,j2).\prod_{j=2}^{\omega(\ell)}\frac{(b_{\ell,j}-1)(b_{\ell,j}-2)}{(c_{\ell,j}-1)(c_{\ell,j}-2)}.

The formula follows from summing over all elements in SS. ∎

Lemma 4.19.

With notation as in 4.6, the number of labelings LL where L1(n)L^{-1}(n) covers a minimal element and Ln31(1)<PL1(n)<PLn31(2)L_{n-3}^{-1}(1)<_{P}L^{-1}(n)<_{P}L_{n-3}^{-1}(2) or Ln31(2)<PL1(n)<PLn31(1)L_{n-3}^{-1}(2)<_{P}L^{-1}(n)<_{P}L_{n-3}^{-1}(1) is

(n1)!(mTj=2ω(m)(bm,j1)(bm,j2)(cm,j1)(cm,j2)).(n-1)!\left(\sum_{m\in T}\prod_{j=2}^{\omega(m)}\frac{(b_{m,j}-1)(b_{m,j}-2)}{(c_{m,j}-1)(c_{m,j}-2)}\right).
Proof.

Let x0=L1(n)x_{0}=L^{-1}(n), and let notation be as in Lemma 4.16 so that L1(n)L^{-1}(n) covers some minimal element \ell. Note that T\ell\in T since Ln31(1),L_{n-3}^{-1}(1), Ln31(2)L_{n-3}^{-1}(2), and Ln31(3)=L1(n)L_{n-3}^{-1}(3)=L^{-1}(n) form a lower order ideal. Moreover, note that Ln31(1)L_{n-3}^{-1}(1) and Ln31(2)L_{n-3}^{-1}(2) are comparable to x0x_{0} if and only if they are in A2A_{2}^{\prime}. Hence, we may apply Lemma 4.16 to see that the probability both Ln31(1)L_{n-3}^{-1}(1) and Ln31(2)L_{n-3}^{-1}(2) are comparable to x0x_{0} is

j=2ω(m)(bm,j1)(bm,j2)(cm,j1)(cm,j2).\prod_{j=2}^{\omega(m)}\frac{(b_{m,j}-1)(b_{m,j}-2)}{(c_{m,j}-1)(c_{m,j}-2)}.

Summing over all the elements in TT will imply the lemma. ∎

Proof of 4.6.

We begin by counting the number of tangled labelings of PP. By Lemma 3.8 in [DK22], if LL is a tangled labeling of an nn-element poset, then L1(n)L^{-1}(n) is minimal. Moreover, a labeling is tangled if and only if Ln21(1)>PLn21(2)L_{n-2}^{-1}(1)>_{P}L_{n-2}^{-1}(2). If LL is tangled, then it follows that L1(n)RTL^{-1}(n)\in R\cup T, since {Ln21(1),Ln21(2)}\{L_{n-2}^{-1}(1),L_{n-2}^{-1}(2)\} forms a lower order ideal and since Ln21(1)>PLn21(2)L_{n-2}^{-1}(1)>_{P}L_{n-2}^{-1}(2). Applying Lemma 4.16, we see that there are

(2) (n1)!(Tj=2ω()b,j1c,j1+Rj=2ω()b,j1c,j1)(n-1)!\left(\sum_{\ell\in T}\prod_{j=2}^{\omega(\ell)}\frac{b_{\ell,j}-1}{c_{\ell,j}-1}+\sum_{\ell\in R}\prod_{j=2}^{\omega(\ell)}\frac{b_{\ell,j}-1}{c_{\ell,j}-1}\right)

tangled labelings.

Now, we enumerate the labelings LL such that Ln3(P)L_{n-3}\not\in\mathcal{L}(P). Lemma 4.4 tells us that if Ln3(P)L_{n-3}\not\in\mathcal{L}(P), then either L1(n1)L^{-1}(n-1) is minimal, L1(n)L^{-1}(n) is minimal, or L1(n)L^{-1}(n) covers a minimal element. Moreover, we know that Y={Ln31(1),Ln31(2),Ln31(3)}Y=\{L_{n-3}^{-1}(1),L_{n-3}^{-1}(2),L_{n-3}^{-1}(3)\} forms a lower order ideal of size 3, and Ln3L_{n-3} restricted to YY is not a linear extension. We condition on the three cases given by Lemma 4.4.

Case (1): We count the labelings LL such that L1(n1)L^{-1}(n-1) is minimal, L1(n1)SL^{-1}(n-1)\not\in S, and Ln31(1)>PLn31(2)L_{n-3}^{-1}(1)>_{P}L_{n-3}^{-1}(2). By Lemma 4.17, there are

(3) (n1)!(Tj=2ω()b,j1c,j1+Rj=2ω()b,j1c,j1)(n-1)!\left(\sum_{\ell\in T}\prod_{j=2}^{\omega(\ell)}\frac{b_{\ell,j}-1}{c_{\ell,j}-1}+\sum_{\ell\in R}\prod_{j=2}^{\omega(\ell)}\frac{b_{\ell,j}-1}{c_{\ell,j}-1}\right)

such labelings. If L1(n1)SL^{-1}(n-1)\in S, because {Ln31(1),Ln31(2),Ln31(3)}\{L_{n-3}^{-1}(1),L_{n-3}^{-1}(2),L_{n-3}^{-1}(3)\} is a lower order ideal, it follows from the definition of SS that Ln31(1)L_{n-3}^{-1}(1) must be the unique parent of both Ln31(2)L_{n-3}^{-1}(2) and Ln31(3)L_{n-3}^{-1}(3). Now, repeatedly applying Lemma 2.8 to Ln31(1)L_{n-3}^{-1}(1) and Ln31(3)L_{n-3}^{-1}(3) tells us the position of L1(n)L^{-1}(n), namely that L1(n)=Ln31(3)SL^{-1}(n)=L_{n-3}^{-1}(3)\in S. Thus, this subcase can be excluded and will be addressed in Case (2) when we assume L1(n)L^{-1}(n) is minimal.

Case (2): We count of labelings LL such that L1(n)L^{-1}(n) is minimal and Ln31(3)<PLn31(1)L_{n-3}^{-1}(3)<_{P}L_{n-3}^{-1}(1) or Ln31(3)<PLn31(2)L_{n-3}^{-1}(3)<_{P}L_{n-3}^{-1}(2). By Lemma 4.18, there are

(4)

(n1)!(2Tj=2ω()b,j1c,j1+2Rj=2ω()b,j1c,j1+Sj=2ω()(b,j1)(b,j2)(c,j1)(c,j2)Tj=2ω()(b,j1)(b,j2)(c,j1)(c,j2))(n-1)!\left(2\displaystyle\sum_{\ell\in T}\prod_{j=2}^{\omega(\ell)}\frac{b_{\ell,j}-1}{c_{\ell,j}-1}+2\sum_{\ell\in R}\prod_{j=2}^{\omega(\ell)}\frac{b_{\ell,j}-1}{c_{\ell,j}-1}\\ +\sum_{\ell\in S}\prod_{j=2}^{\omega(\ell)}\frac{(b_{\ell,j}-1)(b_{\ell,j}-2)}{(c_{\ell,j}-1)(c_{\ell,j}-2)}-\sum_{\ell\in T}\prod_{j=2}^{\omega(\ell)}\frac{(b_{\ell,j}-1)(b_{\ell,j}-2)}{(c_{\ell,j}-1)(c_{\ell,j}-2)}\right)

such labelings.

Case (3): We count the labelings LL such that L1(n)L^{-1}(n) covers a minimal element \ell, Ln31(2)>PLn31(3)L_{n-3}^{-1}(2)>_{P}L_{n-3}^{-1}(3) or Ln31(1)>PLn31(3)L_{n-3}^{-1}(1)>_{P}L_{n-3}^{-1}(3), and L()n1L(\ell)\neq n-1. (The case where L()=n1L(\ell)=n-1 and Ln31(1)>PLn31(3)>PLn31(2)L_{n-3}^{-1}(1)>_{P}L_{n-3}^{-1}(3)>_{P}L_{n-3}^{-1}(2) was counted in Case (1).) We first count the labelings LL such that Ln31(2)>PLn31(3)L_{n-3}^{-1}(2)>_{P}L_{n-3}^{-1}(3) or Ln31(1)>PLn31(3)L_{n-3}^{-1}(1)>_{P}L_{n-3}^{-1}(3). Since YY is a lower order ideal of size 3, it follows that Y\ell\in Y. Thus, we may assume that T\ell\in T. Hence, it is sufficient to count the labelings LL such that L1(n)L^{-1}(n) covers some T\ell\in T and Ln31(1)<PL1(n)<PLn31(2)L_{n-3}^{-1}(1)<_{P}L^{-1}(n)<_{P}L_{n-3}^{-1}(2) or Ln31(2)<PL1(n)<PLn31(1)L_{n-3}^{-1}(2)<_{P}L^{-1}(n)<_{P}L_{n-3}^{-1}(1). We have already done this—the number of such labelings is given in Lemma 4.19. Note that each such labeling is indeed quasi-tangled. To account for the condition L()n1L(\ell)\neq n-1, we enumerate the labelings with L1(n1)TL^{-1}(n-1)\in T, L1(n1)PL1(n)L^{-1}(n-1)\lessdot_{P}L^{-1}(n), and Ln31(1)>PLn31(3)L_{n-3}^{-1}(1)>_{P}L_{n-3}^{-1}(3). An adaptation of Lemma 4.16 allows us to enumerate these labelings, the number of which is given by the term being subtracted in the following expression:

(5) (n1)!(mTj=2ω(m)(bm,j1)(bm,j2)(cm,j1)(cm,j2)1n1mTj=2ω(m)(bm,j2)(cm,j2)).(n-1)!\left(\sum_{m\in T}\prod_{j=2}^{\omega(m)}\frac{(b_{m,j}-1)(b_{m,j}-2)}{(c_{m,j}-1)(c_{m,j}-2)}-\frac{1}{n-1}\sum_{m\in T}\prod_{j=2}^{\omega(m)}\frac{(b_{m,j}-2)}{(c_{m,j}-2)}\right).

Note that the above enumerates the labelings LL such that L1(n)L^{-1}(n) covers a minimal element \ell, Ln31(2)>PLn31(3)L_{n-3}^{-1}(2)>_{P}L_{n-3}^{-1}(3) or Ln31(1)>PLn31(3)L_{n-3}^{-1}(1)>_{P}L_{n-3}^{-1}(3), and L()n1L(\ell)\neq n-1.

Summing (3), (4), and (5) and subtracting (2) gives that the number of quasi-tangled labelings of PP is given by

(n1)!(2Tj=2ω()b,j1c,j11n1Tj=2ω()b,j2c,j2+2Rj=2ω()b,j1c,j1+Sj=2ω()(b,j1)(b,j2)(c,j1)(c,j2))(n-1)!\left(2\displaystyle\sum_{\ell\in T}\prod_{j=2}^{\omega(\ell)}\frac{b_{\ell,j}-1}{c_{\ell,j}-1}-\frac{1}{n-1}\sum_{\ell\in T}\prod_{j=2}^{\omega(\ell)}\frac{b_{\ell,j}-2}{c_{\ell,j}-2}+2\sum_{\ell\in R}\prod_{j=2}^{\omega(\ell)}\frac{b_{\ell,j}-1}{c_{\ell,j}-1}+\sum_{\ell\in S}\prod_{j=2}^{\omega(\ell)}\frac{(b_{\ell,j}-1)(b_{\ell,j}-2)}{(c_{\ell,j}-1)(c_{\ell,j}-2)}\right)

as desired. ∎

5. Enumerating Labelings of Rooted Trees with Sorting Time n1kn-1-k

In light of Lemma 4.4 and the Probability Lemma (Lemma 4.11), it is natural to ask if the methods used in Section 4 can be extended to enumerate the labelings of an nn-element poset PP with sorting time n1kn-1-k. In the following, we give an algorithmic approach for doing so when PP is a rooted tree poset. While, theoretically, this approach could yield a general formula for the labelings with sorting time n1kn-1-k, any such formula would be much too complicated to be practical. Instead, for a fixed kk, we offer an algorithmic approach to enumerating the labelings with sorting time n1kn-1-k. Using this method, it would be possible to write a computer program that computes the number of such labelings for a fixed poset.

In order to do so, we first note that Lemma 4.4 generalizes. In particular, for a fixed kk, one can prove that if LL has sorting time n1kn-1-k, then one of the following holds:

  • L1(nk)L^{-1}(n-k) is minimal;

  • L1(nk+1)L^{-1}(n-k+1) is minimal or covers a minimal element;

  • L1(nk+2)L^{-1}(n-k+2) is minimal, covers a minimal element, or is greater than exactly 2 other elements;

  • L1(n)L^{-1}(n) is greater than at most kk other elements.

Algorithm 5.1.

Let PP be an nn-element rooted tree poset. Then we may enumerate the labelings LL of PP with sorting time n1kn-1-k in the following way:

  1. (1)

    List the possible lower order ideals of size kk other than antichains appearing in PP.

  2. (2)

    For each such lower order ideal occurring in PP, use (1) and the Probability Lemma to count the labelings with sorting time n1kn-1-k. This will involve lots of casework based on the positions of L1(nk),,L1(n)L^{-1}(n-k),\ldots,L^{-1}(n) and lower order ideals of size kk occurring in PP. The proof of 4.6 illustrates this casework for the case k=1k=1 in full generality.

6. Tangled Labelings of Posets

In [DK22], Defant and Kravitz conjectured that any nn-element poset has at most (n1)!(n-1)! tangled labelings (see 1.1). This is not obvious even for classes of posets for which we can explicitly enumerate the tangled labelings (e.g., 4.5).

We can prove the conjecture for inflated rooted forests. The following results from [DK22] will be useful:

Corollary 6.1 ([DK22], Corollary 3.7).

Let PP be an nn-element poset with rr connected components, each having a unique minimal element. Then the number of tangled labelings of PP is

(nr)(n2)!.(n-r)(n-2)!.
Theorem 6.2 ([DK22], Theorem 3.4).

Let PP be an nn-element poset with connected components P1,,PrP_{1},\ldots,P_{r}. Let ni=|Pi|n_{i}=|P_{i}|, and let tit_{i} denote the number of tangled labelings of PiP_{i}. The number of tangled labelings of PP is

(n2)!i=1rti(ni2)!.(n-2)!\sum_{i=1}^{r}\frac{t_{i}}{(n_{i}-2)!}.
Lemma 6.3.

Let PP, PiP_{i}, nin_{i}, and tit_{i} be defined as in the above for i=1,,ri=1,\ldots,r. If there are at most (ni1)!(n_{i}-1)! tangled labelings of each PiP_{i}, then there are at most (nr)(n2)!(n-r)(n-2)! tangled labelings of PP.

Proof.

Substituting (ni1)!ti(n_{i}-1)!\geq t_{i} into

(n2)!i=1rti(ni2)!(n2)!i=1r(ni1)=(n2)!(nr)(n-2)!\sum_{i=1}^{r}\frac{t_{i}}{(n_{i}-2)!}\leq(n-2)!\sum_{i=1}^{r}(n_{i}-1)=(n-2)!(n-r)

gives the bound. ∎

Theorem 6.4.

Let PP be an nn-element inflated rooted forest poset. Then PP has at most (n1)!(n-1)! tangled labelings. Equality holds if and only if PP has a unique minimal element.

Proof.

By Lemma 6.3, it suffices to prove this for PP an inflated rooted tree, where QQ is the rooted tree and φ:PQ\varphi:P\to Q the inflation map. Assume without loss of generality that QQ is reduced. We know that the number of tangled labelings of PP is

(6) (n1)!i=1sj=1ω(i)bi,j1ci,j1.(n-1)!\sum_{i=1}^{s}\prod_{j=1}^{\omega(i)}\frac{b_{i,j}-1}{c_{i,j}-1}.

Let 1,,s\ell_{1},\ldots,\ell_{s} denote the leaves of QQ, and let m1,,msm_{1},\ldots,m_{s} be the unique minimal elements of φ1(1),,φ1(s)\varphi^{-1}(\ell_{1}),\ldots,\varphi^{-1}(\ell_{s}), respectively. Suppose without loss of generality that s1\ell_{s-1} and s\ell_{s} have the same parent in QQ (such leaves exist because QQ is assumed to be reduced). For all j1j\neq 1, bs1,j=bs,jb_{s-1,j}=b_{s,j}; for all jj, cs1,j=cs,jc_{s-1,j}=c_{s,j}. Hence, we may rewrite (6) as

(n1)!(i=1s2j=1ω(i)bi,j1ci,j1+j=2ω(s1)bs1,j1cs1,j1(bs1,1+bs,12cs1,11)).(n-1)!\left(\sum_{i=1}^{s-2}\prod_{j=1}^{\omega(i)}\frac{b_{i,j}-1}{c_{i,j}-1}+\prod_{j=2}^{\omega(s-1)}\frac{b_{s-1,j}-1}{c_{s-1,j}-1}\left(\frac{b_{s-1,1}+b_{s,1}-2}{c_{s-1,1}-1}\right)\right).

Now, let PP^{\prime} be the poset obtained from PP by adding the additional relation ms1Pmsm_{s-1}\lessdot_{P^{\prime}}m_{s}. Note that the resulting poset is still an inflated rooted tree poset and that PP^{\prime} is an inflation of QQ^{\prime}, where QQ^{\prime} is the (reduced) rooted tree poset formed by setting s=s1\ell_{s}=\ell_{s-1} and reducing if necessary. Let ψ:PQ\psi:P^{\prime}\to Q^{\prime} be the corresponding inflation map. Moreover, note that QQ^{\prime} has s1s-1 leaves. For each leaf 1,,s1\ell_{1}^{\prime},\ldots,\ell_{s-1}^{\prime} in QQ^{\prime}, let ui,0,,ui,ω(i)u_{i,0}^{\prime},\ldots,u_{i,\omega(i)}^{\prime} denote the unique path from i\ell_{i}^{\prime} to the root of QQ^{\prime}. Let

bi,j=vQui,j1|ψ1(v)|andci,j=v<Qu,j|ψ1(v)|.b_{i,j}^{\prime}=\sum_{v\leq_{Q^{\prime}}u_{i,j-1}^{\prime}}|\psi^{-1}(v)|\qquad\mathrm{and}\qquad c_{i,j}^{\prime}=\sum_{v<_{Q^{\prime}}u_{\ell,j}^{\prime}}|\psi^{-1}(v)|.

Note that for i{1,,s2}i\in\{1,\ldots,s-2\} and j{1,,ω(i)}j\in\{1,\ldots,\omega(i)\}, bi,j=bi,jb_{i,j}=b_{i,j}^{\prime} and ci,j=ci,jc_{i,j}=c_{i,j}^{\prime}. Moreover, we also know that when i=s1i=s-1, bi,j=bi,jb_{i,j}=b_{i,j}^{\prime} and ci,j=ci,jc_{i,j}=c_{i,j}^{\prime} for j=2,,ω(i)j=2,\ldots,\omega(i). When j=1j=1, we have bs1,1=bs1,1+bs,1b_{s-1,1}^{\prime}=b_{s-1,1}+b_{s,1} and ci,j=ci,jc_{i,j}^{\prime}=c_{i,j}. It follows that the number of tangled labelings of PP^{\prime} is

(n1)!(i=1s2j=1ω(i)bi,j1ci,j1+j=2ω(s1)bs1,j1cs1,j1(bs1,1+bs,11cs1,11)).(n-1)!\left(\sum_{i=1}^{s-2}\prod_{j=1}^{\omega(i)}\frac{b_{i,j}-1}{c_{i,j}-1}+\prod_{j=2}^{\omega(s-1)}\frac{b_{s-1,j}-1}{c_{s-1,j}-1}\left(\frac{b_{s-1,1}+b_{s,1}-1}{c_{s-1,1}-1}\right)\right).

Note that PP^{\prime} has more tangled labelings than PP and that PP^{\prime} has s1s-1 minimal elements.

The result follows from induction and Corollary 6.1. ∎

7. Open Problems

In light of 3.7, it is natural to ask for which rooted tree posets deg()\deg(\partial) is biggest. The following is motivated by the fact that \partial is injective if and only if PP is an antichain. Since the degree of noninvertibility of promotion is smallest when PP is an antichain, it seems natural that deg()\deg(\partial) would be largest when PP is a chain. In Remark 2.4 of [DK22], it is shown that when PP is an nn-element chain, \partial is dynamically equivalent to the bubble sort map from [Knu98] (pages 106-110). Theorem 2.2 of [DP20] tells us that when PP is an nn-element chain, deg()=(n+2)(n+1)/6\deg(\partial)=(n+2)(n+1)/{6}.

Conjecture 7.1.

For any poset PP,

deg()(n+2)(n+1)6.\deg(\partial)\leq\frac{(n+2)(n+1)}{6}.

In other words, the degree of noninvertibility of promotion is largest when PP is a chain.

4.6, in conjunction with the enumeration of the tangled labelings of inflated rooted forests given in [DP20], seems to imply that the inflation operation on posets is very compatible with promotion. Thus, it would be a natural next step to study promotion on inflations of non-rooted trees. For example, it would be interesting to enumerate the tangled labelings of inflations of simple posets such as NN-posets or MM-posets. Doing so might generate new methods for attacking 1.1, which may also be refined in the following way (we require PP to be connected because of Lemma 6.3):

Conjecture 7.2.

Let PP be a connected nn-element poset with ss minimal elements. Then PP has at most (ns)(n2)!(n-s)(n-2)! tangled labelings.

It is also possible to reframe 1.1 in the following way: Let PP be a connected, nn-element poset, and let m1,,msm_{1},\ldots,m_{s} be the minimal elements of PP. Let c:P𝒫([s])c:P\to\mathcal{P}([s]) be a coloring of PP given by ic(x)i\in c(x) if xPmix\geq_{P}m_{i} (here 𝒫([s])\mathcal{P}([s]) denotes the power set of [s]={1,,s}[s]=\{1,\ldots,s\}). The following implies 1.1.

Conjecture 7.3.

With notation as above,

(c(Ln21(1))={s}|L(ms1)=n)(c(Ln21(1))={s}|L(ms)=n).\mathbb{P}\left(c(L_{n-2}^{-1}(1))=\{s\}\,|\,L(m_{s-1})=n\right)\geq\mathbb{P}\left(c(L_{n-2}^{-1}(1))=\{s\}\,|\,L(m_{s})=n\right).

If the above holds, we may apply the same argument as in 6.4 to show that the number of tangled labelings increases when we make ms1msm_{s-1}\lessdot m_{s}. Applying this fact repeatedly would prove 1.1, since posets with a unique minimal element have exactly (n1)!(n-1)! tangled labelings.

4.6 and 4.5 together motivate the following conjectures and question:

Conjecture 7.4.

If PP is an inflated rooted tree poset with deflated leaves, then the number of tangled labelings of PP is less than or equal to the number of quasi-tangled labelings of PP.

Conjecture 7.5.

Let PP be an nn-element poset. Then the number of labelings L:P[n]L:P\to[n] such that Ln3(P)L_{n-3}\not\in\mathcal{L}(P) is at most 3(n1)!3(n-1)!.

Question 7.6.

What is the maximum number of quasi-tangled labelings a poset can have?

Acknowledgments

This research was conducted at the Duluth Summer Mathematics Research Program for Undergraduates at the University of Minnesota Duluth with support from Jane Street Capital, the National Security Agency (grant H98230-22-1-0015), the National Science Foundation (grant DMS-2052036), and Harvard University. My research advisors Noah Kravitz, Colin Defant, and Amanda Burcroff provided generous guidance and feedback during the research process for which I am very grateful. I would like to extend special thanks to Swapnil Garg and Noah Kravitz for their invaluable suggestions during the editing process. I would also like to thank Aleksa Milojević for noticing a mistake in one of my proofs early on in the research process. Finally, I am deeply grateful to Joe Gallian for his support and for giving me the opportunity to participate in his research program.

References

  • [BCFnt] Mathilde Bouvel, Lapo Cioni, and Luca Ferrari, Preimages under the bubblesort operator, preprint.
  • [DK22] Colin Defant and Noah Kravitz, Promotion sorting, Order (2022), 1–18.
  • [DP20] Colin Defant and James Propp, Quantifying noninvertibility in discrete dynamical systems, Electron. J. Combin. 27 (2020), no. 3, Paper No. 3.51, 22. MR 4245164
  • [EG87] Paul Edelman and Curtis Greene, Balanced tableaux, Adv. in Math. 63 (1987), no. 1, 42–99. MR 871081
  • [Hai92] Mark D. Haiman, Dual equivalence with applications, including a conjecture of Proctor, Discrete Math. 99 (1992), no. 1-3, 79–113. MR 1158783
  • [Hua20] Brice Huang, Cyclic descents for general skew tableaux, J. Combin. Theory Ser. A 169 (2020), 105120, 45. MR 3983097
  • [Knu98] Donald E. Knuth, The art of computer programming. Vol. 3, Addison-Wesley, Reading, MA, 1998, Sorting and searching, Second edition [of MR0445948]. MR 3077154
  • [MR94] Claudia Malvenuto and Christophe Reutenauer, Evacuation of labelled graphs, Discrete Math. 132 (1994), no. 1-3, 137–143. MR 1297379
  • [PPR09] T. Kyle Petersen, Pavlo Pylyavskyy, and Brendon Rhoades, Promotion and cyclic sieving via webs, J. Algebraic Combin. 30 (2009), no. 1, 19–41. MR 2519848
  • [Rho10] Brendon Rhoades, Cyclic sieving, promotion, and representation theory, J. Combin. Theory Ser. A 117 (2010), no. 1, 38–76. MR 2557880
  • [Sch63] M. P. Schützenberger, Quelques remarques sur une construction de Schensted, Math. Scand. 12 (1963), 117–128. MR 190017
  • [Sch72] by same author, Promotion des morphismes d’ensembles ordonnés, Discrete Math. 2 (1972), 73–94. MR 299539
  • [Sch76] by same author, Evacuations, Colloquio Internazionale sulle Teorie Combinatorie (Rome, 1973), Tomo I, 1976, pp. 257–264. Atti dei Convegni Lincei, No. 17. MR 0476842
  • [Sta09] Richard P. Stanley, Promotion and evacuation, Electron. J. Combin. 16 (2009), no. 2, Special volume in honor of Anders Björner, Research Paper 9, 24. MR 2515772