Book a Demo!
CoCalc Logo Icon
StoreFeaturesDocsShareSupportNewsAboutPoliciesSign UpSign In
GuillaumeLaplante-Anfossi
GitHub Repository: GuillaumeLaplante-Anfossi/Poissons
Path: blob/main/paper/diagonalsPermutahedra.tex
1017 views
1
\documentclass{amsart}
2
3
\usepackage[T1]{fontenc}
4
\usepackage{enumerate, amsmath, amsfonts, amssymb, amsthm, mathrsfs, wasysym, graphics, graphicx, url, hyperref, hypcap, shuffle, xargs, multicol, overpic, pdflscape, multirow, hvfloat, minibox, accents, array, multido, xifthen, a4wide, ae, aecompl, blkarray, pifont, mathtools, etoolbox, dsfont, stmaryrd}
5
\usepackage[dvipsnames]{xcolor}
6
\usepackage{marginnote}
7
\hypersetup{colorlinks=true, citecolor=darkblue, linkcolor=darkblue}
8
\usepackage[all]{xy}
9
\usepackage[bottom]{footmisc}
10
\usepackage{tikz, tikz-cd, tikz-qtree}
11
\usetikzlibrary{trees, decorations, decorations.markings, shapes, arrows, matrix, calc, fit, intersections, patterns, angles, snakes}
12
\usepackage[external]{forest}
13
%\tikzexternalize
14
\graphicspath{{figures/}{figures/nodes/}}
15
\makeatletter\def\input@path{{figures/}}\makeatother
16
\usepackage{caption}
17
\captionsetup{width=\textwidth}
18
\renewcommand{\topfraction}{1} % possibility to have one page of pictures
19
\renewcommand{\bottomfraction}{1} % possibility to have one page of pictures
20
\usepackage[noabbrev,capitalise]{cleveref}
21
\usepackage[export]{adjustbox}
22
\usepackage{ulem}\normalem
23
\usepackage{xargs}
24
25
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
26
27
% colors
28
\definecolor{darkblue}{rgb}{0,0,0.7} % darkblue color
29
\definecolor{green}{RGB}{57,181,74} % green color
30
\definecolor{violet}{RGB}{147,39,143} % violet color
31
32
\newcommand{\red}{\color{red}} % red command
33
\newcommand{\blue}{\color{blue}} % blue command
34
\newcommand{\orange}{\color{orange}} % orange command
35
\newcommand{\green}{\color{green}} % green command
36
\newcommand{\darkblue}{\color{darkblue}} % darkblue command
37
38
% marginal comments
39
\usepackage{todonotes}
40
%\newcommand{\vincent}[1]{\todo[color=blue!30]{\rm #1 \\ \hfill --- V.}}
41
%\newcommand{\Guillaume}[1]{\todo[color=magenta!30]{\rm #1 \\ \hfill --- G.}}
42
%\newcommand{\Kurt}[1]{\todo[color=orange!30]{\rm #1 \\ \hfill --- K.}}
43
%\newcommand{\BDO}[1]{\todo[color=green!30]{\rm #1 \\ \hfill --- B.}}
44
45
% theorems
46
\newtheorem{theorem}{Theorem}[section]
47
\newtheorem*{theorem*}{Theorem}
48
\newtheorem{corollary}[theorem]{Corollary}
49
\newtheorem{proposition}[theorem]{Proposition}
50
\newtheorem{lemma}[theorem]{Lemma}
51
\newtheorem{conjecture}[theorem]{Conjecture}
52
\newtheorem{construction}[theorem]{Construction}
53
54
\theoremstyle{definition}
55
\newtheorem{definition}[theorem]{Definition}
56
\newtheorem{example}[theorem]{Example}
57
\newtheorem{remark}[theorem]{Remark}
58
\newtheorem{question}[theorem]{Question}
59
\newtheorem{problem}[theorem]{Problem}
60
\newtheorem{notation}[theorem]{Notation}
61
\newtheorem{assumption}[theorem]{Assumption}
62
\crefname{notation}{Notation}{Notations}
63
\crefname{problem}{Problem}{Problems}
64
65
% math special letters
66
\newcommand{\R}{\mathbb{R}} % reals
67
\newcommand{\N}{\mathbb{N}} % naturals
68
\newcommand{\Z}{\mathbb{Z}} % integers
69
\newcommand{\C}{\mathbb{C}} % complex
70
\newcommand{\I}{\mathbb{I}} % set of integers
71
\newcommand{\fA}{\mathfrak{A}} % alternating group
72
\newcommand{\fB}{\mathfrak{S}^\textsc{b}} % signed symmetric group
73
\newcommand{\cA}{\mathcal{A}} % algebra
74
\newcommand{\cC}{\mathcal{C}} % collection
75
\newcommand{\cS}{\mathcal{S}} % ground set
76
\newcommand{\uR}{\underline{R}} % underline set
77
\newcommand{\uS}{\underline{S}} % underline set
78
\newcommand{\uT}{\underline{T}} % underline set
79
\newcommand{\oS}{\overline{S}} % overline set
80
\newcommand{\ucS}{\underline{\cS}} % underline ground set
81
\renewcommand{\b}[1]{{\boldsymbol{#1}}} % bold letters
82
\newcommand{\bb}[1]{\mathbb{#1}} % bb letters
83
\newcommand{\f}[1]{\mathfrak{#1}} % frak letters
84
\newcommand{\h}{\widehat} % hat letters
85
\newcommand{\Un}{\mathrm{U}} % unordered IJ
86
\newcommand{\Or}{\mathrm{O}} % ordered IJ
87
\newcommand{\PT}{\mathrm{PT}} %planar trees
88
\newcommand{\K}{\mathrm{K}} %associahedra
89
\newcommand{\J}{\mathrm{J}} %multiplihedra
90
91
% math commands
92
\newcommand{\set}[2]{\left\{ #1 \;\middle|\; #2 \right\}} % set notation
93
\newcommand{\bigset}[2]{\big\{ #1 \;\big|\; #2 \big\}} % big set notation
94
\newcommand{\Bigset}[2]{\Big\{ #1 \;\Big|\; #2 \Big\}} % Big set notation
95
\newcommand{\setangle}[2]{\left\langle #1 \;\middle|\; #2 \right\rangle} % set notation
96
\newcommand{\ssm}{\smallsetminus} % small set minus
97
\newcommand{\dotprod}[2]{\left\langle \, #1 \; \middle| \; #2 \, \right\rangle} % dot product
98
\newcommand{\symdif}{\,\triangle\,} % symmetric difference
99
\newcommand{\one}{\b{1}} % the all one vector
100
\newcommand{\eqdef}{\mbox{\,\raisebox{0.2ex}{\scriptsize\ensuremath{\mathrm:}}\ensuremath{=}\,}} % :=
101
\newcommand{\defeq}{\mbox{~\ensuremath{=}\raisebox{0.2ex}{\scriptsize\ensuremath{\mathrm:}} }} % =:
102
\newcommand{\simplex}{\b{\triangle}} % simplex
103
\renewcommand{\implies}{\Rightarrow} % imply sign
104
\newcommand{\transpose}[1]{{#1}^t} % transpose matrix
105
\renewcommand{\complement}[1]{\bar{#1}} % complement
106
\newcommand{\card}[1]{\##1} % cardinality
107
108
% operators
109
\DeclareMathOperator{\conv}{conv} % convex hull
110
\DeclareMathOperator{\vect}{vect} % linear span
111
\DeclareMathOperator{\cone}{cone} % cone hull
112
\DeclareMathOperator{\Inv}{Inv} % inversions
113
\DeclareMathOperator{\inv}{inv} % inversions
114
\DeclareMathOperator{\Ima}{Im} %Image d'une fonction
115
\DeclareMathOperator{\std}{std}
116
117
%colored nested trees
118
\newcommand{\bluea}[1]{\textcolor{MidnightBlue}{\boldsymbol{\left(\right.}} #1 \textcolor{MidnightBlue}{\boldsymbol{\left.\right)}}}
119
\newcommand{\reda}[1]{\textcolor{Red!60}{\boldsymbol{\left(\right.}} #1 \textcolor{Red!60}{\boldsymbol{\left.\right)}}}
120
\newcommand{\purplea}[1]{\textcolor{Purple!80}{\boldsymbol{\left[\right.}} #1 \textcolor{Purple!80}{\boldsymbol{\left.\right]}}}
121
122
% others
123
\newcommand{\ie}{\textit{i.e.}~} % id est
124
\newcommand{\resp}{resp.~} % id est
125
\newcommand{\eg}{\textit{e.g.}~} % exempli gratia
126
\newcommand{\Eg}{\textit{E.g.}~} % exempli gratia
127
\newcommand{\apriori}{\textit{a priori}} % a priori
128
\newcommand{\viceversa}{\textit{vice versa}} % vice versa
129
\newcommand{\versus}{\textit{vs.}~} % versus
130
\newcommand{\aka}{\textit{a.k.a.}~} % also known as
131
\newcommand{\perse}{\textit{per se}} % per se
132
\newcommand{\ordinal}{\textsuperscript{th}} % th for ordinals
133
\newcommand{\ordinalst}{\textsuperscript{st}} % st for ordinals
134
\newcommand{\defn}[1]{\textsl{\darkblue #1}} % emphasis of a definition
135
\newcommand{\para}[1]{\medskip\noindent\uline{#1.}} % paragraph
136
\renewcommand{\topfraction}{1} % possibility to have one page of pictures
137
\renewcommand{\bottomfraction}{1} % possibility to have one page of pictures
138
\newcommand{\imagetop}[1]{\vtop{\null\hbox{#1}}} % image aligned top
139
\newcommand{\imagebot}[1]{\vbox{\hbox{#1}\null}} % image aligned bot
140
\newcommand{\OEIS}[1]{\cite[{\rm \href{http://oeis.org/#1}{\texttt{#1}}}]{OEIS}}
141
142
%Hyperplane arrangements
143
\renewcommand{\b}[1]{\boldsymbol{#1}} % bold
144
\newcommandx{\arrangement}[1][1 = A]{\mathcal{#1}} % arrangement
145
\newcommand{\HH}{\mathbb{H}} % hyperplane
146
\newcommandx{\braidArrangement}[1][1 = n]{\arrangement[B]_{#1}} % braid arrangement
147
\newcommandx{\multiBraidArrangement}[2][1 = n, 2 = \ell]{\arrangement[B]_{#1}^{#2}} % (l,n)-braid arrangement
148
\newcommandx{\rank}[1][1 = \arrangement]{\operatorname{rk}(#1)} % rank
149
\newcommandx{\facePoset}[1][1 = \arrangement]{\mathsf{Fa}(#1)} % face poset
150
\newcommandx{\fPol}[2][1 = \arrangement, 2 = x]{\b{f}_{#1}(#2)} % face polynomial
151
\newcommandx{\bPol}[2][1 = \arrangement, 2 = x]{\b{b}_{#1}(#2)} % bounded face polynomial
152
\newcommandx{\flatPoset}[1][1 = \arrangement]{\mathsf{Fl}(#1)} % intersection poset
153
\newcommandx{\charPol}[2][1 = \arrangement, 2 = y]{\b{\chi}_{#1}(#2)} % characteristic polynomial
154
\newcommandx{\mobPol}[3][1 = \arrangement, 2 = x, 3 = y]{\b{\mu}_{#1}(#2, #3)} % Mobius polynomial
155
\newcommandx{\weirdPol}[2][1 = \arrangement, 2 = x]{\b{\pi}_{#1}(#2)} % weird polynomial
156
\newcommandx{\partitionPoset}[1][1 = n]{\b{\Pi}_{#1}} % partition poset
157
\newcommandx{\forestPoset}[2][1 = n, 2 = \ell]{\b{\Phi}_{#1}^{#2}} % partition forest poset
158
\newcommandx{\rainbowForests}[2][1 = n, 2 = \ell]{\b{\Psi}_{#1}^{#2}} % rainbow forests
159
\newcommandx{\rainbowTrees}[2][1 = n, 2 = \ell]{\b{\rm{T}}_{#1}^{#2}} % rainbow trees
160
\newcommandx{\Perm}[1][1 = n]{\mathsf{Perm}(#1)} % permutahedron
161
\newcommandx{\Asso}[1][1 = n]{\mathsf{Asso}(#1)} % associahedron
162
\newcommandx{\Cube}[1][1 = n]{\mathsf{Cube}(#1)} % cube
163
\newcommandx{\Simplex}[1][1 = n]{\mathsf{Simplex}(#1)} % simplex
164
\makeatletter
165
\def\rightharpoonupfill@{\arrowfill@\relbar\relbar\rightharpoonup}
166
\newcommand{\overrightharpoonup}{%
167
\mathpalette{\overarrow@\rightharpoonupfill@}}
168
\makeatother
169
\newcommandx{\order}[1]{\smash{\overrightharpoonup{#1}}} % ordered
170
\newcommandx{\orderedPartitionPoset}[1][1 = n]{\order{\b{\Pi}}_{#1}} % oriented partition poset
171
\newcommandx{\orderedForestPoset}[2][1 = n, 2 = \ell]{\order{\b{\Phi}}_{#1}^{#2}} % oriented partition forest poset
172
173
%Operadic diagonals
174
\newcommand{\SU}{\mathrm{SU}}
175
\newcommand{\LA}{\mathrm{LA}}
176
\newcommand{\SUD}{\triangle^{\mathrm{SU}}}
177
\newcommand{\LAD}{\triangle^{\mathrm{LA}}}
178
\newcommand{\SCP}{\mathrm{SCP}}
179
\newcommand{\PolySub}{\mathsf{Poly}}
180
\newcommand{\EC}{\mathcal{E}} %essential complimentary partitions
181
\newcommand{\OP}{\triangle} %ordered partitions
182
\newcommand{\BT}{\mathcal{B}} %bipartite trees
183
\newcommand{\Ainf}{\ensuremath{\mathrm{A}_\infty}}
184
\newcommand{\op}{\mathrm{op}}
185
\newcommand{\tr}{\mathrm{tr}}
186
\newcommand{\id}{\mathrm{id}}
187
188
%Shift operators
189
%\vphantom{1} ensures sequential arrows aligned.
190
\newcommand{\rightshift}[1]{\overrightarrow{\vphantom{1}{#1}}}
191
\newcommand{\rightshiftk}[2]{\overrightarrow{\vphantom{1}{#1}}^{#2}}
192
\newcommand{\leftshift}[1]{\overleftarrow{\vphantom{1}{#1}}}
193
\newcommand{\leftshiftk}[2]{\overleftarrow{\vphantom{1}{#1}}^{#2}}
194
%Subdiv cube defs
195
\newcommand{\divcube}[1]{\Box_{#1}}
196
\newcommandx{\hour}[1][1 = v]{\raisebox{0.065cm}{\rotatebox[origin=c]{90}{$\bowtie$}}_{#1}}
197
\newcommand{\maxsubdivpairsv}{M_v}
198
199
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
200
201
%% from https://tex.stackexchange.com/questions/356564/macro-for-rounded-polygon-around-some-nodes
202
%% answer of bitt.j
203
%Necessary for the mypoly command%
204
\usepackage{tkz-euclide}
205
\usepackage{xstring}
206
207
%------------------------%
208
%---The mypoly command---%
209
%------------------------%
210
211
%--Getting the last Element of a list--%
212
\def\splicelist#1{
213
\StrCount{#1}{,}[\numofelem]
214
\ifnum\numofelem>0\relax
215
\StrBehind[\numofelem]{#1}{,}[\mylast]%
216
\else
217
\let\mylast#1%
218
\fi
219
}
220
221
%--The mypoly macro--%
222
%How to use:
223
%\myroundpoly[decorative commands]{list of names of nodes}{distance}
224
%list of names has to be given in clockwise order
225
\newcommand{\hedge}[3][very thick,color=black]{
226
%Get the last element
227
\splicelist{#2}
228
%Calculate the auxiliary coordinates for the arcs
229
\foreach \vertex [remember=\vertex as \succvertex
230
(initially \mylast)] in {#2}{
231
\coordinate (\succvertex-next) at ($(\succvertex)!#3!90:(\vertex)$);
232
\coordinate (\vertex-previous) at ($(\vertex)!#3!-90:(\succvertex)$);
233
\draw[#1] (\succvertex-next) -- (\vertex-previous);
234
}
235
%Draw the arcs
236
\foreach \vertex in {#2}{
237
\tkzDrawArc[#1](\vertex,\vertex-next)(\vertex-previous)
238
}
239
}
240
241
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
242
243
% formating the part command
244
\makeatletter
245
\def\part{\@startsection{part}{1}%
246
\z@{.7\linespacing\@plus\linespacing}{.8\linespacing}%
247
{\LARGE\sffamily\centering}}
248
%\@addtoreset{section}{part}
249
\makeatother
250
\renewcommand{\thepart}{\Roman{part}}
251
%\renewcommand{\thesection}{\arabic{part}.\arabic{section}}
252
253
% formating the table of contents
254
\setcounter{tocdepth}{2}
255
\makeatletter
256
\def\l@part{\@tocline{1}{8pt}{0pc}{}{}}
257
\def\l@section{\@tocline{1}{4pt}{0pc}{}{}}
258
\makeatother
259
\let\oldtocpart=\tocpart
260
\renewcommand{\tocpart}[2]{\sc\large\oldtocpart{#1}{#2}}
261
\let\oldtocsection=\tocsection
262
\renewcommand{\tocsection}[2]{\bf\oldtocsection{#1}{#2}}
263
\let\oldtocsubsubsection=\tocsubsubsection
264
\renewcommand{\tocsubsubsection}[2]{\quad\oldtocsubsubsection{#1}{#2}}
265
266
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
267
268
\title{Cellular diagonals of permutahedra}
269
270
\thanks{
271
BDO was partially supported by the French ANR grants ALCOHOL (ANR-19-CE40-0006), CARPLO (ANR-20-CE40-0007), HighAGT (ANR-20-CE40-0016) and S3 (ANR-20-CE48-0010).
272
GLA and KS were supported by the Australian Research Council Future Fellowship FT210100256.
273
KS was supported by an Australian Government Research Training Program (RTP) Scholarship.
274
VP was a CNRS researcher at \'Ecole Polytechnique when this work was done, and was partially supported by the French ANR grant CHARMS (ANR-19-CE40-0017) and by the French--Austrian project PAGCAP (ANR-21-CE48-0020 \& FWF I 5788).
275
}
276
277
\author[B. Delcroix-Oger]{B\'er\'enice Delcroix-Oger}
278
\address[B\'{e}r\'{e}nice Delcroix-Oger]{Institut Montpelli\'erain Alexander Grothendieck, Universit\'e de Montpellier, France}
279
\email{berenice.delcroix-oger@umontpellier.fr}
280
\urladdr{\url{https://oger.perso.math.cnrs.fr/}}
281
282
\author[G. Laplante-Anfossi]{Guillaume Laplante-Anfossi}
283
\address[Guillaume Laplante-Anfossi]{School of Mathematics and Statistics, The University of Melbourne, Victoria, Australia}
284
\email{guillaume.laplanteanfossi@unimelb.edu.au}
285
\urladdr{\url{https://guillaumelaplante-anfossi.github.io/}}
286
287
\author[V.~Pilaud]{Vincent Pilaud}
288
\address[Vincent Pilaud]{Universitat de Barcelona}
289
\email{vincent.pilaud@ub.edu}
290
\urladdr{\url{https://www.ub.edu/comb/vincentpilaud/}}
291
292
\author[K. Stoeckl]{Kurt Stoeckl}
293
\address[Kurt Stoeckl]{School of Mathematics and Statistics, The University of Melbourne, Victoria, Australia}
294
\email{kstoeckl@student.unimelb.edu.au}
295
\urladdr{\url{https://kstoeckl.github.io/}}
296
297
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
298
299
\begin{document}
300
301
\begin{abstract}
302
We provide a systematic enumerative and combinatorial study of geometric cellular diagonals on the permutahedra.
303
304
In the first part of the paper, we study the combinatorics of certain hyperplane arrangements obtained as the union of $\ell$ generically translated copies of the classical braid arrangement.
305
Based on Zaslavsky's theory, we derive enumerative results on the faces of these arrangements involving combinatorial objects named partition forests and rainbow forests.
306
This yields in particular nice formulas for the number of regions and bounded regions in terms of exponentials of generating functions of Fuss-Catalan numbers.
307
By duality, the specialization of these results to the case $\ell = 2$ gives the enumeration of any geometric diagonal of the permutahedron.
308
309
In the second part of the paper, we study diagonals which respect the operadic structure on the family of permutahedra.
310
We show that there are exactly two such diagonals, which are moreover isomorphic.
311
We describe their facets by a simple rule on paths in partition trees, and their vertices as pattern-avoiding pairs of permutations.
312
We show that one of these diagonals is a topological enhancement of the Saneblidze--Umble diagonal, and unravel a natural lattice structure on their sets of facets.
313
314
In the third part of the paper, we use the preceding results to show that there are precisely two isomorphic topological cellular operadic structures on the families of operahedra and multiplihedra, and exactly two infinity-isomorphic geometric universal tensor products of homotopy operads and A-infinity morphisms.
315
\end{abstract}
316
317
\vspace*{-1.4cm}
318
\maketitle
319
320
\vspace*{-.7cm}
321
\centerline{\includegraphics[scale=.75]{diagonalPermutahedron3}}
322
\vspace*{-.4cm}
323
324
\newpage
325
\enlargethispage{1.3cm}
326
\tableofcontents
327
328
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
329
330
\newpage
331
\section*{Introduction}
332
\label{s:introduction}
333
334
The purpose of this article is to study \emph{cellular diagonals} on the \emph{permutahedra}, which are cellular maps homotopic to the usual \emph{thin diagonal} $\triangle : P \to P \times P,~ x \mapsto (x,x)$ (\cref{def:thinDiagonal,def:cellularDiagonal}).
335
Such diagonals, and in particular coherent families that we call \emph{operadic diagonals} (\cref{def:operadicDiagonal}), are of interest in algebraic geometry and topology: via the theory of Fulton--Sturmfels~\cite{FultonSturmfels}, they give explicit formulas for the cup product on Losev--Manin toric varieties~\cite{LosevManin}; they define universal tensor products of permutadic $\Ainf$-algebras~\cite{LodayRonco-permutads,Markl}; they define a coproduct on permutahedral sets, which are models of two-fold loop spaces~\cite{SaneblidzeUmble}, and their study is needed to pursue the work of H. J. Baues aiming at defining explicit combinatorial models for higher iterated loop spaces~\cite{Baues}; using the canonical projections to the operahedra, associahedra and multiplihedra, they define universal tensor products of homotopy operads, $\Ainf$-algebras and $\Ainf$-morphisms, respectively~\cite{LaplanteAnfossi,LaplanteAnfossiMazuir}.
336
337
Cellular diagonals for face-coherent families of polytopes are a fundamental object in algebraic topology.
338
The Alexander--Whitney diagonal for simplices~\cite{EilenbergMacLane}, and the Serre map for cubes~\cite{Serre}, allow one to define the cup product in singular simplicial and cubical cohomology.
339
These two diagonals are also needed in the study of iterated loop spaces~\cite{Baues}, while other diagonals are needed in the study of the homology of fibered spaces~\cite{Saneblidze-freeLoopFibration,SaneblidzeRivera, Proute}.
340
In another direction, cellular diagonals allow one to define universal tensor products in homotopical algebra.
341
The seminal case of the \emph{associahedra} has a rich history: the first algebraic diagonal was found by S.~Saneblidze and R.~Umble~\cite{SaneblidzeUmble}, followed by a second one by M.~Markl and S.~Shnider~\cite{MarklShnider}, which was conjectured to coincide with the first one.
342
This has recently been shown to hold~\cite{SaneblidzeUmble-comparingDiagonals}, while a topological enhancement of the \emph{magical formula} of~\cite{MarklShnider} was provided by N.~Masuda, H.~Thomas, A.~Tonks and B.~Vallette~\cite{MasudaThomasTonksVallette}.
343
344
In~\cite{MasudaThomasTonksVallette}, the authors re-introduced the powerful technique of Fulton--Sturmfels~\cite{FultonSturmfels}, which came from the theory of fiber polytopes of~\cite{BilleraSturmfels}, to define a topological cellular diagonal of the associahedra.
345
We shall call such a diagonal a \emph{geometric diagonal} (\cref{def:geometricDiagonal}).
346
There are two remarkable features of this diagonal (or more precisely this family of diagonals, one for the Loday associahedron in each dimension).
347
First, it respects the operadic structure of the associahedra (in fact, forces a unique topological cellular operad structure on them!), that is, the fact that each face of an associahedron is isomorphic to a product of lower-dimensional associahedra.
348
Second, it satisfies the \emph{magical formula} of J.-L. Loday: the faces in the image of the diagonal are given by the pairs of faces which are comparable in the Tamari order (see \cref{sec:cellularDiagonals,rem:magicalFormula} for a precise statement).
349
This magical formula for the associahedra recently lead to new enumerative results for Tamari intervals~\cite{BostanChyzakPilaud}.
350
351
Building on~\cite{MasudaThomasTonksVallette}, a general theory of geometric diagonals was developed in~\cite{LaplanteAnfossi}.
352
In particular, a combinatorial formula describing the image of the diagonal of any polytope was given~\cite[Thm.~1.26]{LaplanteAnfossi}.
353
The topological operad structure of~\cite{MasudaThomasTonksVallette} on the associahedra was generalized to the family of \emph{operahedra}, which comprise the family of permutahedra, and encodes the notion of homotopy operad.
354
Cellular diagonals of the operahedra do \emph{not} satisfy the magical formula, and the combinatorial difficulty of describing their image is what prompted the development of the theory in~\cite{LaplanteAnfossi}.
355
In fact, there is an interesting dichotomy between the families of polytopes which satisfy the magical formula (simplices, cubes, freehedra, associahedra) and those who do not (permutahedra, multiplihedra, operahedra).
356
357
Since the operahedra are \emph{generalized permutahedra}~\cite{Postnikov}, their operadic diagonals are completely determined by the operadic diagonals of permutahedra (see~\cite[Sect.~1.6]{LaplanteAnfossi}), which is the purpose of study of the present paper.
358
359
The first cellular diagonal of the permutahedra was obtained at the algebraic level by S.~Sanebli\-dze and R.~Umble~\cite{SaneblidzeUmble}.
360
We shall call this diagonal the \emph{original $\SU$ diagonal}.
361
The first topological cellular diagonal of the permutahedra was defined in~\cite{LaplanteAnfossi}, we shall call it the \emph{geometric $\LA$ diagonal}.
362
Both of these families of diagonals are \emph{operadic}, \ie they respect the product structure on the faces of permutahedra (this property is called ``comultiplicativity" in~\cite{SaneblidzeUmble}).
363
More precisely, the algebraic structure encoded by the permutahedra is that of permutadic $\Ainf$-algebra~\cite{LodayRonco-permutads,Markl}.
364
365
The toric varieties associated with the permutahedra are called Losev--Manin varieties, introduced in~\cite{LosevManin}.
366
At this level, the operadic structure is that of a reconnectad~\cite{DotsenkoKeilthyLyskov}.
367
The cohomology ring structure was studied by A.~Losev and Y.~Manin, and quite extensively since then, see for instance~\cite{BergstromMinabe, Lin}.
368
Our current work brings a completely combinatorially explicit description of the cup product; it would be interesting to know if this new description can lead to new results, or how it can be used to recover existing ones.
369
370
The first part of the paper derives enumerative results for the iterations of any geometric diagonal of the permutahedra.
371
According to the Fulton--Sturmfels formula~\cite{FultonSturmfels} (see \cref{prop:diagonalCommonRefinement} and \cref{rem:Fulton--Sturmfels}), this amounts to the study of hyperplane arrangements made of generically translated copies of the braid arrangement.
372
The second part studies in depth the combinatorics of operadic diagonals of the permutahedra, providing in particular a topological enhancement of the original $\SU$ diagonal, while the third part derives consequences of this combinatorial study in the field of homotopical algebra.
373
We now proceed to introduce separately each part in more detail.
374
375
%%%%%%%%%%%%%%%
376
377
\subsection*{\cref{part:multiBraidArrangements}. Combinatorics of multiple braid arrangements}
378
379
As the dual of the permutahedron~$\Perm$ is the classical braid arrangement~$\braidArrangement$, the dual of a diagonal of the permutahedron~$\Perm$ is a hyperplane arrangement~$\multiBraidArrangement[n][2]$ made of $2$ generically translated copies of the braid arrangement~$\braidArrangement$.
380
In the first part of the paper, we therefore study the combinatorics of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$, defined as the union of $\ell$ generically translated copies of the braid arrangement~$\braidArrangement$ (\cref{def:multiBraidArrangement}).
381
We are mainly interested in the~$\ell = 2$ case for the enumeration of the faces of the diagonals of the permutahedron~$\Perm$, but the general $\ell$ case is not much harder and corresponds algebraically to the enumeration of the faces of cellular $\ell$-gonals of the permutahedron~$\Perm$.
382
%We observe that the flats of~$\multiBraidArrangement$ are in bijection with forests of partitions and that the faces of~$\multiBraidArrangement$ are in bijection with certain forests of ordered partitions.
383
384
\cref{sec:flatPoset} is dedicated to the combinatorial description of the flat poset of~$\multiBraidArrangement$ and its enumerative consequences.
385
We first observe that the flats of~$\multiBraidArrangement$ are in bijection with \emph{$(\ell,n)$-partition forests}, defined as $\ell$-tuples of (unordered) partitions of~$[n]$ whose intersection hypergraph is a hyperforest (\cref{def:partitionForests}).
386
As this description is independent of the translations of the different copies (as long as these translations are generic), we obtain by T.~Zaslavsky's theory that the number of \mbox{$k$-dimensional} faces and of bounded faces of~$\multiBraidArrangement$ only depends on~$k$, $\ell$, and~$n$. % (and not on the specific translation vectors choosen to construct~$\multiBraidArrangement$).
387
In fact, we obtain the following formula for the M\"obius polynomial of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ in terms of pairs of $(\ell,n)$-partition forests.
388
389
\begin{theorem*}[\cref{thm:MobiusPolynomialMultiBraidArrangement}]
390
The M\"obius polynomial of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ is given by
391
\[
392
\mobPol[\multiBraidArrangement] = x^{n-1-\ell n} y^{n-1-\ell n} \sum_{\b{F} \le \b{G}} \prod_{i \in [\ell]} x^{\card{F_i}} y^{\card{G_i}} \prod_{p \in G_i} (-1)^{\card{F_i[p]}-1} (\card{F_i[p]}-1)! \; ,
393
\]
394
where~$\b{F} \le \b{G}$ ranges over all intervals of the $(\ell,n)$-partition forest poset, and~$F_i[p]$ denotes the restriction of the partition~$F_i$ to the part~$p$ of~$G_i$.
395
\end{theorem*}
396
397
This formula is not particularly easy to handle, but it turns out to simplify to very elegant formulas for the number of vertices, regions, and bounded regions of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$.
398
Namely, using an alternative combinatorial description of the $(\ell,n)$-partition forests in terms of $(\ell, n)$-rainbow forests and a colored analogue of the classical Pr\"ufer code for permutations, we first obtain the number of vertices of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$.
399
400
\begin{theorem*}[\cref{thm:verticesMultiBraidArrangement}]
401
The number of vertices of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ is
402
\[
403
f_0(\multiBraidArrangement) = \ell \big( (\ell-1) n + 1 \big)^{n-2}.
404
\]
405
\end{theorem*}
406
407
This result can even be refined according to the dimension of the flats of the different copies intersected to obtain the vertices of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$.
408
409
\begin{theorem*}[\cref{thm:verticesRefinedMultiBraidArrangement}]
410
For any~$k_1, \dots, k_\ell$ such that~$0 \le k_i \le n-1$ for~$i \in [\ell]$ and~${\sum_{i \in [\ell]} k_i = n-1}$, the number of vertices~$v$ of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ such that the smallest flat of the $i$\ordinal{} copy of~$\braidArrangement$ containing~$v$ has dimension~$n-k_i-1$ is given by
411
\[
412
n^{\ell-1} \binom{n-1}{k_1, \dots, k_\ell} \prod_{i \in [\ell]} (n-k_i)^{k_i-1}.
413
\]
414
\end{theorem*}
415
416
We then consider the regions of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$.
417
We first obtain a very simple exponential formula for its characteristic polynomial.
418
419
\begin{theorem*}[\cref{thm:characteristicPolynomialMultiBraidArrangement}]
420
The characteristic polynomial~$\charPol[\multiBraidArrangement]$ of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ is given~by
421
\[
422
\charPol[\multiBraidArrangement] = \frac{(-1)^n n!}{y} \, [z^n] \, \exp \bigg( - \sum_{m \ge 1} \frac{F_{\ell,m} \, y \, z^m}{m} \bigg) ,
423
\]
424
where~$\displaystyle F_{\ell,m} \eqdef \frac{1}{(\ell-1)m+1} \binom{\ell m}{m}$ is the Fuss-Catalan number.
425
\end{theorem*}
426
427
Evaluating the characteristic polynomial at~$y = -1$ and~$y = 1$ respectively, we obtain by T.~Zaslavsky's theory the numbers of regions and bounded regions of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$.
428
429
\begin{theorem*}[\cref{thm:regionsMultiBraidArrangement}]
430
The numbers of regions and of bounded regions of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ are given by
431
\begin{align*}
432
f_{n-1}(\multiBraidArrangement)
433
& = n! \, [z^n] \exp \Bigg( \sum_{m \ge 1} \frac{F_{\ell,m} \, z^m}{m} \Bigg) \\
434
\text{and}\qquad
435
b_{n-1}(\multiBraidArrangement)
436
%& = - n! \, [z^n] \exp \Bigg( - \sum_{m \ge 1} \frac{F_{\ell,m} \, z^m}{m} \Bigg)
437
& = (n-1)! \, [z^{n-1}] \exp \bigg( (\ell-1) \sum_{m \ge 1} F_{\ell,m} \, z^m \bigg),
438
\end{align*}
439
where~$\displaystyle F_{\ell,m} \eqdef \frac{1}{(\ell-1)m+1} \binom{\ell m}{m}$ is the Fuss-Catalan number.
440
\end{theorem*}
441
442
Finally, \cref{sec:facePoset} is dedicated to the combinatorial description of the face poset of~$\multiBraidArrangement$.
443
We observe that the faces of~$\multiBraidArrangement$ are in bijection with certain \emph{ordered} $(\ell,n)$-partition forests, defined as $\ell$-tuples of ordered partitions of~$[n]$ whose underlying unordered partitions form an (unordered) $(\ell,n)$-partition forest (\cref{def:orderedPartitionForest}).
444
Here, which ordered $(\ell,n)$-partition forests actually appear as faces of~$\multiBraidArrangement$ depends on the choice of the translations of the different copies.
445
We provide a combinatorial description of the possible orderings of a $(\ell,n)$-partition forest compatible with some given translations in terms of certain paths in the forest (\cref{prop:PFtoOPF1,prop:PFtoOPF2}), and a combinatorial characterization of the ordered partition forests which appear for some given translations in terms of the circuits of a certain oriented graph (\cref{prop:characterizationOPFs}).
446
447
%%%%%%%%%%%%%%%
448
449
\subsection*{\cref{part:diagonalsPermutahedra}. Diagonals of permutahedra}
450
451
We present cellular diagonals, the Fulton--Sturmfels method, the magical formula and specialize the results of \cref{part:multiBraidArrangements} to the permutahedra in \cref{sec:cellularDiagonals}.
452
Then, we initiate in \cref{sec:operadicDiagonals} the study of operadic diagonals (\cref{def:operadicDiagonal}).
453
These are families of diagonals of the permutahedra which are compatible with the property that faces of permutahedra are product of lower-dimensional permutahedra.
454
455
\begin{theorem*}[\cref{thm:unique-operadic,thm:bijection-operadic-diagonals}]
456
There are exactly four operadic geometric diagonals of the permutahedra, the geometric $\LA$ and $\SU$ diagonals and their opposites, and only the first two respect the weak order on permutations.
457
Moreover, their cellular images are isomorphic as posets.
458
\end{theorem*}
459
460
It turns out that the facets and vertices of operadic diagonals admit elegant combinatorial descriptions.
461
The following is a consequence of a general geometrical result, that holds for any diagonal (\cref{prop:PFtoOPF1}).
462
463
\begin{theorem*}[\cref{thm:facet-ordering}]
464
A pair of ordered partitions $(\sigma,\tau)$ forming a partition tree is a facet of the $\LA$ (\resp $\SU$) geometric diagonal if and only if the minimum (\resp maximum) of every directed path between two consecutive blocks of $\sigma$ or $\tau$ is oriented from $\sigma$ to $\tau$ (\resp from $\tau$ to $\sigma$).
465
\end{theorem*}
466
467
\pagebreak
468
Vertices of operadic diagonals are pairs of permutations, and form a strict subset of intervals of the weak order.
469
They admit an analogous description in terms of pattern-avoidance.
470
471
\begin{theorem*}[\cref{thm:patterns}]
472
A pair of permutations of $[n]$ is a vertex of the $\LA$ (\resp $\SU$) diagonal if and only if for any~$k\geq 1$ and for any $I=\{i_1, \dots, i_k\},J=\{j_1, \dots, j_k\} \subset [k]$ such that $i_1=1$ (\resp $j_k=2k$) it avoids the patterns
473
\begin{align}
474
(j_1 i_1 j_2 i_2 \cdots j_k i_k,\ i_2 j_1 i_3 j_2 \cdots i_k j_{k-1} i_1 j_k), \tag{LA} \\
475
\text{ \resp } (j_1 i_1 j_2 i_2 \cdots j_k i_k, \ i_1 j_k i_2 j_1 \cdots i_{k-1} j_{k-2}i_k j_{k-1}), \tag{SU}
476
\end{align}
477
For each $k \ge 1$, there are $\binom{2k-1}{k-1,k}(k-1)!k!$ such patterns, which are $(21,12)$ for $k=1$, and the following for $k=2$
478
\begin{itemize}
479
\item $\LA$ avoids
480
$(3142,2314), (4132,2413),
481
(2143,3214), (4123,3412),
482
(2134,4213), (3124,4312)$,
483
\item $\SU$ avoids
484
$(1243,2431),(1342,3421),
485
(2143,1432),(2341,3412),
486
(3142,1423),(3241,2413)$.
487
\end{itemize}
488
\end{theorem*}
489
490
In \cref{sec:shifts}, we introduce \emph{shifts} that can be performed on the facets of operadic diagonals.
491
These allow us to show that the geometric $\SU$ diagonal is a topological enhancement of the original $\SU$ diagonal.
492
493
\begin{theorem*}[\cref{thm:recover-SU}]
494
The original and geometric $\SU$ diagonals coincide.
495
\end{theorem*}
496
497
The proof of this result, quite technical, proceeds by showing the equivalence between $4$ different descriptions of the diagonal: the original, $1$-shift, $m$-shift and geometric $\SU$ diagonals (\cref{subsec:topological-SU}).
498
This brings a positive answer to~\cite[Rem.~2.19]{LaplanteAnfossi}, showing that the original $\SU$ diagonal can be recovered from a choice of chambers in the fundamental hyperplane arrangement of the permutahedron.
499
Our formulas for the number of facets also agrees with the experimental count made in~\cite{VejdemoJohansson}.
500
Moreover, it provides a new proof that all known diagonals on the associahedra coincide~\cite{SaneblidzeUmble-comparingDiagonals}.
501
Indeed, since the family of vectors inducing the geometric $\SU$ diagonal all have strictly decreasing coordinates, the diagonal induced on the associahedron is given by the magical formula~\cite[Thm.~2]{MasudaThomasTonksVallette}, see also~\cite[Prop.~3.8]{LaplanteAnfossi}.
502
503
The above theorem also allows us to translate the different combinatorial descriptions of the facets of operadic diagonals from one to the other, compiled in the following table.
504
505
\begin{figure}[h!]
506
\begin{center}
507
\begin{tabular}{c|c|c}
508
Description & $\SU$ diagonal & $\LA$ diagonal \\
509
\hline
510
Original & \cite{SaneblidzeUmble} & \cref{def:classical-LA} \\
511
Geometric & \cref{thm:minimal} & \cite{LaplanteAnfossi} \\
512
Path extrema & \cref{thm:facet-ordering} & \cref{thm:facet-ordering} \\
513
$1$-shifts & \cref{def:classical-SU} & \cref{def:classical-LA} \\
514
$m$-shifts & \cref{def:classical-SU} & \cref{def:classical-LA} \\
515
Lattice & \cref{prop:shift lattice} & \cref{prop:shift lattice} \\
516
Cubical & \cite{SaneblidzeUmble} & \cref{prop:LA-cubical} \\
517
Matrix & \cite{SaneblidzeUmble} & \cref{subsec:matrix}
518
\end{tabular}
519
\end{center}
520
\end{figure}
521
522
In \cref{sec:Shift-lattice}, we show that the facets of operadic diagonals are disjoint unions of lattices, that we call the \emph{shift lattices}.
523
These lattices are isomorphic to a product of chains, and are indexed by the permutations of $[n]$.
524
Moreover, while the pairs of facets of operadic diagonals are intervals of the facial weak order (\cref{sec:facial-weak-order}), the shift lattices are not sub-lattices of this order's product (see \cref{fig: Inversion and lattice counter example}).
525
526
Finally, we present the alternative cubical (\cref{sec:Cubical}) and matrix (\cref{subsec:matrix}) descriptions of the $\SU$ diagonal from \cite{SaneblidzeUmble,SaneblidzeUmble-comparingDiagonals}, providing proofs of their equivalence with the other descriptions, and giving their $\LA$ counterparts.
527
The existence of this cubical description, based on a subdivision of the cube combinatorially isomorphic to the permutahedron, finds its conceptual root in the bar-cobar resolution of the associative permutad.
528
Indeed, this resolution is encoded by the dual subdivision of the permutahedron, which is cubical since $\Perm$ is a simple polytope, and a diagonal can be obtained from the classical Serre diagonal via retraction, in the same fashion as for the associahedra, see \cite{MarklShnider, Loday-diagonal} and \cite[Sec. 5.1]{LaplanteAnfossiMazuir}.
529
530
%%%%%%%%%%%%%%%
531
532
\pagebreak
533
\subsection*{\cref{part:higherAlgebraicStructures}. Higher algebraic structures}
534
535
In this shorter third part of the paper, we derive some higher algebraic consequences of the preceding results, which were the original motivation for the present study.
536
They concern the \emph{operahedra}, a family of polytopes indexed by planar trees, which encode (non-symmetric non-unital) homotopy operads \cite{LaplanteAnfossi}, and the \emph{multiplihedra}, a family of polytopes indexed by $2$-colored nested linear trees, which encode $\Ainf$-morphisms \cite{LaplanteAnfossiMazuir}.
537
Both of these admit realizations ``\`a la Loday", which generalize the Loday realizations of the associahedra.
538
The faces of an operahedron are in bijection with \emph{nestings}, or parenthesization, of the corresponding planar tree, while the faces of a multiplihedron are in bijection with $2$-colored nestings of the corresponding linear tree.
539
The main results concerning the operahedra are summarized as follows.
540
541
\begin{theorem*}[\cref{thm:operahedra,thm:top-iso,thm:infinity-iso}]
542
There are exactly
543
\begin{enumerate}
544
\item two geometric operadic diagonals of the Loday operahedra, the $\LA$ and $\SU$ diagonals,
545
\item two geometric topological cellular colored operad structures on the Loday operahedra,
546
\item two geometric universal tensor products of homotopy operads,
547
\end{enumerate}
548
which agree with the generalized Tamari order on fully nested trees.
549
Moreover, the two topological operad structures are isomorphic, and the two tensor products are not strictly isomorphic, but are related by an $\infty$-isotopy.
550
\end{theorem*}
551
552
As the associahedra and the permutahedra are part of the family of operahedra, we get analogous results for $\Ainf$-algebras and permutadic $\Ainf$-algebras.
553
The main results concerning the multiplihedra are summarized as follows.
554
555
\begin{theorem*}[{\cref{thm:multiplihedra,thm:top-iso-2,thm:infinity-iso-2}}]
556
There are exactly
557
\begin{enumerate}
558
\item two geometric operadic diagonals of the Forcey multiplihedra, the $\LA$ and $\SU$ diagonals,
559
\item two geometric topological cellular operadic bimodule structures (over the Loday associahedra) on the Forcey multiplihedra,
560
\item two compatible geometric universal tensor products of $\Ainf$-algebras and $\Ainf$-morphisms,
561
\end{enumerate}
562
which agree with the Tamari-type order on atomic $2$-colored nested linear trees.
563
Moreover, the two topological operadic bimodule structures are isomorphic, and the two tensor products are not strictly isomorphic, but are related by an $\infty$-isotopy.
564
\end{theorem*}
565
566
Here, by the adjective ``geometric", we mean diagonal, operadic structure and tensor product which are obtained geometrically on the polytopes via the Fulton--Sturmfelds method.
567
By ``universal", we mean formulas for the tensor products which apply to \emph{any} pair of homotopy operads or $\Ainf$-morphisms.
568
569
However, the isomorphisms of topological operads (\resp operadic bimodules) takes place in a category of polytopes $\PolySub$ for which the morphisms are \emph{not} affine maps \cite[Def. 4.13]{LaplanteAnfossi}, and it does \emph{not} commute with the diagonal maps (\cref{ex:iso-not-Hopf,ex:iso-not-Hopf-2}).
570
Moreover, the pairs of faces in the image of the two operadic diagonals are in general not in bijection (see \cref{ex:operahedra-LA-SU,ex:multiplihedra-LA-SU}), yielding different (but $\infty$-isomorphic) tensor products of homotopy operads (\resp \mbox{$\Ainf$-morphisms}).
571
572
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
573
574
\section*{Acknowledgements}
575
576
We are indebted to Matthieu Josuat-Vergès for taking part in the premises of this paper, in particular for conjecturing the case $\ell = 2$ of \cref{thm:verticesRefinedMultiBraidArrangement}, for working on the proof of \cref{thm:patterns}, and for implementing some sage code used during the project.
577
GLA is grateful to Hugh Thomas for raising the question to count the facets of the $\LA$ diagonal and for preliminary discussions during a visit at the LACIM in the summer of 2021 where the project started, and to the Max Planck Institute for Mathematics in Bonn where part of this work was carried out.
578
We are grateful to Sylvie Corteel for pointing out the relevance of the dual perspective for the purpose of counting.
579
VP is grateful to the organizers (Karim Adiprasito, Alexey Glazyrin, Isabella Novic, and Igor Pak) of the workshop ``Combinatorics and Geometry of Convex Polyhedra'' held at the Simons Center for Geometry and Physics in March 2023 for the opportunity to present a preliminary version of our enumerative results, and to Pavel Galashin for asking about the characteristic polynomial of the multiple braid arrangement during this presentation.
580
We thank Samson Saneblidze and Ron Umble for useful correspondence.
581
582
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
583
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
584
585
\clearpage
586
\part{Combinatorics of multiple braid arrangements}
587
\label{part:multiBraidArrangements}
588
589
In this first part, we study the combinatorics of hyperplane arrangements obtained as unions of generically translated copies of the braid arrangement.
590
In \cref{sec:arrangements}, we first recall some classical facts on the enumeration of hyperplane arrangements (\cref{subsec:arrangements}), present the classical braid arrangement (\cref{subsec:braidArrangement}), and define our multiple braid arrangements (\cref{subsec:multiBraidArrangement}).
591
Then in \cref{sec:flatPoset}, we describe their flat posets in terms of partition forests (\cref{subsec:partitionForests}) and rainbow forests (\cref{subsec:rainbowForests}), from which we derive their M\"obius polynomials (\cref{subsec:MobiusPolynomialMultiBraidArrangement}), and some surprising formulas for their numbers of vertices (\cref{subsec:verticesMultiBraidArrangement}) and regions (\cref{subsec:regionsMultiBraidArrangement}).
592
Finally, in \cref{sec:facePoset}, we describe their face posets in terms of ordered partition forests (\cref{subsec:orderedPartitionForests}), and explore some combinatorial criteria to describe the ordered partition forests that appear as faces of a given multiple braid arrangement (\cref{subsec:PFtoOPF,subsec:criterionOPF}).
593
594
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
595
596
\section{Recollection on hyperplane arrangements and braid arrangements}
597
\label{sec:arrangements}
598
599
%%%%%%%%%%%%%%%
600
601
\subsection{Hyperplane arrangements}
602
\label{subsec:arrangements}
603
604
We first briefly recall classical results on the combinatorics of affine hyperplane arrangements, in particular the enumerative connection between their intersection posets and their face lattices due to T.~Zaslavsky~\cite{Zaslavsky}.
605
606
\begin{definition}
607
A finite affine real \defn{hyperplane arrangement} is a finite set~$\arrangement$ of affine hyperplanes in~$\R^d$.
608
\end{definition}
609
610
\begin{definition}
611
A \defn{region} of~$\arrangement$ is a connected component of~$\R^d \ssm \bigcup_{H \in \arrangement} H$.
612
The \defn{faces} of~$\arrangement$ are the closures of the regions of~$\arrangement$ and all their intersections with a hyperplane of~$\arrangement$.
613
The \defn{face poset} of~$\arrangement$ is the poset~$\facePoset$ of faces of~$\arrangement$ ordered by inclusion.
614
The \defn{$f$-polynomial}~$\fPol$ and \defn{$b$-polynomial}~$\bPol$ of~$\arrangement$ are the polynomials
615
\[
616
\fPol \eqdef \sum_{k = 0}^d f_k(\arrangement) \, x^k
617
\qquad\text{and}\qquad
618
\bPol \eqdef \sum_{k = 0}^d b_k(\arrangement) \, x^k ,
619
\]
620
where~$f_k(\arrangement)$ denotes the number of $k$-dimensional faces of~$\arrangement$, while~$b_k(\arrangement)$ denotes the number of bounded $k$-dimensional faces of~$\arrangement$.
621
\end{definition}
622
623
\begin{definition}
624
A \defn{flat} of~$\arrangement$ is a non-empty affine subspace of~$\R^d$ that can be obtained as the intersection of some hyperplanes of~$\arrangement$.
625
The \defn{flat poset} of~$\arrangement$ is the poset~$\flatPoset$ of flats of~$\arrangement$ ordered by reverse inclusion.
626
\end{definition}
627
628
\begin{definition}
629
\label{def:MobiusPolynomial}
630
The \defn{M\"obius polynomial}~$\mobPol(x,y)$ of~$\arrangement$ is the polynomial defined by
631
\[
632
\mobPol \eqdef \sum_{F \le G} \mu_{\flatPoset}(F,G) \, x^{\dim(F)} \, y^{\dim(G)},
633
\]
634
where~$F \le G$ ranges over all intervals of the flat poset~$\flatPoset$, and~$\mu_{\flatPoset}(F,G)$ denotes the \defn{M\"obius function} on the flat poset~$\flatPoset$ defined as usual by
635
\[
636
\mu_{\flatPoset}(F, F) = 1
637
\qquad\text{and}\qquad
638
\sum_{F \le G \le H} \mu_{\flatPoset}(F,G) = 0
639
\]
640
for all~$F < H$ in~$\flatPoset$.
641
\end{definition}
642
643
\begin{remark}
644
Our definition of the M\"obius polynomial slightly differs from that of~\cite{Zaslavsky} as we use the dimension of~$F$ instead of its codimension, in order to simplify slightly the following statement.
645
\end{remark}
646
647
\begin{theorem}[{\cite[Thm.~A]{Zaslavsky}}]
648
\label{thm:Zaslavsky}
649
The $f$-polynomial, the $b$-polynomial, and the M\"obius polynomial of the hyperplane arrangement~$\arrangement$ are related by
650
\[
651
\fPol = \mobPol[\arrangement][-x][-1]
652
\qquad\text{and}\qquad
653
\bPol = \mobPol[\arrangement][-x][1].
654
\]
655
\end{theorem}
656
657
\begin{example}
658
%
659
\begin{figure}
660
% \begin{overpic}[scale=.9]{intersectionPoset}
661
% \put(72.5, -2){$1$}
662
% \put(51, 10){$-1$}
663
% \put(61, 10){$-1$}
664
% \put(71, 10){$-1$}
665
% \put(82, 10){$-1$}
666
% \put(94, 10){$-1$}
667
% \put(51.5, 32){$2$}
668
% \put(66, 32){$1$}
669
% \put(80, 32){$1$}
670
% \put(93.5, 32){$2$}
671
% \end{overpic}
672
% \caption{A hyperplane arrangement (left) and its intersection poset with its M\"obius function (right).}
673
\centerline{\includegraphics[scale=.9]{intersectionPoset}}
674
\caption{A hyperplane arrangement (left) and its intersection poset (right).}
675
\label{fig:arrangement}
676
\end{figure}
677
%
678
For the arrangement~$\arrangement$ of $5$ hyperplanes of \cref{fig:arrangement}, we have
679
\[
680
\mobPol = x^2y^2 - 5x^2y + 6x^2 + 5xy - 10x + 4 ,
681
\]
682
so that
683
\[
684
\fPol = \mobPol[\arrangement][-x][-1] = 12 \, x^2 + 15 \, x + 4
685
\qquad \text{and}\qquad
686
\bPol = \mobPol[\arrangement][-x][1] = 2 \, x^2 + 5 \, x + 4 .
687
\]
688
\end{example}
689
690
\begin{remark}
691
\label{rem:characteristicPolynomial}
692
The coefficient of~$x^d$ in the M\"obius polynomial~$\mobPol$ gives the more classical \defn{characteristic polynomial}
693
\[
694
\charPol \eqdef [x^d] \, \mobPol = \sum_F \mu_{\flatPoset}(\R^d,F) \, y^{\dim(F)} .
695
\]
696
By \cref{thm:Zaslavsky}, we thus have
697
\[
698
f_d(\arrangement) = (-1)^d \, \charPol[\arrangement][-1]
699
\qquad\text{and}\qquad
700
b_d(\arrangement) = (-1)^d \, \charPol[\arrangement][1].
701
\]
702
\end{remark}
703
704
%%%%%%%%%%%%%%%
705
706
\subsection{The braid arrangement}
707
\label{subsec:braidArrangement}
708
709
We now briefly recall the classical combinatorics of the braid arrangement.
710
See \cref{fig:facePosetBraidArrangement3,fig:intersectionPosetBraidArrangement3,fig:intersectionPosetBraidArrangement4} for illustrations when~$n = 3$ and~$n = 4$.
711
712
%\begin{figure}
713
% \centerline{\includegraphics[scale=.9]{figures/intersectionPosetBraidArrangement3Full}}
714
% \caption{The braid arrangement $\braidArrangement[3]$ (left), its flat poset~$\flatPoset[{\braidArrangement[3]}]$ (middle), and the partition poset~$\partitionPoset[3]$ (right).}
715
% \label{fig:braidArrangement3}
716
%\end{figure}
717
718
\afterpage{
719
\begin{figure}
720
\centerline{\includegraphics[scale=.6]{figures/facePosetBraidArrangement3}}
721
\caption{The face poset~$\facePoset[{\braidArrangement[3]}]$ of the braid arrangement $\braidArrangement[3]$ (left), where faces are represented as cones (middle) or as ordered set partitions of~$[3]$ (right).}
722
\label{fig:facePosetBraidArrangement3}
723
\end{figure}
724
}
725
726
\afterpage{
727
\begin{figure}
728
\centerline{\includegraphics[scale=.6]{figures/intersectionPosetBraidArrangement3}}
729
\caption{The flat poset~$\flatPoset[{\braidArrangement[3]}]$ of the braid arrangement $\braidArrangement[3]$, where flats are represented as intersections of hyperplanes (left) or as set partitions of~$[3]$ (right).}
730
\label{fig:intersectionPosetBraidArrangement3}
731
\end{figure}
732
}
733
734
\afterpage{
735
\begin{figure}
736
\centerline{\includegraphics[scale=.26]{figures/intersectionPosetBraidArrangement4}}
737
\caption{The flat poset~$\flatPoset[{\braidArrangement[4]}]$ of the braid arrangement $\braidArrangement[4]$, where flats are represented as intersections of hyperplanes (top) or as set partitions of~$[4]$ (bottom).}
738
\label{fig:intersectionPosetBraidArrangement4}
739
\end{figure}
740
}
741
742
\begin{definition}
743
Fix~$n \ge 1$ and denote by~$\HH$ the hyperplane of~$\R^n$ defined by~$\sum_{s \in [n]} x_s = 0$.
744
The \defn{braid arrangement}~$\braidArrangement$ is the arrangement of the hyperplanes~$\set{\b{x} \in \HH}{x_s = x_t}$ for all~${1 \le s < t \le n}$.
745
\end{definition}
746
747
\begin{remark}
748
\label{rem:essential}
749
Note that we have decided to work in the space~$\HH$ rather than in the space~$\R^n$.
750
The advantage is that the braid arrangement~$\braidArrangement$ in~$\HH$ is essential, so that we can speak of its rays.
751
Working in~$\R^n$ would change rays to walls, and would multiply all M\"obius polynomials by a factor~$xy$.
752
\end{remark}
753
754
The combinatorics of the braid arrangement~$\braidArrangement$ is well-known.
755
The descriptions of its face and flat posets involve both ordered and unordered set partitions.
756
To avoid confusions, we will always mark with an arrow the ordered structures (ordered set partitions, ordered partition forests, etc.).
757
Hence, the letter $\pi$ denotes an unordered set partition (the order is irrelevant, neither inside each part, nor between two distinct parts), while~$\order{\pi}$ denotes an ordered set partition (the order inside each part is irrelevant, but the order between distinct parts is relevant).
758
759
The braid arrangement~$\braidArrangement$ has a $k$-dimensional face
760
\[
761
\Phi(\order{\pi}) \eqdef \set{\b{x} \in \R^n}{x_s \le x_t \text{ for all $s,t$ such that the part of~$s$ is weakly before the part of~$t$ in $\order{\pi}$}}
762
\]
763
for each ordered set partition~$\order{\pi}$ of~$[n]$ into~$k+1$ parts, or equivalently, for each surjection from~$[n]$ to~$[k+1]$.
764
The face poset~$\facePoset[\braidArrangement]$ is thus isomorphic to the refinement poset~$\orderedPartitionPoset$ on ordered set partitions, where an ordered partition~$\order{\pi}$ is smaller than an ordered partition~$\order{\omega}$ if each part of~$\order{\pi}$ is the union of an interval of consecutive parts in~$\order{\omega}$.
765
In particular, it has a single vertex corresponding to the ordered partition~$[n]$, $2^n-2$ rays corresponding to the proper nonempty subsets of~$[n]$ (ordered partitions of~$[n]$ into~$2$ parts), and $n!$ regions corresponding to the permutations of~$[n]$ (ordered partitions of~$[n]$ into~$n$ parts).
766
As an example, \cref{fig:facePosetBraidArrangement3} illustrates the face poset of the braid arrangement~$\braidArrangement[3]$.
767
768
The braid arrangement~$\braidArrangement$ has a $k$-dimensional~flat
769
\[
770
\Psi(\pi) \eqdef \set{\b{x} \in \R^n}{x_s = x_t \text{ for all $s, t$ which belong to the same part of~$\pi$}}
771
\]
772
for each unordered set partition~$\pi$ of~$[n]$ into $k+1$ parts.
773
The flat poset~$\flatPoset[\braidArrangement]$ is thus isomorphic to the refinement poset~$\partitionPoset$ on set partitions of~$[n]$, where a partition~$\pi$ is smaller than a partition~$\omega$ if each part of~$\pi$ is contained in a part of~$\omega$.
774
For instance, \cref{fig:intersectionPosetBraidArrangement3,fig:intersectionPosetBraidArrangement4} illustrate the flat posets of the braid arrangements~$\braidArrangement[3]$ and~$\braidArrangement[4]$.
775
Note that the refinement in~$\orderedPartitionPoset$ and in~$\partitionPoset$ are in opposite direction.
776
777
The M\"obius function of the set partitions poset~$\partitionPoset$ is given by
778
\[
779
\mu_{\partitionPoset}(\pi, \omega) = \prod_{p \in \omega} (-1)^{\card{\pi[p]}-1}(\card{\pi[p]}-1)! \ ,
780
\]
781
where~$\pi[p]$ denotes the restriction of the partition~$\pi$ to the part~$p$ of the partition~$\omega$, and $\card{\pi[p]}$ denotes its number of parts.
782
See for instance~\cite{Birkhoff, Rota}.
783
The M\"obius polynomial of the braid arrangement~$\braidArrangement$ is given by
784
\[
785
\mobPol[\braidArrangement] = \sum_{k \in [n]} x^{k-1} S(n,k) \prod_{i \in [k-1]} (y-i) ,
786
\]
787
where~$S(n,k)$ denotes the Stirling number of the second kind \OEIS{A008277}, \ie the number of set partitions of~$[n]$ into~$k$ parts.
788
For instance
789
\begin{align*}
790
\mobPol[{\braidArrangement[1]}] & = 1 \\
791
\mobPol[{\braidArrangement[2]}] & = x y - x + 1 = x (y - 1) + 1 \\
792
\mobPol[{\braidArrangement[3]}] & = x^2 y^2 - 3 x^2 y + 2 x^2 + 3 x y - 3 x + 1 = x^2 (y - 1) (y - 2) + 3 x (y - 1) + 1\\
793
\mobPol[{\braidArrangement[4]}] & = x^3 y^3 - 6 x^3 y^2 + 11 x^3 y - 6 x^3 + 6 x^2 y^2 - 18 x^2 y + 12 x^2 + 7 x y - 7 x + 1 \\
794
& = x^3 (y - 1) (y - 2) (y - 3) + 6 x^2 (y - 1) (y - 2) + 7 x (y - 1) + 1.
795
\end{align*}
796
In particular, the characteristic polynomial of the braid arrangement~$\braidArrangement$ is given by
797
\[
798
\charPol[\braidArrangement] = (y-1) (y-2) \dots (y-n-1).
799
\]
800
Working in~$\R^n$ rather than in~$\HH$ would lead to an additional~$y$ factor in this formula, which might be more familiar to the reader.
801
See \cref{rem:essential}.
802
803
Finally, we will consider the evaluation of the M\"obius polynomial~$\mobPol[\braidArrangement]$ at~$y = 0$:
804
\[
805
\weirdPol[n] \eqdef \mobPol[\braidArrangement][x][0] = \sum_{k \in [n]} (-1)^{k-1} \, (k-1)! \, S(n,k) \, x^{k-1}.
806
\]
807
The coefficients of this polynomial are given by the sequence \OEIS{A028246}.
808
We just observe here that it is connected to the $\b{f}$-polynomial of~$\braidArrangement$.
809
810
\begin{lemma}
811
We have~$\weirdPol[n] = (1-x) \, \fPol[\braidArrangement]$.
812
\end{lemma}
813
814
\begin{proof}
815
This lemma is equivalent to the equality
816
\begin{equation*}
817
\sum_{k=1}^n (-1)^{k-1} (k-1)! \, S(n,k) \, x^{k-1} = (1-x) \sum_{k=1}^{n-1} (-1)^{k-1} \, k! \, S(n-1,k) \, x^{k-1}.
818
\end{equation*}
819
Distributing $(1-x)$ in the right hand side gives:
820
\begin{gather*}
821
(1-x) \sum_{k=1}^{n-1} (-1)^{k-1} \, k! \, S(n-1,k) \, x^{k-1} \\
822
= \sum_{k=1}^{n-1} k! \, S(n-1,k) \, (-x)^{k-1} + \sum_{k=1}^{n-1} k! \, S(n-1,k) \, (-x)^{k} \\
823
= \sum_{k=1}^{n-1} k! \, S(n-1,k) \, (-x)^{k-1} + \sum_{k=2}^{n} (k-1)! \, S(n-1,k-1) \, (-x)^{k-1} + (n-1)! \, S(n-1, n-1) \, (-x)^{n-1} \\
824
= S(n-1,1) \, (-x)^0 + \sum_{k=2}^{n-1} (k-1)! \, \big( S(n-1,k-1) + k \, S(n-1,k) \big) \, (-x)^{k-1}.
825
\end{gather*}
826
The result thus follows from the inductive formula on Stirling numbers of the second kind
827
\[
828
S(n+1,k) = k \, S(n,k) + S(n,k-1)
829
\]
830
for $0<k<n$.
831
\end{proof}
832
833
%%%%%%%%%%%%%%%
834
835
\subsection{The $(\ell,n)$-braid arrrangement}
836
\label{subsec:multiBraidArrangement}
837
838
\enlargethispage{.7cm}
839
We now focus on the following specific hyperplane arrangements, illustrated in \cref{fig:multiBraidArrangements}.
840
We still denote by~$\HH$ the hyperplane of~$\R^n$ defined by~$\sum_{s \in [n]} x_s = 0$.
841
842
\begin{figure}[t]
843
\centerline{
844
\begin{tabular}{c@{\hspace{.7cm}}c@{\hspace{.7cm}}c@{\hspace{.7cm}}c}
845
\includegraphics[scale=.4]{multiBraidArrangement1}
846
&
847
\includegraphics[scale=.4]{multiBraidArrangement2}
848
&
849
\includegraphics[scale=.4]{multiBraidArrangement3}
850
&
851
\includegraphics[scale=.4]{multiBraidArrangement4}
852
\\
853
$\ell = 1$ & $\ell = 2$ & $\ell = 3$ & $\ell = 4$
854
\end{tabular}
855
}
856
\caption{The $(\ell,3)$-braid arrangements for~$\ell \in [4]$.}
857
\label{fig:multiBraidArrangements}
858
\end{figure}
859
860
%\begin{definition}
861
%\label{def:multiBraidArrangement}
862
%The \defn{$(\ell,n)$-braid arrangement}~$\multiBraidArrangement$ is the arrangement in~$\HH$ obtained as the union of $\ell$ generically translated copies of the braid arrangement~$\braidArrangement$ (that is, the $\b{a}$-braid arrangement for some generic matrix~$\b{a} \in M_{\ell,n-1}(\R)$).
863
%\end{definition}
864
%
865
%This definition is slightly misleading, as the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ a priori depends on the (generic) translation vectors chosen for each copy.
866
%We will see that many combinatorial aspects of~$\multiBraidArrangement$, in particular its flat poset and thus its M\"obius, $f$- and $b$-polynomials, are in fact independent of the translation vectors as long as they are generic.
867
%However, the combinatorial description of the face poset of~$\multiBraidArrangement$ will depend on the translation vectors, so that we introduce the following more precise notation recording the translations.
868
%
869
%\begin{definition}
870
%A matrix~$\b{a} \eqdef (a_{i,j}) \in M_{\ell,n-1}(\R)$ is \defn{generic} if any linear dependence among its coefficients~$\sum_{i,j} \lambda_{i,j} \, a_{i,j} = 0$ splits into $\ell$ linear dependences~$\sum_j \lambda_{i,j} \, a_{i,j} = 0$ for~$i \in [\ell]$.
871
%\end{definition}
872
%
873
%\begin{definition}
874
%\label{def:multiBraidArrangementPrecise}
875
%For any integers~$\ell,n \geq 1$, and any matrix~$\b{a} \eqdef (a_{i,j}) \in M_{\ell,n-1}(\R)$, the \defn{$\b{a}$-braid arrangement}~$\multiBraidArrangement(\b{a})$ is the arrangement of hyperplanes~$\set{\b{x} \in \HH}{x_s - x_t = A_{i,s,t}}$ for all~${1 \le s < t \le n}$ and~$i \in [\ell]$, where $A_{i,s,t} \eqdef \smash{\sum_{s \le j < t} a_{i,j}}$.
876
%\end{definition}
877
878
\begin{definition}
879
\label{def:multiBraidArrangementPrecise}
880
For any integers~$\ell,n \geq 1$, and any matrix~$\b{a} \eqdef (a_{i,j}) \in M_{\ell,n-1}(\R)$, the \defn{$\b{a}$-braid arrangement}~$\multiBraidArrangement(\b{a})$ is the arrangement of hyperplanes~$\set{\b{x} \in \HH}{x_s - x_t = A_{i,s,t}}$ for all~${1 \le s < t \le n}$ and~$i \in [\ell]$, where $A_{i,s,t} \eqdef \smash{\sum_{s \le j < t} a_{i,j}}$.
881
\end{definition}
882
883
In other words, the $\b{a}$-braid arrangement~$\multiBraidArrangement(\b{a})$ is the union of~$\ell$ copies of the braid arrangement~$\braidArrangement$ translated according to the matrix~$\b{a}$.
884
Of course, the $\b{a}$-braid arrangement~$\multiBraidArrangement(\b{a})$ highly depends on~$\b{a}$.
885
In this paper, we are interested in the case where~$\b{a}$ is generic in the following sense.
886
887
\begin{definition}
888
A matrix~$\b{a} \eqdef (a_{i,j}) \in M_{\ell,n-1}(\R)$ is \defn{generic} if for any~$i_1, \dots, i_k \in [\ell]$ and distinct $r_1, \dots, r_k \in [n]$, the equality~$\sum_{j \in [k]} A_{i_j, r_{j-1}, r_j} = 0$ implies~$i_1 = \dots = i_k$ (with the notation~$A_{i,s,t} \eqdef \smash{\sum_{s \le j < t} a_{i,j}}$ and the convention~$r_0 = r_k$).
889
\end{definition}
890
891
We will see that many combinatorial aspects of~$\multiBraidArrangement(\b{a})$, in particular its flat poset and thus its M\"obius, $f$- and $b$-polynomials, are in fact independent of the matrix~$\b{a}$ as long as it is generic.
892
We therefore consider the following definition.
893
894
\begin{definition}
895
\label{def:multiBraidArrangement}
896
The \defn{$(\ell,n)$-braid arrangement}~$\multiBraidArrangement$ is the arrangement in~$\HH$ obtained as the union of $\ell$ generically translated copies of the braid arrangement~$\braidArrangement$ (that is, any $\b{a}$-braid arrangement for some generic matrix~$\b{a} \in M_{\ell,n-1}(\R)$).
897
\end{definition}
898
899
%\begin{remark}
900
%\label{rem:multiBraidArrangement}
901
%In practice, consider the hyperplanes~$\set{\b{x} \in \HH}{x_s - x_t = A_{i,s,t}}$ for all~$1 \le s < t \le n$ and~$i \in [\ell]$, where $A_{i,s,t} \eqdef \smash{\sum_{s \le j < t} a_{i,j}}$ for an arbitrary generic matrix~$(a_{i,j}) \in M_{\ell,n-1}(\R)$.
902
%\end{remark}
903
904
The objective of \cref{part:multiBraidArrangements} is to explore the combinatorics of these multiple braid arrangements.
905
We have split our presentation into two sections:
906
\begin{itemize}
907
\item In \cref{sec:flatPoset}, we describe the flat poset~$\flatPoset[\multiBraidArrangement]$ of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ in terms of $(\ell,n)$-partition forests (\cref{subsec:partitionForests}) and labeled $(\ell,n)$-rainbow forests (\cref{subsec:rainbowForests}), which enables us to derive its M\"obius, $f$- and $b$- polynomials (\cref{subsec:MobiusPolynomialMultiBraidArrangement}), from which we extract interesting formulas for the number of vertices (\cref{subsec:verticesMultiBraidArrangement}) and regions (\cref{subsec:regionsMultiBraidArrangement}). Note that all these results are independent of the translation matrix.
908
\item In \cref{sec:facePoset}, we describe the face poset~$\facePoset[\multiBraidArrangement(\b{a})]$ of the $\b{a}$-braid arrangement~$\multiBraidArrangement(\b{a})$ in terms of ordered $(\ell,n)$-partition forests (\cref{subsec:orderedPartitionForests}). In contrast to the flat poset, this description of the face poset depends on the translation matrix~$\b{a}$. For a given choice of~$\b{a}$, we describe in particular the ordered $(\ell,n)$-partitions forests with a given underlying (unordered) $(\ell,n)$-partition forest (\cref{subsec:PFtoOPF}). We then give a criterion to decide whether a given ordered $(\ell,n)$-partition forest corresponds to a face of~$\multiBraidArrangement(\b{a})$ (\cref{subsec:criterionOPF}).
909
\end{itemize}
910
911
\begin{remark}
912
Note that each hyperplane of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ is orthogonal to a root~$\b{e}_i-\b{e}_j$ of the type~$A$ root system.
913
Many such arrangements have been studied previously, for instance, the \defn{Shi arrangement}~\cite{Shi1, Shi2}, the \defn{Catalan arrangement}~\cite[Sect.~7]{PostnikovStanley}, the \defn{Linial arrangement}~\cite[Sect.~8]{PostnikovStanley}, the \defn{generic arrangement} of~\cite[Sect.~5]{PostnikovStanley}, or the \defn{discriminantal arrangements} of~\cite{ManinSchechtman,BayerBrandt}.
914
We refer to the work of A.~Postnikov and R.~Stanley~\cite{PostnikovStanley} and of O.~Bernardi~\cite{Bernardi} for much more references.
915
However, in all these examples, either the copies of the braid arrangement are perturbed, or they are translated non-generically.
916
We have not been able to find the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ properly treated in the literature.
917
\end{remark}
918
919
\begin{remark}
920
Part of our discussion on the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ could actually be developed for a hyperplane arrangement~$\arrangement^\ell$ obtained as the union of~$\ell$ generically translated copies of an arbitrary linear hyperplane arrangement~$\arrangement$.
921
Similarly to \cref{prop:flatPosetMultiBraidArrangement}, the flat poset~$\flatPoset[\arrangement^\ell]$ is isomorphic to the lower set of the $\ell$\ordinal{} Cartesian power of the flat poset~$\flatPoset$ induced by the $\ell$-tuples whose meet in the flat poset~$\flatPoset$ is the bottom element~$\b{0}$ (these are sometimes called strong antichains) and which are minimal for this property.
922
Similar to \cref{thm:MobiusPolynomialMultiBraidArrangement}, this yields a general formula for the M\"obius polynomial of~$\arrangement^\ell$ in terms of the M\"obius function of the flat poset~$\flatPoset$.
923
Here, we additionally benefit from the nice properties of the M\"obius polynomial of the braid arrangement~$\braidArrangement$ to obtain appealing formulas for the vertices, regions and bounded regions of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ (see \cref{thm:verticesMultiBraidArrangement,thm:verticesRefinedMultiBraidArrangement,thm:characteristicPolynomialMultiBraidArrangement,thm:regionsMultiBraidArrangement}).
924
We have therefore decided to restrict our attention to the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$.
925
\end{remark}
926
927
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
928
929
\section{Flat poset and enumeration of~$\multiBraidArrangement$}
930
\label{sec:flatPoset}
931
932
In this section, we describe the flat poset of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ in terms of \mbox{$(\ell,n)$-partition} forests and derive explicit formulas for its $f$-vector.
933
Remarkably, the flat poset (and thus the M\"obius, $f$- and $b$- polynomials) of~$\multiBraidArrangement$ is independent of the translation vectors as long as they are generic.
934
935
%%%%%%%%%%%%%%%
936
937
\subsection{Partition forests}
938
\label{subsec:partitionForests}
939
940
We first introduce the main characters of this section, which will describe the combinatorics of the flat poset of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ of \cref{def:multiBraidArrangement}.
941
942
\begin{definition}
943
\label{def:intersectionHypergraph}
944
The \defn{intersection hypergraph} of a $\ell$-tuple~$\b{F} \eqdef (F_1, \dots, F_\ell)$ of set partitions of~$[n]$ is the $\ell$-regular $\ell$-partite hypergraph on all parts of all the partitions~$F_i$ for~${i \in [\ell]}$, with a hyperedge connecting the parts containing~$j$ for each~$j \in [n]$.
945
\end{definition}
946
947
\begin{definition}
948
\label{def:partitionForests}
949
An \defn{$(\ell,n)$-partition forest} (\resp \defn{tree}) is a $\ell$-tuple~$\b{F} \eqdef (F_1, \dots, F_\ell)$ of set partitions of~$[n]$ whose intersection hypergraph is a hyperforest (\resp hypertree).
950
See \cref{fig:forests}.
951
The \defn{dimension} of~$\b{F}$ is~$\smash{{\dim(\b{F}) \eqdef n - 1 - \ell n + \sum_{i \in [\ell]} \card{F_i}}}$.
952
The \defn{$(\ell,n)$-partition forest poset} is the poset~$\forestPoset$ on $(\ell,n)$-partition forests ordered by componentwise refinement.
953
%
954
\begin{figure}[b]
955
\centerline{\includegraphics[scale=.9]{forests}}
956
\caption{Some $(3,6)$-partition forests (top) with their intersection hypergraphs (middle) and the corresponding labeled $(3,6)$-rainbow forests (bottom). The last two are trees. The order of the colors in the bottom pictures is red, green, blue.}
957
\label{fig:forests}
958
\end{figure}
959
\end{definition}
960
961
In other words, $\forestPoset$ is the lower set of the $\ell$\ordinal{} Cartesian power of the partition poset~$\partitionPoset$ induced by $(\ell,n)$-partition forests.
962
Note that the maximal elements of~$\forestPoset$ are the $(\ell, n)$-partition trees.
963
964
The following statement is illustrated in \cref{fig:intersectionPosetMultiBraidArrangement32}.
965
%
966
\begin{figure}
967
\centerline{\includegraphics[scale=.9]{figures/intersectionPosetMultiBraidArrangement32Full}}
968
\caption{The $(2,3)$-braid arrangement $\multiBraidArrangement[3][2]$ (left), and its flat poset (right), where flats are represented as intersections of hyperplanes (top), as $(2,3)$-partitions forests (middle), and as labeled $(2,3)$-rainbow forests (bottom).}
969
\label{fig:intersectionPosetMultiBraidArrangement32}
970
\end{figure}
971
972
\begin{proposition}
973
\label{prop:flatPosetMultiBraidArrangement}
974
The flat poset~$\flatPoset[\multiBraidArrangement]$ of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ is isomorphic to the $(\ell,n)$-partition forest poset.
975
\end{proposition}
976
977
\begin{proof}
978
Consider that~$\multiBraidArrangement$ is the $\b{a}$-braid arrangement~$\multiBraidArrangement(\b{a})$ for some generic matrix~$\b{a}$.
979
%Consider the hyperplanes~$\set{\b{x} \in \HH}{x_s - x_t = A_{i,s,t}}$ described in \cref{def:multiBraidArrangementPrecise}.
980
In view of our discussion in \cref{subsec:braidArrangement}, observe that, for each~${i \in [\ell]}$, each set partition~$\pi$ of~$[n]$ corresponds to a $(\card{\pi})$-dimensional flat
981
\[
982
\Psi_i(\pi) \eqdef \set{\b{x} \in \HH}{x_s - x_t = A_{i,s,t} \text{ for all $s,t$ in the same part of $\pi$} }
983
\]
984
of the $i$\ordinal{} copy of the braid arrangement~$\braidArrangement$.
985
The flats of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ are thus all of the form
986
\[
987
\Psi(\b{F}) \eqdef \bigcap_{i \in [\ell]} \Psi_i(F_i)
988
\]
989
for certain $\ell$-tuples~$\b{F} \eqdef (F_1, \dots, F_\ell)$ of set partitions of~$[n]$.
990
Since the matrix~$\b{a}$ is generic, $\Psi(\b{F})$ is non-empty if and only if the intersection hypergraph of~$\b{F}$ is acyclic.
991
Moreover, $\Psi(\b{F})$ is included in~$\Psi(\b{G})$ if and only if~$\b{F}$ refines~$\b{G}$ componentwise.
992
Hence, the flat poset of~$\multiBraidArrangement$ is isomorphic to the $(\ell,n)$-partition forest poset.
993
Finally, notice that the codimension of the flat~$\Psi(\b{F})$ is the sum of the codimensions of the flats~$\Psi_i(F_i)$ for~$i \in [\ell]$, so that~$\dim(\b{F}) \eqdef n - 1 - \ell n + \sum_{i \in [\ell]} \card{F_i} $ is indeed the dimension of the flat~$\Psi(\b{F})$.
994
\end{proof}
995
996
%%%%%%%%%%%%%%%
997
998
\subsection{M\"obius polynomial}
999
\label{subsec:MobiusPolynomialMultiBraidArrangement}
1000
1001
We now derive from \cref{def:MobiusPolynomial,prop:flatPosetMultiBraidArrangement} the M\"obius polynomial of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$.
1002
1003
\begin{theorem}
1004
\label{thm:MobiusPolynomialMultiBraidArrangement}
1005
The M\"obius polynomial of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ is given by
1006
\[
1007
\mobPol[\multiBraidArrangement] = x^{n-1-\ell n} y^{n-1-\ell n} \sum_{\b{F} \le \b{G}} \prod_{i \in [\ell]} x^{\card{F_i}} y^{\card{G_i}} \prod_{p \in G_i} (-1)^{\card{F_i[p]}-1} (\card{F_i[p]}-1)! \; ,
1008
\]
1009
where~$\b{F} \le \b{G}$ ranges over all intervals of the $(\ell,n)$-partition forest poset~$\forestPoset$, and~$F_i[p]$ denotes the restriction of the partition~$F_i$ to the part~$p$ of~$G_i$.
1010
\end{theorem}
1011
1012
\begin{proof}
1013
Observe that for~$\b{F} \eqdef (F_1, \dots, F_\ell)$ and~$\b{G} \eqdef (G_1, \dots, G_\ell)$ in~$\forestPoset$, we have
1014
\[
1015
[\b{F}, \b{G}] = \prod_{i \in [\ell]} [F_i, G_i] \simeq \prod_{i \in [\ell]} \prod_{p \in G_i} \partitionPoset[{\card{F_i[p]}}].
1016
\]
1017
Recall that the M\"obius function is multiplicative:
1018
\(
1019
\mu_{P \times Q} \big( (p,q), (p’,q’) \big) = \mu_P(p,p’) \cdot \mu_Q(q,q’),
1020
\)
1021
for all~$p, p' \in P$ and~$q, q' \in Q$.
1022
Hence, we obtain that
1023
\[
1024
\mu_{\forestPoset}(\b{F}, \b{G}) = \prod_{i \in [\ell]} \prod_{p \in G_i} (-1)^{\card{F_i[p]}-1} (\card{F_i[p]}-1)! .
1025
\]
1026
Hence, we derive from \cref{def:MobiusPolynomial,prop:flatPosetMultiBraidArrangement} that
1027
\begin{align*}
1028
\mobPol[\multiBraidArrangement]
1029
& = \sum_{\b{F} \le \b{G}} \mu_{\forestPoset}(\b{F}, \b{G}) \, x^{\dim(\b{F})} \, y^{\dim(\b{G})} \\
1030
& = x^{n-1-\ell n} y^{n-1-\ell n} \sum_{\b{F} \le \b{G}} \prod_{i \in [\ell]} x^{\card{F_i}} y^{\card{G_i}} \prod_{p \in G_i} (-1)^{\card{F_i[p]}-1} (\card{F_i[p]}-1)! .
1031
\qedhere
1032
\end{align*}
1033
\end{proof}
1034
1035
By using the polynomial
1036
\[
1037
\weirdPol[n] \eqdef \mobPol[\braidArrangement][x][0] = \sum_{k \in [n]} (-1)^{k-1} \, (k-1)! \, S(n,k) \, x^{k-1}
1038
\]
1039
introduced at the end of \cref{subsec:braidArrangement}, the M\"obius polynomial~$\mobPol[\multiBraidArrangement]$ can also be expressed as follows.
1040
1041
\pagebreak
1042
\begin{proposition}
1043
\label{prop:alternativeFormulaMobiusPolynomialMultiBraidArrangement}
1044
The M\"obius polynomial of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ is given by
1045
\[
1046
\mobPol[\multiBraidArrangement] = x^{(n-1)(1-\ell)} \sum_{G \in \forestPoset} y^{n-1-\ell n+\sum_{i \in [\ell]} \card{G_i}} \prod_{i \in [\ell]} \weirdPol[\card{G_i}].
1047
\]
1048
\end{proposition}
1049
1050
\begin{proof}
1051
As already mentioned, the $(\ell,n)$-partition forest poset~$\forestPoset$ is a lower set of the $\ell$\ordinal{} Cartesian power of the partition poset~$\partitionPoset$.
1052
In other words, given a $(\ell,n)$-partition forest $\b{G} \eqdef (G_1, \dots, G_\ell)$, any $\ell$-tuple~$\b{F} \eqdef (F_1, \dots, F_\ell)$ of partitions satisfying~$F_i \le_{\partitionPoset[n]} G_i$ for all~$i \in [\ell]$ is a $(\ell,n)$-partition forest.
1053
Hence, we obtain from \cref{def:MobiusPolynomial,prop:flatPosetMultiBraidArrangement} that
1054
\begin{align*}
1055
\mobPol[\multiBraidArrangement]
1056
& = \sum_{G \in \forestPoset} y^{n-\ell n - 1 + \sum_{i \in [\ell]} \card{G_i}} \prod_{i \in [\ell]} \sum_{F_i \leq_{\partitionPoset[n]} G_i } \mu_{\partitionPoset[n]}(F_i,G_i) \, x^{n-\ell n - 1 + \sum_{i \in [\ell]} \card{F_i}}, \\
1057
& = \sum_{G \in \forestPoset} y^{n-\ell n - 1 + \sum_{i \in [\ell]} \card{G_i}} x^{(n-1)(1-\ell)} \prod_{i \in [\ell]} \sum_{\pi_i \in \partitionPoset[\card{G_i}]} \mu_{\partitionPoset[\card{G_i}]}(\pi_i,\hat{1}) \, x^{\card{\pi_i}-1},
1058
\end{align*}
1059
where $\hat{1}$ denotes the maximal element in $\partitionPoset[\card{G_i}]$ and $\pi_i$ is obtained from $F_i$ by merging elements in the same part of $G_i$.
1060
The result follows since~$\weirdPol[\card{G_i}] = \sum_{\pi_i \in \partitionPoset[\card{G_i}]} \mu_{\partitionPoset[\card{G_i}]}(\pi_i,\hat{1}) \, x^{\card{\pi_i}-1}$.
1061
%Hence, we obtain
1062
%\[
1063
%\mobPol[\multiBraidArrangement] = \sum_{G \in \forestPoset} y^{n-\ell n - 1 + \sum_{i \in [\ell]} \card{G_i}} x^{(n-1)(1-\ell)} \prod_{i \in [\ell]} \weirdPol[\card{G_i}].
1064
%\qedhere
1065
%\]
1066
\end{proof}
1067
1068
From \cref{thm:Zaslavsky,thm:MobiusPolynomialMultiBraidArrangement}, we thus obtain the face numbers and bounded face numbers of~$\multiBraidArrangement$, whose first few values are gathered in \cref{table:fvectorMultiBraidArrangements}.
1069
1070
%\begin{table}
1071
%\centerline{
1072
% \begin{tabular}{c@{\hspace{.7cm}}c}
1073
% \begin{tabular}[t]{c|cccc|c}
1074
% $n \backslash k$ & $0$ & $1$ & $2$ & $3$ & $\Sigma$ \\
1075
% \hline
1076
% $1$ & $1$ &&&& $1$ \\
1077
% $2$ & $2$ & $1$ &&& $3$ \\
1078
% $3$ & $6$ & $6$ & $1$ && $13$ \\
1079
% $4$ & $24$ & $36$ & $14$ & $1$ & $75$
1080
% \end{tabular}
1081
% &
1082
% \begin{tabular}[t]{c|cccc|c}
1083
% $n \backslash k$ & $0$ & $1$ & $2$ & $3$ & $\Sigma$ \\
1084
% \hline
1085
% $1$ & $1$ &&&& $1$ \\
1086
% $2$ & $3$ & $2$ &&& $5$ \\
1087
% $3$ & $17$ & $24$ & $8$ && $49$ \\
1088
% $4$ & $149$ & $324$ & $226$ & $50$ & $749$
1089
% \end{tabular}
1090
% \\[2cm]
1091
% \begin{tabular}[t]{c|cccc|c}
1092
% $n \backslash k$ & $0$ & $1$ & $2$ & $3$ & $\Sigma$ \\
1093
% \hline
1094
% $1$ & $1$ &&&& $1$ \\
1095
% $2$ & $0$ & $1$ &&& $1$ \\
1096
% $3$ & $0$ & $0$ & $1$ && $1$ \\
1097
% $4$ & $0$ & $0$ & $0$ & $1$ & $1$
1098
% \end{tabular}
1099
% &
1100
% \begin{tabular}[t]{c|cccc|c}
1101
% $n \backslash k$ & $0$ & $1$ & $2$ & $3$ & $\Sigma$ \\
1102
% \hline
1103
% $1$ & $1$ &&&& $1$ \\
1104
% $2$ & $1$ & $2$ &&& $3$ \\
1105
% $3$ & $5$ & $12$ & $8$ && $25$ \\
1106
% $4$ & $43$ & $132$ & $138$ & $50$ & $363$
1107
% \end{tabular}
1108
% \\[2cm]
1109
% $\ell = 1$ & $\ell = 2$
1110
% \\[.8cm]
1111
% \begin{tabular}[t]{c|cccc|c}
1112
% $n \backslash k$ & $0$ & $1$ & $2$ & $3$ & $\Sigma$ \\
1113
% \hline
1114
% $1$ & $1$ &&&& $1$ \\
1115
% $2$ & $4$ & $3$ &&& $7$ \\
1116
% $3$ & $34$ & $54$ & $21$ && $109$ \\
1117
% $4$ & $472$ & $1152$ & $924$ & $243$ & $2791$
1118
% \end{tabular}
1119
% &
1120
% \begin{tabular}[t]{c|cccc|c}
1121
% $n \backslash k$ & $0$ & $1$ & $2$ & $3$ & $\Sigma$ \\
1122
% \hline
1123
% $1$ & $1$ &&&& $1$ \\
1124
% $2$ & $5$ & $4$ &&& $9$ \\
1125
% $3$ & $57$ & $96$ & $40$ && $193$ \\
1126
% $4$ & $1089$ & $2808$ & $2396$ & $676$ & $6969$
1127
% \end{tabular}
1128
% \\[2cm]
1129
% \begin{tabular}[t]{c|cccc|c}
1130
% $n \backslash k$ & $0$ & $1$ & $2$ & $3$ & $\Sigma$ \\
1131
% \hline
1132
% $1$ & $1$ &&&& $1$ \\
1133
% $2$ & $2$ & $3$ &&& $5$ \\
1134
% $3$ & $16$ & $36$ & $21$ && $73$ \\
1135
% $4$ & $224$ & $684$ & $702$ & $243$ & $1853$
1136
% \end{tabular}
1137
% &
1138
% \begin{tabular}[t]{c|cccc|c}
1139
% $n \backslash k$ & $0$ & $1$ & $2$ & $3$ & $\Sigma$ \\
1140
% \hline
1141
% $1$ & $1$ &&&& $1$ \\
1142
% $2$ & $3$ & $4$ &&& $7$ \\
1143
% $3$ & $33$ & $72$ & $40$ && $145$ \\
1144
% $4$ & $639$ & $1944$ & $1980$ & $676$ & $5239$
1145
% \end{tabular}
1146
% \\[2cm]
1147
% $\ell = 3$ & $\ell = 4$
1148
% \end{tabular}
1149
% }
1150
% \vspace{.3cm}
1151
% \caption{The face numbers (top) and the bounded face numbers (bottom) of the $(\ell,n)$-braid arrangements for~$\ell, n \in [4]$.}
1152
% \label{table:fvectorMultiBraidArrangements}
1153
%\end{table}
1154
1155
\begin{table}[t]
1156
\centerline{\scalebox{.8}{
1157
\begin{tabular}{c@{\hspace{.7cm}}c@{\hspace{.7cm}}c@{\hspace{.7cm}}c}
1158
$\ell = 1$ & $\ell = 2$ & $\ell = 3$ & $\ell = 4$
1159
\\[.2cm]
1160
\begin{tabular}[t]{c|cccc|c}
1161
$n \backslash k$ & $0$ & $1$ & $2$ & $3$ & $\Sigma$ \\
1162
\hline
1163
$1$ & $1$ &&&& $1$ \\
1164
$2$ & $2$ & $1$ &&& $3$ \\
1165
$3$ & $6$ & $6$ & $1$ && $13$ \\
1166
$4$ & $24$ & $36$ & $14$ & $1$ & $75$
1167
\end{tabular}
1168
&
1169
\begin{tabular}[t]{c|cccc|c}
1170
$n \backslash k$ & $0$ & $1$ & $2$ & $3$ & $\Sigma$ \\
1171
\hline
1172
$1$ & $1$ &&&& $1$ \\
1173
$2$ & $3$ & $2$ &&& $5$ \\
1174
$3$ & $17$ & $24$ & $8$ && $49$ \\
1175
$4$ & $149$ & $324$ & $226$ & $50$ & $749$
1176
\end{tabular}
1177
&
1178
\begin{tabular}[t]{c|cccc|c}
1179
$n \backslash k$ & $0$ & $1$ & $2$ & $3$ & $\Sigma$ \\
1180
\hline
1181
$1$ & $1$ &&&& $1$ \\
1182
$2$ & $4$ & $3$ &&& $7$ \\
1183
$3$ & $34$ & $54$ & $21$ && $109$ \\
1184
$4$ & $472$ & $1152$ & $924$ & $243$ & $2791$
1185
\end{tabular}
1186
&
1187
\begin{tabular}[t]{c|cccc|c}
1188
$n \backslash k$ & $0$ & $1$ & $2$ & $3$ & $\Sigma$ \\
1189
\hline
1190
$1$ & $1$ &&&& $1$ \\
1191
$2$ & $5$ & $4$ &&& $9$ \\
1192
$3$ & $57$ & $96$ & $40$ && $193$ \\
1193
$4$ & $1089$ & $2808$ & $2396$ & $676$ & $6969$
1194
\end{tabular}
1195
\\[2.3cm]
1196
\begin{tabular}[t]{c|cccc|c}
1197
$n \backslash k$ & $0$ & $1$ & $2$ & $3$ & $\Sigma$ \\
1198
\hline
1199
$1$ & $1$ &&&& $1$ \\
1200
$2$ & $0$ & $1$ &&& $1$ \\
1201
$3$ & $0$ & $0$ & $1$ && $1$ \\
1202
$4$ & $0$ & $0$ & $0$ & $1$ & $1$
1203
\end{tabular}
1204
&
1205
\begin{tabular}[t]{c|cccc|c}
1206
$n \backslash k$ & $0$ & $1$ & $2$ & $3$ & $\Sigma$ \\
1207
\hline
1208
$1$ & $1$ &&&& $1$ \\
1209
$2$ & $1$ & $2$ &&& $3$ \\
1210
$3$ & $5$ & $12$ & $8$ && $25$ \\
1211
$4$ & $43$ & $132$ & $138$ & $50$ & $363$
1212
\end{tabular}
1213
&
1214
\begin{tabular}[t]{c|cccc|c}
1215
$n \backslash k$ & $0$ & $1$ & $2$ & $3$ & $\Sigma$ \\
1216
\hline
1217
$1$ & $1$ &&&& $1$ \\
1218
$2$ & $2$ & $3$ &&& $5$ \\
1219
$3$ & $16$ & $36$ & $21$ && $73$ \\
1220
$4$ & $224$ & $684$ & $702$ & $243$ & $1853$
1221
\end{tabular}
1222
&
1223
\begin{tabular}[t]{c|cccc|c}
1224
$n \backslash k$ & $0$ & $1$ & $2$ & $3$ & $\Sigma$ \\
1225
\hline
1226
$1$ & $1$ &&&& $1$ \\
1227
$2$ & $3$ & $4$ &&& $7$ \\
1228
$3$ & $33$ & $72$ & $40$ && $145$ \\
1229
$4$ & $639$ & $1944$ & $1980$ & $676$ & $5239$
1230
\end{tabular}
1231
\end{tabular}
1232
}}
1233
% \vspace{.3cm}
1234
\caption{The face numbers (top) and the bounded face numbers (bottom) of the $(\ell,n)$-braid arrangements for~$\ell, n \in [4]$.}
1235
\label{table:fvectorMultiBraidArrangements}
1236
\end{table}
1237
1238
\begin{corollary}
1239
\label{coro:fbvectorsMultiBraidArrangement}
1240
The $f$- and $b$-polynomials of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ are given by
1241
\begin{align*}
1242
\fPol[\multiBraidArrangement] & = x^{n-1-\ell n}\sum_{\b{F} \le \b{G}} \prod_{i \in [\ell]} x^{\card{F_i}} \prod_{p \in G_i} (\card{F_i[p]}-1)!\\
1243
\text{and}\qquad
1244
\bPol[\multiBraidArrangement] & = (-1)^\ell x^{n-1-\ell n} \sum_{\b{F} \le \b{G}} \prod_{i \in [\ell]} x^{\card{F_i}} \prod_{p \in G_i} -(\card{F_i[p]}-1)! ,
1245
\end{align*}
1246
where~$\b{F} \le \b{G}$ ranges over all intervals of the $(\ell,n)$-partition forest poset~$\forestPoset$, and~$F_i[p]$ denotes the restriction of the partition~$F_i$ to the part~$p$ of~$G_i$.
1247
\end{corollary}
1248
1249
\begin{example}
1250
For~$n = 1$, we have
1251
\[
1252
\mobPol[{\multiBraidArrangement[1][\ell]}] = \fPol[{\multiBraidArrangement[1][\ell]}] = \bPol[{\multiBraidArrangement[1][\ell]}] = 1.
1253
\]
1254
For~$n = 2$, we have
1255
\[
1256
\mobPol[{\multiBraidArrangement[2][\ell]}] = xy-\ell x+\ell,
1257
\quad
1258
\fPol[{\multiBraidArrangement[2][\ell]}] = (\ell+1)x+\ell
1259
\quad\text{and}\quad
1260
\bPol[{\multiBraidArrangement[2][\ell]}] = (\ell-1)x+\ell.
1261
\]
1262
The case~$n = 3$ is already more interesting.
1263
Consider the set partitions~$P \eqdef \big\{ \{1\}, \{2\}, \{3\} \big\}$, $Q_i \eqdef \big\{ \{i\}, [3] \ssm \{i\} \big\}$ for~$i \in [3]$, and~$R \eqdef \big\{ [3] \big\}$.
1264
Observe that the $(\ell,3)$-partition forests are all of the form
1265
\begin{gather*}
1266
\b{F} \eqdef P^\ell,
1267
\quad
1268
\b{G}_i^p \eqdef P^p Q_i P^{\ell-p-1}, % \text{ for } p \le \ell-1 \text{ and } i \in [3],
1269
\quad
1270
\b{H}_{i,j}^{p,q} \eqdef P^p Q_i P^{\ell-p-q-2} Q_j P^q \;\text{($i \ne j$)} % \text{ for } p + q \le \ell-2 \text{ and } i \ne j \in [3]
1271
\quad\text{or}\quad
1272
\b{K}^p \eqdef P^p R P^{\ell-p-1}. % \text{ for } p \le \ell-1.
1273
\end{gather*}
1274
(where we write a tuple of partitions of~$[3]$ as a word on~$\{P, Q_1, Q_2, Q_3, R\}$). %, and $p$ and $q$ are such that the total length is~$\ell$).
1275
%\begin{gather*}
1276
%\b{F} \eqdef (\underbrace{P, \dots, P}_\ell), \\
1277
%\b{G}_i^p \eqdef (\underbrace{P, \dots, P}_p, Q_i, \underbrace{P, \dots, P}_{\ell-p-1}) \text{ for } p \le \ell-1 \text{ and } i \in [3],
1278
%\\
1279
%\b{H}_{i,j}^{p,q} \eqdef (\underbrace{P, \dots, P}_p, Q_i, \underbrace{P, \dots, P}_{\ell-p-q-2}, Q_j, \underbrace{P, \dots, P}_q) \text{ for } p + q \le \ell-2 \text{ and } i \ne j \in [3],
1280
%\\
1281
%\text{or}\quad
1282
%\b{K}^p \eqdef (\underbrace{P, \dots, P}_p, R, \underbrace{P, \dots, P}_{\ell-p-1}) \text{ for } p \le \ell-1.
1283
%\end{gather*}
1284
Moreover, the cover relations in the $(\ell,3)$-partition forest poset are precisely the relations
1285
\[
1286
\b{F} \le \b{G}_i^p \hspace{-.4cm} \begin{array}{l} \rotatebox[origin=c]{45}{$\le$} \raisebox{.2cm}{$\b{H}_{i,j}^{p,q}$} \\[.1cm] \quad \le \b{K}^p \\[.1cm] \rotatebox[origin=c]{-45}{$\le$} \raisebox{-.2cm}{$\b{H}_{j,i}^{\ell-q-1, \ell-p-1}$} \end{array}
1287
\]
1288
for~$i \ne j$ and~$p, q$ such that~$p + q \le \ell-2$.
1289
Hence, we have
1290
\begin{align*}
1291
\mobPol[{\multiBraidArrangement[3][\ell]}] & = x^2 y^2 - 3 \ell x^2 y + \ell (3 \ell - 1) x^2 + 3 \ell x y - 3 \ell (2 \ell - 1) x + \ell (3 \ell - 2) , \\
1292
\fPol[{\multiBraidArrangement[3][\ell]}] & = (3 \ell^2 + 2 \ell + 1) x^2 + 6 \ell^2 x + \ell (3 \ell - 2), \\
1293
\text{and}\qquad
1294
\bPol[{\multiBraidArrangement[3][\ell]}] & = (3 \ell^2 - 4 \ell + 1) x^2 + 6 \ell (\ell - 1) x + \ell (3 \ell - 2).
1295
\end{align*}
1296
Observe that~$3 \ell^2 + 2 \ell + 1$ is~\OEIS{A056109}, that~$\ell (3 \ell - 2)$ is~\OEIS{A000567}, and that ${3 \ell^2 - 4 \ell + 1}$ is~\OEIS{A045944}.
1297
%\vincenti{There is a weird connection between the first and the last. Namely, $3 \ell^2 - 4 \ell + 1 = 3 (\ell - 1)^2 + 2 (\ell - 1)$. Is there a bijective explanation on the arrangements?}
1298
\end{example}
1299
1300
%%%%%%%%%%%%%%%
1301
1302
\subsection{Rainbow forests}
1303
\label{subsec:rainbowForests}
1304
1305
In order to obtain more explicit formulas for the number of vertices and regions of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ in \cref{subsec:verticesMultiBraidArrangement,subsec:regionsMultiBraidArrangement}, we now introduce another combinatorial model for $(\ell,n)$-partition forests which is more adapted to their enumeration.
1306
1307
\begin{definition}
1308
\label{def:rainbowForest}
1309
An \defn{$\ell$-rainbow coloring} of a rooted plane forest~$F$ is an assignment of colors of~$[\ell]$ to the non-root nodes of~$F$ such that
1310
\begin{enumerate}[(i)]
1311
\item there is no monochromatic edge,
1312
\item the colors of siblings are increasing from left to right.
1313
\end{enumerate}
1314
We denote by~$\|F\|$ the number of nodes of~$F$ and by~$\card{F}$ the number of trees of the forest~$F$ (\ie its number of connected components).
1315
An \defn{$(\ell,n)$-rainbow forest} (\resp \defn{tree}) is a \mbox{$\ell$-rainbow} colored forest (\resp tree) with $\|F\| = n$ nodes.
1316
We denote by~$\rainbowForests$ (\resp $\rainbowTrees$) the set of $(\ell,n)$-rainbow forests (\resp trees), and set~$\rainbowForests[][\ell] \eqdef \bigsqcup_n \rainbowForests$ (\resp $\rainbowTrees[][\ell] \eqdef \bigsqcup_n \rainbowTrees$).
1317
\end{definition}
1318
1319
For instance, we have listed the $14$ $(2,4)$-rainbow trees in \cref{fig:rainbowTrees}\,(top).
1320
This figure actually illustrates the following statement.
1321
1322
\begin{lemma}
1323
\label{lem:FussCatalan}
1324
The $(\ell,m)$-rainbow trees are counted by the \defn{Fuss-Catalan number}
1325
\[
1326
\card{\rainbowTrees[m][\ell]} = F_{\ell,m} \eqdef \frac{1}{(\ell-1)m+1} \binom{\ell m}{m} \qquad \text{\OEIS{A062993}}.
1327
\]
1328
\end{lemma}
1329
1330
\begin{proof}
1331
We can transform a $\ell$-rainbow tree~$R$ to an $\ell$-ary tree~$T$ as illustrated in \cref{fig:rainbowTrees}.
1332
Namely, the parent of a node~$N$ in~$T$ is the previous sibling colored as~$N$ in~$R$ if it exists, and the parent of~$N$ in~$R$ otherwise.
1333
This classical map is a bijection from $\ell$-rainbow trees to $\ell$-ary trees, which are counted by the Fuss-Catalan numbers~\cite{Klarner, HiltonPedersen}.
1334
%
1335
\begin{figure}
1336
\centerline{\includegraphics[scale=.7]{rainbowTrees}}
1337
\caption{The $14$ $(2,4)$-rainbow trees (top) and $14$ binary trees (bottom), and the simple bijection between them (middle). The order of the colors is red, blue.}
1338
\label{fig:rainbowTrees}
1339
\end{figure}
1340
\end{proof}
1341
1342
\begin{remark}
1343
\label{rem:functionalEquationFussCatalan}
1344
Recall that the corresponding generating function~$F_\ell(z) \eqdef \sum_{m \ge 0} F_{\ell,m} \, z^m$ satisfies the functional equation
1345
\[
1346
F_\ell(z) = 1 + z \, F_\ell(z)^\ell.
1347
\]
1348
\end{remark}
1349
1350
\begin{table}
1351
\centerline{\scalebox{.8}{
1352
\begin{tabular}[t]{c|ccccccccc}
1353
$m \backslash \ell$ & $1$ & $2$ & $3$ & $4$ & $5$ & $6$ & $7$ & $8$ & $9$ \\
1354
\hline
1355
$1$ & $1$ & $1$ & $1$ & $1$ & $1$ & $1$ & $1$ & $1$ & $1$ \\
1356
$2$ & $1$ & $2$ & $3$ & $4$ & $5$ & $6$ & $7$ & $8$ & $9$ \\
1357
$3$ & $1$ & $5$ & $12$ & $22$ & $35$ & $51$ & $70$ & $92$ & $117$ \\
1358
$4$ & $1$ & $14$ & $55$ & $140$ & $285$ & $506$ & $819$ & $1240$ & $1785$ \\
1359
$5$ & $1$ & $42$ & $273$ & $969$ & $2530$ & $5481$ & $10472$ & $18278$ & $29799$ \\
1360
$6$ & $1$ & $132$ & $1428$ & $7084$ & $23751$ & $62832$ & $141778$ & $285384$ & $527085$ \\
1361
$7$ & $1$ & $429$ & $7752$ & $53820$ & $231880$ & $749398$ & $1997688$ & $4638348$ & $9706503$ \\
1362
$8$ & $1$ & $1430$ & $43263$ & $420732$ & $2330445$ & $9203634$ & $28989675$ & $77652024$ & $184138713$ \\
1363
$9$ & $1$ & $4862$ & $246675$ & $3362260$ & $23950355$ & $115607310$ & $430321633$ & $1329890705$ & $3573805950$
1364
\end{tabular}
1365
}}
1366
% \vspace{.3cm}
1367
\caption{The Fuss-Catalan numbers~$F_{\ell,m} = \frac{1}{(\ell-1)m+1} \binom{\ell m}{m}$ for~$\ell,m \in [9]$. See \OEIS{A062993}.}
1368
\end{table}
1369
1370
\begin{definition}
1371
For a $(\ell,n)$-rainbow forest~$F$, we define
1372
\[
1373
\omega(F) \eqdef \prod_{i \in [\ell]} \prod_{N \in F} \card{C_i(N)}! ,
1374
\]
1375
where~$N$ ranges over all nodes of~$F$ and~$C_i(N)$ denotes the children of~$N$ colored by~$i$.
1376
\end{definition}
1377
1378
\begin{definition}
1379
\label{def:labelingRainbowForest}
1380
A \defn{labeling} of a $(\ell,n)$-rainbow forest~$F$ is a bijective map from the nodes of~$F$ to~$[n]$ such that
1381
\begin{enumerate}[(i)]
1382
\item the label of each root is minimal in its tree,
1383
\item the labels of siblings with the same color are increasing from left to right.
1384
\end{enumerate}
1385
\end{definition}
1386
1387
\begin{lemma}
1388
\label{lem:labelingRainbowForest}
1389
The number~$\lambda(F)$ of labelings of a $(\ell,n)$-rainbow forest~$F$ is given by
1390
\[
1391
\lambda(F) = \frac{n!}{\omega(F) \prod\limits_{T \in F} \|T\|} .
1392
\]
1393
\end{lemma}
1394
1395
\begin{proof}
1396
Out of all~$n!$ bijective maps from the nodes of~$F$ to~$[n]$, only~$1/\prod_{T \in F} \|T\|$ satisfy Condition~(i) of \cref{def:labelingRainbowForest}, and only $1/\prod_{i \in [\ell]} \prod_{N \in F} \card{C_i(N)}! = 1/\omega(F)$ satisfy Condition~(ii) of \cref{def:labelingRainbowForest}.
1397
\end{proof}
1398
1399
The following statement is illustrated in \cref{fig:forests}.
1400
1401
\begin{proposition}
1402
\label{prop:bijectionForests}
1403
There is a bijection from $(\ell,n)$-partition forests to labeled $(\ell,n)$-rainbow forests, such that if the partition forest~$\b{F}$ is sent to the labeled rainbow forest~$F$, then
1404
\[
1405
\dim(\b{F}) = \card{F}-1
1406
\qquad\text{and}\qquad
1407
\mu_{\forestPoset}(\HH, \b{F}) = (-1)^{n-\card{F}} \, \omega(F).
1408
\]
1409
\end{proposition}
1410
1411
\begin{proof}
1412
From a labeled $(\ell,n)$-rainbow forest~$F$, we construct a $(\ell,n)$-partition forest~$\b{F} \eqdef (F_1, \dots, F_\ell)$ whose $i$\ordinal{} partition~$F_i$ has a part~$\{N\} \cup C_i(N)$ for each node~$N$ of~$F$ not colored~$i$.
1413
Condition~(i) of \cref{def:rainbowForest} ensures that each $F_i$ is indeed a partition.
1414
1415
Conversely, start from a $(\ell,n)$-partition forest~$\b{F} \eqdef (F_1, \dots, F_i)$.
1416
Consider the colored clique graph~$K_{\b{F}}$ on~$[n]$ obtained by replacing each part in~$F_i$ by a clique of edges colored by~$i$.
1417
For each~$1 < j \le n$, there is a unique shortest path in~$K_{\b{F}}$ from the vertex~$j$ to the smallest vertex in the connected component of~$j$.
1418
Define the parent~$p$ of~$j$ to be the next vertex along this path, and color the node~$j$ by the color of the edge between~$j$ and~$p$.
1419
This defines a labeled $(\ell,n)$-rainbow forest~$F$.
1420
1421
Finally, observe that
1422
\begin{gather*}
1423
\dim(\b{F}) = n - 1 - \ell n + \sum_{i \in [\ell]} \card{F_i} = \card{F}-1, \qquad\text{and} \\
1424
\mu_{\forestPoset}(\HH, \b{F}) = \prod\limits_{i \in [\ell]} \prod\limits_{p \in F_i} (-1)^{\card{p}-1} (\card{p}-1) = \prod\limits_{i \in [\ell]} \prod\limits_{N \in F} (-1)^{\card{C_i(N)}} \card{C_i(N)}! = (-1)^{n-\card{F}} \, \omega(F).
1425
\qedhere
1426
\end{gather*}
1427
\end{proof}
1428
1429
We now transport via this bijection the partial order of the flat lattice on rainbow forests.
1430
For a node~$a$ of a forest~$F$, we denote by $\operatorname{Root}(a)$ the root of the tree of~$F$ containing~$a$.
1431
The following statement is illustrated in \cref{fig:CoverRelRF}, choosing $c$ to be green, $a$ to be $5$ and $b$ to be $7$.
1432
1433
\begin{figure}
1434
\centerline{\includegraphics[scale=1]{CoverRelRFFig}}
1435
\vspace{-.3cm}
1436
\caption{A covering relation described in \cref{prop:CoverRelRF}, choosing $c$ to be green, $a$ to be $5$ and $b$ to be $7$.}
1437
\label{fig:CoverRelRF}
1438
\end{figure}
1439
1440
\begin{proposition}
1441
\label{prop:CoverRelRF}
1442
In the flat poset~$\flatPoset[\multiBraidArrangement]$ labeled by rainbow forests using~\cref{prop:flatPosetMultiBraidArrangement,prop:bijectionForests}, a rainbow forest~$F$ is covered by a rainbow forest~$G$ if and only $G$ can be obtained from $F$ by:
1443
\begin{enumerate}
1444
\item choosing a color $c$, and two vertices $a$ and $b$ not colored with~$c$ and with~$\operatorname{Root}(a)<\operatorname{Root}(b)$,
1445
\item shifting the colors along the path from~$\operatorname{Root}(b)$ to~$b$, so that each node along this path is now colored by the former color of its child and~$b$ is not colored anymore,
1446
\item rerooting at~$b$ the tree containing~$b$ at~$b$, and coloring $b$ with~$c$,
1447
\item adding an edge~$(a,b)$ and replacing the edge~$(b,e)$ by an edge~$(a,e)$ for each child $e$ of~$b$ colored with~$c$.
1448
\end{enumerate}
1449
\end{proposition}
1450
1451
\begin{proof}
1452
Let us first remark that the graph obtained by these operations is indeed a rainbow forest.
1453
First, we add an edge between two distinct connected components, so that the result is indeed acyclic.
1454
Moreover, the condition on the color of $a$ and on the deletion of edges between $b$ and vertices of color $c$ ensures that we do not add an edge between two vertices of the same color.
1455
Note that the parent of $b$ inherits the color of $b$ which is not $c$.
1456
1457
Let us recall that the cover relations in the flat poset~$\flatPoset[\multiBraidArrangement]$ are given in terms of $(\ell,n)$-partition forests by choosing a partition $\pi$ of the partition tuple (which corresponds directly to choosing a color), choosing two parts $\pi_a$ and $\pi_b$ in the partition $\pi$, and merging them, without creating a loop in the intersection hypergraph.
1458
1459
By choosing two vertices in different connected components of the rainbow forest, we are sure that the intersection hypergraph obtained by adding an edge is still acyclic.
1460
1461
The last point that has to be explained is the link between the condition on the color of~$a$ and~$b$ and merging two parts in the same partition.
1462
If one of the two nodes, say $a$ for instance is of color~$c$, then it belongs to the same part of $\pi$ as its parent $z$.
1463
The merging is the same if we choose~$z$ which is not colored $c$.
1464
Moreover, as $b$ is in a different connected component, the corresponding two parts are distinct in $\pi$.
1465
Finally, a part is just a corolla so the merging corresponds to building a corolla with $a$, $b$ and their children of color $c$.
1466
\end{proof}
1467
1468
We finally recast \cref{prop:alternativeFormulaMobiusPolynomialMultiBraidArrangement} in terms of rainbow forests.
1469
1470
\begin{proposition}
1471
The M\"obius polynomial of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ is given by
1472
\[
1473
\mobPol[\multiBraidArrangement] = x^{(n-1)(1-\ell)} \sum_{G \in \rainbowForests} y^{n-1+\card{E(G)}} \prod_{i \in [\ell]} \weirdPol[n-\card{E(G,i)}].
1474
\]
1475
\end{proposition}
1476
1477
\begin{remark}
1478
To further simplify this expression, we would need to count the number of rainbow forests with a prescribed number of colored edges.
1479
However, this number does not admit a known multiplicative formula, up to our knowledge. When there is only one color, the corresponding sequence (counting non-colored forests on $n$ nodes and $k$ edges, rooted in the minimal label of each connected component) is \OEIS{A138464}.
1480
\end{remark}
1481
1482
%%%%%%%%%%%%%%%
1483
1484
\subsection{Enumeration of vertices of $\multiBraidArrangement$}
1485
\label{subsec:verticesMultiBraidArrangement}
1486
1487
We now use the labeled $(\ell,n)$-rainbow forests of \cref{subsec:rainbowForests} to derive more explicit formulas for the number of vertices of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$.
1488
The first few values are gathered in \cref{table:verticesMultiBraidArrangement}.
1489
1490
\begin{table}
1491
\centerline{\scalebox{.8}{
1492
\begin{tabular}[t]{c|cccccccc}
1493
$n \backslash \ell$ & $1$ & $2$ & $3$ & $4$ & $5$ & $6$ & $7$ & $8$ \\ % & $9$ \\
1494
\hline
1495
$1$ & $1$ & $1$ & $1$ & $1$ & $1$ & $1$ & $1$ & $1$ \\ % & $1$ \\
1496
$2$ & $1$ & $2$ & $3$ & $4$ & $5$ & $6$ & $7$ & $8$ \\ % & $9$ \\
1497
$3$ & $1$ & $8$ & $21$ & $40$ & $65$ & $96$ & $133$ & $176$ \\ % & $225$ \\
1498
$4$ & $1$ & $50$ & $243$ & $676$ & $1445$ & $2646$ & $4375$ & $6728$ \\ % & $9801$ \\
1499
$5$ & $1$ & $432$ & $3993$ & $16384$ & $46305$ & $105456$ & $208537$ & $373248$ \\ % & $620289$ \\
1500
$6$ & $1$ & $4802$ & $85683$ & $521284$ & $1953125$ & $5541126$ & $13119127$ & $27350408$ \\ % & $51883209$ \\
1501
$7$ & $1$ & $65536$ & $2278125$ & $20614528$ & $102555745$ & $362797056$ & $1029059101$ & $2500000000$ \\ % & $5415228513$ \\
1502
$8$ & $1$ & $1062882$ & $72412707$ & $976562500$ & $6457339845$ & $28500625446$ & $96889010407$ & $274371577992$ % \\ & $678770015625$ \\
1503
% $9$ & $1$ & $20000000$ & $2681615217$ & $53971714048$ & $474659385665$ & $2614905943296$ & $10657046640625$ & $35184372088832$ & $99426586671873$
1504
\end{tabular}
1505
}}
1506
% \vspace{.3cm}
1507
\caption{The numbers $f_0(\multiBraidArrangement) = \ell \big( (\ell-1) n + 1 \big)^{n-2}$ of vertices of~$\multiBraidArrangement$ for~$\ell,n \in [8]$.}
1508
\label{table:verticesMultiBraidArrangement}
1509
\end{table}
1510
1511
\begin{theorem}
1512
\label{thm:verticesMultiBraidArrangement}
1513
The number of vertices of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ is
1514
\[
1515
f_0(\multiBraidArrangement) = \ell \big( (\ell-1) n + 1 \big)^{n-2}.
1516
\]
1517
\end{theorem}
1518
1519
\begin{proof}
1520
By \cref{prop:flatPosetMultiBraidArrangement,prop:bijectionForests}, we just need to count the labeled $(\ell,n)$-rainbow trees.
1521
A common reasoning for counting Cayley trees is the use of its Prüfer code defined by recursively pruning the smallest leaf while writing down the label of its parent.
1522
This bijection can be adapted to colored Cayley trees by writing down the label of the parent colored by the color of the pruned leaf.
1523
This leads to a bijection with certain colored words of length $n-1$.
1524
Namely, there are two possibilities:
1525
\begin{itemize}
1526
\item either the pruned leaf is attached to the node~$1$ and it can have all $\ell$ colors,
1527
\item or it is attached to one of the $n-1$ other nodes and it can only have $\ell-1$ colors.
1528
\end{itemize}
1529
Note that the last letter in the Prüfer code (obtained by removing the last edge) is necessarily the root $1$, with $\ell$ possible different colors.
1530
Hence, there are
1531
\[
1532
\big( \ell+(n-1)(\ell-1) \big)^{n-2} \ell = \ell \big( (\ell-1) n + 1 \big)^{n-2}
1533
\]
1534
such words.
1535
Similar ideas were used in~\cite{Lewis}.
1536
\end{proof}
1537
1538
We can refine the formula of \cref{thm:verticesMultiBraidArrangement} according to the dimension of the flats of the different copies intersected to obtain the vertices of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$.
1539
1540
\begin{theorem}
1541
\label{thm:verticesRefinedMultiBraidArrangement}
1542
For any~$k_1, \dots, k_\ell$ such that~$0 \le k_i \le n-1$ for~$i \in [\ell]$ and~${\sum_{i \in [\ell]} k_i = n-1}$, the number of vertices~$v$ of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ such that the smallest flat of the $i$\ordinal{} copy of~$\braidArrangement$ containing~$v$ has dimension~$n-k_i-1$ is given by
1543
\[
1544
n^{\ell-1} \binom{n-1}{k_1, \dots, k_\ell} \prod_{i \in [\ell]} (n-k_i)^{k_i-1}.
1545
\]
1546
\end{theorem}
1547
1548
\begin{proof}
1549
By \cref{prop:flatPosetMultiBraidArrangement,prop:bijectionForests}, we just need to count the labeled $(\ell,n)$-rainbow trees with~$k_i$ nodes colored by~$i$.
1550
Forgetting the labels, the $(\ell,n)$-rainbow trees with~$k_i$ nodes colored by~$i$ are precisely the spanning trees of the complete multipartite graph~$K_{k_1, \dots, k_\ell, 1}$ (where the last~$1$ stands for the uncolored root).
1551
Using a Pr\"ufer code similar to that of the proof of \cref{thm:verticesMultiBraidArrangement}, R.~Lewis proved in~\cite{Lewis} that the latter are counted by~${n^{\ell-1} \prod_{i \in [\ell]} (n-k_i)^{k_i-1}}$.
1552
Finally, the possible labelings are counted by the multinomial coefficient~$\binom{n-1}{k_1, \dots, k_\ell}$.
1553
\end{proof}
1554
1555
%%%%%%%%%%%%%%%
1556
1557
\subsection{Enumeration of regions and bounded regions of $\multiBraidArrangement$}
1558
\label{subsec:regionsMultiBraidArrangement}
1559
1560
\enlargethispage{.2cm}
1561
We finally use the labeled $(\ell,n)$-rainbow forests of \cref{subsec:rainbowForests} to derive more explicit formulas for the number of regions and bounded regions of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$.
1562
The first few values are gathered in \cref{table:regionsMultiBraidArrangement,table:boundedRegionsMultiBraidArrangement}.
1563
We first compute the characteristic polynomial of~$\multiBraidArrangement$.
1564
1565
\afterpage{
1566
\begin{table}
1567
\centerline{\scalebox{.8}{
1568
\begin{tabular}[t]{c|cccccccc}
1569
$n \backslash \ell$ & $1$ & $2$ & $3$ & $4$ & $5$ & $6$ & $7$ & $8$ \\ % & $9$ \\
1570
\hline
1571
$1$ & $1$ & $1$ & $1$ & $1$ & $1$ & $1$ & $1$ & $1$ \\ % & $1$ \\
1572
$2$ & $2$ & $3$ & $4$ & $5$ & $6$ & $7$ & $8$ & $9$ \\ % & $10$ \\
1573
$3$ & $6$ & $17$ & $34$ & $57$ & $86$ & $121$ & $162$ & $209$ \\ % & $262$ \\
1574
$4$ & $24$ & $149$ & $472$ & $1089$ & $2096$ & $3589$ & $5664$ & $8417$ \\ % & $11944$ \\
1575
$5$ & $120$ & $1809$ & $9328$ & $29937$ & $73896$ & $154465$ & $287904$ & $493473$ \\ % & $793432$ \\
1576
$6$ & $720$ & $28399$ & $241888$ & $1085157$ & $3442816$ & $8795635$ & $19376064$ & $38323753$ \\ % & $69841072$ \\
1577
$7$ & $5040$ & $550297$ & $7806832$ & $49075065$ & $200320816$ & $625812385$ & $1629858672$ & $3720648337$ \\ % & $7686190000$ \\
1578
$8$ & $40320$ & $12732873$ & $302346112$ & $2666534049$ & $14010892416$ & $53536186825$ & $164859458688$ & $434390214657$ % \\ & $1017282905344$ \\
1579
% $9$ & $362880$ & $343231361$ & $13682809216$ & $169423639713$ & $1146173002496$ & $5357227099105$ & $19506923076096$ & $59328538244801$ & $157507267166848$
1580
\end{tabular}
1581
}}
1582
% \vspace{.3cm}
1583
\caption{The numbers $f_{n-1}(\multiBraidArrangement)$ of regions of~$\multiBraidArrangement$ for~$\ell,n \in [8]$.}
1584
\label{table:regionsMultiBraidArrangement}
1585
\end{table}
1586
}
1587
1588
\afterpage{
1589
\begin{table}
1590
\centerline{\scalebox{.8}{
1591
\begin{tabular}[t]{c|cccccccc}
1592
$n \backslash \ell$ & $1$ & $2$ & $3$ & $4$ & $5$ & $6$ & $7$ & $8$ \\ % & $9$ \\
1593
\hline
1594
$1$ & $1$ & $1$ & $1$ & $1$ & $1$ & $1$ & $1$ & $1$ \\ % & $1$ \\
1595
$2$ & $0$ & $1$ & $2$ & $3$ & $4$ & $5$ & $6$ & $7$ \\ % & $8$ \\
1596
$3$ & $0$ & $5$ & $16$ & $33$ & $56$ & $85$ & $120$ & $161$ \\ % & $208$ \\
1597
$4$ & $0$ & $43$ & $224$ & $639$ & $1384$ & $2555$ & $4248$ & $6559$ \\ % & $9584$ \\
1598
$5$ & $0$ & $529$ & $4528$ & $17937$ & $49696$ & $111745$ & $219024$ & $389473$ \\ % & $644032$ \\
1599
$6$ & $0$ & $8501$ & $120272$ & $663363$ & $2354624$ & $6455225$ & $14926176$ & $30583847$ \\ % & $57255488$ \\
1600
$7$ & $0$ & $169021$ & $3968704$ & $30533409$ & $138995776$ & $464913325$ & $1268796096$ & $2996735329$ \\ % & $6353133184$ \\
1601
$8$ & $0$ & $4010455$ & $156745472$ & $1684352799$ & $9841053184$ & $40179437975$ & $129465630720$ & $352560518527$ % \\ & $846588258944$ \\
1602
% $9$ & $0$ & $110676833$ & $7216242688$ & $108413745057$ & $813420601856$ & $4055310777025$ & $15431698810368$ & $48461340225473$ & $131823966149632$
1603
\end{tabular}
1604
}}
1605
% \vspace{.3cm}
1606
\caption{The numbers $b_{n-1}(\multiBraidArrangement)$ of bounded regions of~$\multiBraidArrangement$ for~$\ell,n \in [8]$.}
1607
\label{table:boundedRegionsMultiBraidArrangement}
1608
\end{table}
1609
}
1610
1611
\begin{theorem}
1612
\label{thm:characteristicPolynomialMultiBraidArrangement}
1613
The characteristic polynomial~$\charPol[\multiBraidArrangement]$ of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ is given~by
1614
\[
1615
\charPol[\multiBraidArrangement] = \frac{(-1)^n n!}{y} \, [z^n] \, \exp \bigg( - \sum_{m \ge 1} \frac{F_{\ell,m} \, y \, z^m}{m} \bigg) ,
1616
\]
1617
where~$\displaystyle F_{\ell,m} \eqdef \frac{1}{(\ell-1)m+1} \binom{\ell m}{m}$ is the Fuss-Catalan number.
1618
\end{theorem}
1619
1620
\begin{proof}
1621
By \cref{thm:MobiusPolynomialMultiBraidArrangement,prop:bijectionForests}, the characteristic polynomial~$\charPol[\multiBraidArrangement]$ is
1622
\[
1623
\charPol[\multiBraidArrangement] = \sum_{\b{F} \in \forestPoset} \mu_{\forestPoset}(\HH, \b{F}) \, y^{\dim(\b{F})} = \sum_{F \in \rainbowForests} \lambda(F) \, (-1)^{n-\card{F}} \, \omega(F) \, y^{\card{F}-1}.
1624
\]
1625
From \cref{lem:labelingRainbowForest}, we observe that
1626
\[
1627
\frac{\lambda(F) \, \omega(F) \, (-y)^{\card{F}} \, z^{\|F\|}}{\|F\|!} = \prod_{T \in F} \frac{-y \, z^{\|T\|}}{\|T\|} ,
1628
\]
1629
where $T$ ranges over the trees of~$F$.
1630
Now using that rainbow forests are exactly sets of rainbow trees, we obtain that
1631
\[
1632
\sum_{F \in \rainbowForests[][\ell]} \frac{ \lambda(F) \, \omega(F) \, (-y)^{\card{F}} \, z^{\|F\|}}{\|F\|!} = \sum_{F \in \rainbowForests[][\ell]} \prod_{T \in F} \frac{-y \, z^{\|T\|}}{\|T\|} = \exp \bigg( \sum_{T \in \rainbowTrees[][\ell]} \frac{-y \, z^{\|T\|}}{\|T\|} \bigg).
1633
\]
1634
From \cref{lem:FussCatalan}, we obtain that
1635
\[
1636
\exp \bigg( \sum_{T \in \rainbowTrees[][\ell]} \frac{-y \, z^{\|T\|}}{\|T\|} \bigg) = \exp \bigg( - \sum_{m \ge 1} \frac{F_{\ell,m} \, y \, z^m}{m} \bigg).
1637
\]
1638
We conclude that
1639
\begin{align*}
1640
\charPol[\multiBraidArrangement]
1641
& = \sum_{F \in \rainbowForests} \lambda(F) \, (-1)^{n-\card{F}} \, \omega(F) \, y^{\card{F}-1} \\
1642
& = \frac{(-1)^n \, n!}{y} [z^n] \sum_{F \in \rainbowForests[][\ell]} \frac{ \lambda(F) \, \omega(F) \, (-y)^{\card{F}} \, z^{\|F\|}}{\|F\|!} \\
1643
& = \frac{(-1)^n n!}{y} \, [z^n] \, \exp \bigg( - \sum_{m \ge 1} \frac{F_{\ell,m} \, y \, z^m}{m} \bigg).
1644
\qedhere
1645
\end{align*}
1646
\end{proof}
1647
1648
From the characteristic polynomial of~$\multiBraidArrangement$ and \cref{rem:characteristicPolynomial}, we obtain its numbers of regions and bounded regions.
1649
1650
\begin{theorem}
1651
\label{thm:regionsMultiBraidArrangement}
1652
The numbers of regions and of bounded regions of the $(\ell,n)$-braid arrangement~$\multiBraidArrangement$ are given by
1653
\begin{align*}
1654
f_{n-1}(\multiBraidArrangement)
1655
& = n! \, [z^n] \exp \Bigg( \sum_{m \ge 1} \frac{F_{\ell,m} \, z^m}{m} \Bigg) \\
1656
\text{and}\qquad
1657
b_{n-1}(\multiBraidArrangement)
1658
%& = - n! \, [z^n] \exp \Bigg( - \sum_{m \ge 1} \frac{F_{\ell,m} \, z^m}{m} \Bigg)
1659
& = (n-1)! \, [z^{n-1}] \exp \bigg( (\ell-1) \sum_{m \ge 1} F_{\ell,m} \, z^m \bigg),
1660
\end{align*}
1661
where~$\displaystyle F_{\ell,m} \eqdef \frac{1}{(\ell-1)m+1} \binom{\ell m}{m}$ is the Fuss-Catalan number.
1662
\end{theorem}
1663
1664
\begin{proof}
1665
By \cref{rem:characteristicPolynomial}, we obtain from \cref{thm:characteristicPolynomialMultiBraidArrangement} that
1666
\begin{align*}
1667
f_{n-1}(\multiBraidArrangement) & = (-1)^{n-1} \charPol[\multiBraidArrangement][-1] = n! \, [z^n] \, \exp \bigg( \sum_{m \ge 1} \frac{F_{\ell,m} \, z^m}{m} \bigg), \\
1668
% \text{and}\qquad
1669
b_{n-1}(\multiBraidArrangement) & = (-1)^{n-1} \charPol[\multiBraidArrangement][1] = - n! \, [z^n] \, \exp \bigg( - \sum_{m \ge 1} \frac{F_{\ell,m} \, z^m}{m} \bigg).
1670
\end{align*}
1671
To conclude, we thus just need to observe that
1672
\(
1673
U_\ell(z) = \frac{\partial}{\partial z} V_\ell(z)
1674
\)
1675
where
1676
\[
1677
U_\ell(z) \eqdef \exp \bigg( (\ell-1) \sum_{m \ge 1} F_{\ell,m} \, z^m \bigg)
1678
\qquad\text{and}\qquad
1679
V_\ell(z) \eqdef - \exp \bigg( - \sum_{m \ge 1} \frac{F_{\ell,m} \, z^m}{m} \bigg).
1680
\]
1681
For this, consider the generating functions
1682
\[
1683
F_\ell(z) \eqdef \sum_{m \ge 0} F_{\ell,m} \, z^m
1684
\qquad\text{and}\qquad
1685
G_\ell(z) \eqdef \sum_{m \ge 1} \frac{F_{\ell,m} \, z^m}{m}.
1686
\]
1687
Recall from \cref{rem:functionalEquationFussCatalan} that~$F_\ell(z)$ satisfies the functional equation
1688
\[
1689
F_\ell(z) = 1 + z \, F_\ell(z)^\ell.
1690
\]
1691
We thus obtain that
1692
\[
1693
F_\ell'(z) \big( 1 - \ell \, z \, F_\ell(z)^{\ell-1} \big) = F_\ell(z)^\ell
1694
\quad\text{and}\quad
1695
F_\ell(z) \big( 1 - \ell \, z \, F_\ell(z)^{\ell-1} \big) = 1 - (\ell-1) \, z \, F_\ell(z)^\ell.
1696
\]
1697
Combining these two equations, we get
1698
\begin{equation}
1699
\label{eq:diff}
1700
F_\ell(z)^{\ell+1} = F_\ell'(z) \big( 1 - (\ell-1) \, z \, F_\ell(z)^\ell \big).
1701
\end{equation}
1702
Observe now that
1703
\begin{equation}
1704
\label{eq:GF}
1705
z \, G_\ell'(z) = F_\ell(z) - 1 = z \, F_\ell(z)^\ell
1706
\qquad\text{and}\qquad
1707
G_\ell''(z) = \ell \, F_\ell(z)^{\ell-1} \, F_\ell'(z).
1708
\end{equation}
1709
Hence
1710
\[
1711
U_\ell(z) = \exp \big( (\ell-1) \, (F_\ell(z) - 1) \big) = \exp \big( (\ell-1) \, z \, G_\ell’(z) \big)
1712
\]
1713
and
1714
\[
1715
V_\ell'(z) = \frac{\partial}{\partial z} - \exp \big( \! - G_\ell(z) \big) = G_\ell'(z) \exp \big( -G_\ell(z) \big).
1716
\]
1717
Consider now the function
1718
\[
1719
W_\ell(z) = V_\ell'(z) / U_\ell(z) = G_\ell'(z) \exp \big( \! - G_\ell(z) - (\ell-1) \, z \, G_\ell'(z) \big).
1720
\]
1721
Clearly, $W_\ell(0) = 1$.
1722
Moreover, using~\eqref{eq:GF}, we obtain that its derivative is
1723
\begin{align*}
1724
W_\ell'(z)
1725
& = \Big( G_\ell''(z) \big(1 - (\ell-1) \, z \, G_\ell'(z) \big) - \ell \, G_\ell'(z)^2 \Big) \exp \big( \! - G_\ell(z) - (\ell-1) \, z \, G_\ell'(z) \big) \\
1726
& = \ell \, F_\ell(z)^{\ell-1} \Big( F_\ell'(z) \big( 1 - (\ell-1) \, z \, F_\ell(z)^\ell \big) - F_\ell(z)^{\ell+1} \Big) \exp \big( \! - G_\ell(z) - (\ell-1) \, z \, G_\ell'(z) \big),
1727
\end{align*}
1728
which vanishes by~\eqref{eq:diff}.
1729
\end{proof}
1730
1731
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1732
1733
\section{Face poset and combinatorial description of~$\multiBraidArrangement(\b{a})$}
1734
\label{sec:facePoset}
1735
1736
In this section, we describe the face poset of the $\b{a}$-braid arrangement~$\multiBraidArrangement(\b{a})$ in terms of ordered $(\ell,n)$-partition forests.
1737
This section highly depends on the choice of the translation matrix~$\b{a}$.
1738
1739
%%%%%%%%%%%%%%%
1740
1741
\subsection{Ordered partition forests}
1742
\label{subsec:orderedPartitionForests}
1743
1744
We now introduce the combinatorial objects that will be used to encode the faces of the $\b{a}$-braid arrangement~$\multiBraidArrangement(\b{a})$ of \cref{def:multiBraidArrangementPrecise}.
1745
1746
\begin{definition}
1747
\label{def:orderedPartitionForest}
1748
An \defn{ordered $(\ell,n)$-partition forest} (\resp \defn{tree}) is an $\ell$-tuple~$\order{\b{F}} \eqdef (\order{F_1}, \dots, \order{F_\ell})$ of ordered set partitions of~$[n]$ such that the corresponding $\ell$-tuple~$\b{F} \eqdef (F_1, \dots, F_\ell)$ of unordered set partitions of~$[n]$ forms an $(\ell,n)$-partition forest (\resp tree).
1749
The \defn{ordered $(\ell,n)$-partition forest poset} is the poset~$\orderedForestPoset$ on ordered $(\ell,n)$-partition forests ordered by componentwise refinement.
1750
In other words, $\orderedForestPoset$ is the subposet of the $\ell$\ordinal{} Cartesian power of the ordered partition poset~$\orderedPartitionPoset$ induced by ordered $(\ell,n)$-partition forests.
1751
Note that the maximal elements of~$\orderedForestPoset$ are the ordered $(\ell, n)$-partition trees.
1752
\end{definition}
1753
1754
The following statement is the analogue of \cref{prop:flatPosetMultiBraidArrangement}, and is illustrated in \cref{fig:B23a,fig:B23b}.
1755
1756
\begin{proposition}
1757
\label{prop:facePosetMultiBraidArrangement}
1758
The face poset~$\facePoset[\multiBraidArrangement(\b{a})]$ of the $\b{a}$-braid arrangement~$\multiBraidArrangement(\b{a})$ is isomorphic to an upper set~$\orderedForestPoset(\b{a})$ of the ordered $(\ell,n)$-partition forest poset~$\orderedForestPoset$.
1759
\end{proposition}
1760
1761
\begin{proof}
1762
The proof is based on that of \cref{prop:flatPosetMultiBraidArrangement}.
1763
A face of~$\multiBraidArrangement(\b{a})$ is an intersection of faces of the $\ell$ copies of~$\multiBraidArrangement$, hence corresponds to an $\ell$-tuple of ordered partitions of~$[n]$.
1764
Moreover, the flats supporting these faces intersect, so that the corresponding unordered partitions must form an $(\ell,n)$-partition forest.
1765
Hence, each face of~$\multiBraidArrangement(\b{a})$ corresponds to a certain ordered $(\ell,n)$-partition forest.
1766
Moreover, the inclusion of faces of~$\multiBraidArrangement(\b{a})$ translates to the componentwise refinement on ordered partitions.
1767
Finally, by genericity, it is immediate that we obtain an upper set of this componentwise refinement order.
1768
\end{proof}
1769
1770
\begin{figure}
1771
\centerline{\includegraphics[scale=.9]{B23a}}
1772
\caption{Labelings of the faces of the arrangement~$\multiBraidArrangement[3][2](\b{a})$ for~$\b{a} = \begin{bmatrix} 0 & 0 \\ -1 & -1 \end{bmatrix}$.}
1773
\label{fig:B23a}
1774
\end{figure}
1775
1776
\begin{figure}
1777
\centerline{\includegraphics[scale=.9]{B23b}}
1778
\caption{Labelings of the faces of the arrangement~$\multiBraidArrangement[3][2](\b{a})$ for~$\b{a} = \begin{bmatrix} 0 & 0 \\ 1 & -2 \end{bmatrix}$.}
1779
\label{fig:B23b}
1780
\end{figure}
1781
1782
We now fix a generic translation matrix~$\b{a} \eqdef (a_{i,j})$ and still denote by~$A_{i,s,t} \eqdef \smash{\sum_{s \le j < t} a_{i,j}}$ for all~$1 \le s < t \le n$ and~$i \in [\ell]$ (and often write~$A_{i,t,s}$ for~$-A_{i,s,t}$).
1783
The objective of this section is to describe
1784
\begin{itemize}
1785
\item the ordered $(\ell,n)$-partitions forests of the upper set~$\orderedForestPoset(\b{a})$ with a given underlying (unordered) $(\ell,n)$-partition forest (\cref{subsec:PFtoOPF}),
1786
\item a criterion to decide whether a given ordered $(\ell,n)$-partition forest belongs to the upper set~$\orderedForestPoset(\b{a})$, \ie corresponds to a face of~$\multiBraidArrangement(\b{a})$ (\cref{subsec:criterionOPF}).
1787
\end{itemize}
1788
1789
%%%%%%%%%%%%%%%
1790
1791
\subsection{From partition forests to ordered partition forests}
1792
\label{subsec:PFtoOPF}
1793
1794
In this section, we describe the ordered $(\ell,n)$-partitions forests~$\order{\b{F}}$ of the upper set~$\orderedForestPoset(\b{a})$ with a given underlying $(\ell,n)$-partition forest~$\b{F}$.
1795
We denote by~$cc(\b{F})$ the connected components of~$\b{F}$, meaning the partition of~$[n]$ given by the hyperedge labels of the connected components of the intersection hypergraph of~$\b{F}$.
1796
We first observe that the choice of~$\b{a}$ fixes the order of the parts in a common connected component of~$\b{F}$.
1797
1798
\begin{proposition}
1799
\label{prop:PFtoOPF1}
1800
Consider a $(\ell,n)$-partition forest~$\b{F} \eqdef (F_1, \dots, F_\ell)$, and two integers~$s,t \in [n]$ labeling two hyperedges in the same connected component of the intersection hypergraph of~$\b{F}$.
1801
Assume that the unique path from~$s$ to~$t$ in the hypergraph of~$\b{F}$ passes through the hyperedges labeled by~$s = r_0, \dots, r_q = t$ and through parts of the partitions~$F_{i_1}, \dots, F_{i_q}$.
1802
Then for any ordered $(\ell,n)$-partition forest~$\order{\b{F}} \eqdef (\order{F_1}, \dots, \order{F_\ell})$ of the upper set~$\orderedForestPoset(\b{a})$ with underlying $(\ell,n)$-partition forest~$\b{F}$ and any~$i \in [\ell]$, the order of~$s$ and~$t$ in~$\order{F}_i$ is given by the sign of~$A_{i,s,t} - \sum_{p \in [q]} A_{i_p, r_{p-1}, r_p}$.
1803
\end{proposition}
1804
1805
\begin{proof}
1806
Consider any point~$\b{x}$ in the face of~$\multiBraidArrangement(\b{a})$ corresponding to~$\order{\b{F}}$.
1807
Along the path from~$s$ to~$t$, we have~$x_{r_{p-1}} - x_{r_p} = A_{i_p, r_{p-1}, r_p}$ for each~$p \in [q]$.
1808
Hence, we obtain that
1809
\[
1810
x_s - x_t = \sum_{p \in [q]} (x_{r_{p-1}} - x_{r_p}) = \sum_{p \in [q]} A_{i_p, r_{p-1}, r_p}.
1811
\]
1812
The order of~$s,t$ in~$\order{F}_i$ is given by the sign of~$A_{i,s,t} - (x_s - x_t)$, hence of~$A_{i,s,t} - \sum_{p \in [q]} A_{i_p, r_{p-1}, r_p}$.
1813
\end{proof}
1814
1815
We now describe the different ways to order the parts in distinct connected components of~$\b{F}$.
1816
For this, we need the following posets.
1817
1818
\begin{definition}
1819
Consider a $(\ell,n)$-partition forest~$\b{F}$ and denote by~$cc(\b{F})$ the connected components of~$\b{F}$.
1820
For each pair~$s,t \in [n]$ in distinct connected components of~$cc(\b{F})$, we define the chain~$<_{s,t}$ on the $\ell$ triples~$(i,s,t)$ for~$i \in [\ell]$ given by the order of the values~$A_{i,s,t}$.
1821
The \defn{inversion poset}~$\Inv(\b{F}, \b{a})$ is then the poset obtained by quotienting the disjoint union of the chains~$<_{s,t}$ (for all~$s,t \in [n]$ in distinct connected components of~$cc(\b{F})$) by the equivalence relation~$(i,s,t) \equiv (i,s',t')$ if~$s$ and~$s'$ belong to the same part of~$F_i$ and~$t$ and~$t'$ belong to the same part of~$F_i$.
1822
We say that a subset~$X$ of~$\Inv(\b{F}, \b{a})$ is antisymmetric if~$(i,s,t) \in X \iff (i,t,s) \notin X$.
1823
\end{definition}
1824
1825
\begin{proposition}
1826
\label{prop:PFtoOPF2}
1827
The ordered $(\ell,n)$-partition forests of the upper set~$\orderedForestPoset(\b{a})$ with a given underlying $(\ell,n)$-partition forest~$\b{F}$ are in bijection with the antisymmetric lower sets of the inversion poset~$\Inv(\b{F}, \b{a})$.
1828
\end{proposition}
1829
1830
\begin{proof}
1831
Consider an ordered $(\ell,n)$-partition forest~$\order{\b{F}}$ of the upper set~$\orderedForestPoset(\b{a})$.
1832
Let~$\b{x}$ be any point of the face of~$\multiBraidArrangement(\b{a})$ corresponding to~$\order{\b{F}}$.
1833
For each pair~$s,t \in [n]$ in distinct connected components of~$cc(\b{F})$, let~$I_{s,t}(\order{\b{F}})$ be the set of indices~$i \in [\ell]$ such that~$x_s - x_t < A_{i,s,t}$.
1834
Note that~$I_{s,t}$ is by definition a lower set of the chain~$<_{s,t}$ of~$\Inv(\b{F}, \b{a})$.
1835
Hence, $I(\order{F}) \eqdef \bigcup_{s,t} I_{s,t} / {\equiv}$ is a lower set of~$\Inv(\b{F}, \b{a})$.
1836
Moreover, it is clearly antisymmetric since
1837
\[
1838
(i,s,t) \in I(\order{F}) \iff x_s - x_t < A_{i,s,t} \iff x_t - x_s > A_{i,t,s} \iff (i,t,s) \notin I(\order{F}).
1839
\]
1840
1841
Conversely, given an antisymmetric lower set~$I$ of~$\Inv(\b{F}, \b{a})$, we can reconstruct an ordered $(\ell,n)$-partition forest~$\order{\b{F}}$ by ordering each pair~$s,t \in [n]$ in~$\order{F}_i$
1842
\begin{itemize}
1843
\item according to \cref{prop:PFtoOPF1} (hence independently of~$I$) if $s$ and~$t$ belong to the same connected component of~$\b{F}$,
1844
\item according to~$I$ if~$s$ and~$t$ belong to distinct connected components of~$\b{F}$. Namely, we place the block of~$\order{F}_i$ containing~$s$ before the block of~$\order{F}_i$ containing~$t$ if and only if~$(i,s,t) \in I$.
1845
\end{itemize}
1846
It is then straightforward to check that the resulting ordered $(\ell,n)$-partition forest belongs to the upper set~$\orderedForestPoset(\b{a})$, by exhibiting a point~$\b{x}$ in of the corresponding face of~$\multiBraidArrangement(\b{a})$.
1847
\end{proof}
1848
1849
%%%%%%%%%%%%%%%
1850
1851
\subsection{A criterion for ordered partition forests}
1852
\label{subsec:criterionOPF}
1853
1854
We now consider a given ordered $(\ell,n)$-partition forest~$\order{F}$ and provide a criterion to decide if it belongs to the upper set~$\orderedForestPoset(\b{a})$ corresponding to the faces of~$\multiBraidArrangement(\b{a})$.
1855
For this, we need the following directed graph associated to~$\order{\b{F}}$.
1856
1857
\begin{definition}
1858
For an ordered partition~$\order{\pi} \eqdef \order{\pi}_1 | \cdots | \order{\pi}_k$ of~$[n]$, we denote by~$D_{\order{\pi}}$ the directed graph on~$[n]$ with an arc~$\max(\order{\pi}_j) \to \min(\order{\pi}_{j+1})$ for each~$j \in [k-1]$ and a cycle~${x_1 \to \dots \to x_p \to x_1}$ for each part~$\order{\pi}_j = \{x_1 < \dots < x_p\}$.
1859
Note that~$D_{\order{\pi}}$ has~$n$ vertices and~$n + k$ arcs.
1860
For an ordered $(\ell,n)$-partition forest~$\order{\b{F}} \eqdef (\order{F_1}, \dots, \order{F_\ell})$, we denote by~$D_{\order{\b{F}}}$ the superposition of the directed graphs~$D_{\order{F}_i}$ for~$i \in [\ell]$, where the arcs of~$D_{\order{F}_i}$ are labeled by~$i$.
1861
\end{definition}
1862
1863
\begin{proposition}
1864
\label{prop:characterizationOPFs}
1865
An ordered $(\ell,n)$-partition forest~$\order{\b{F}}$ belongs to the upper set~$\orderedForestPoset(\b{a})$ if and only if $\sum_{\alpha \in \gamma} A_{i(\alpha), s(\alpha), t(\alpha)} \ge 0$ for any (simple) oriented cycle~$\gamma$ in~$D_{\order{\b{F}}}$, where each arc~$\alpha \in \gamma$ has label~$i(\alpha)$, source~$s(\alpha)$, and target~$t(\alpha)$.
1866
\end{proposition}
1867
1868
\begin{proof}
1869
Consider an ordered $(\ell,n)$-partition forest~$\order{\b{F}} \eqdef (\order{F}_1, \dots, \order{F}_\ell)$.
1870
For each~$i \in [\ell]$, denote by
1871
\begin{itemize}
1872
\item $m_i$ the number of arcs of~$D_{\order{F}_i}$
1873
\item $M_i$ the incidence matrix of~$D_{\order{F}_i}$, with $m_i$ rows and $n$ columns, with a row for each arc~$\alpha$ of~$D_{\order{F}_i}$ containing a $-1$ in column~$s(\alpha)$, a $1$ in column~$t(\alpha)$, and $0$ elsewhere,
1874
\item $\b{z}_i$ the column vector in~$\R^{m_i}$ with a row for each arc~$\alpha$ of~$D_{\order{F}_i}$ containing the value~$A_{i(\alpha), s(\alpha), t(\alpha)}$.
1875
\end{itemize}
1876
Then a point~$\b{x} \in \R^n$ belongs to the face of the $i$\ordinal{} braid arrangement corresponding to~$\order{F}_i$ if and only if it satisfies~$M_i \, \b{x} \le \b{z}_i$.
1877
Hence, $\order{\b{F}}$ appears as a face of the $\b{a}$-braid arrangement if and only if there exists~$\b{x} \in \R^n$ such that~$M \, \b{x} \le \b{z}$, where~$M$ is the $(m \times n)$-matrix (where~$m \eqdef \sum_{i \in [\ell]} m_i$), obtained by piling the matrices~$M_i$ for~$i \in [\ell]$ and similarly, $\b{z}$ is the column vector obtained by piling the vectors~$\b{z}_i$.
1878
A direct application of the Farkas lemma (see \eg \cite[Prop.~1.7]{Ziegler}), there exists~$\b{x} \in \R^n$ such that~$M \, \b{x} \le \b{z}$ if and only if~$\b{c} \b{z} \ge 0$ for any~$\b{c} \in (\R^m)^*$ with~$\b{c} \ge \b{0}$ and~$\b{c} M = \b{0}$.
1879
Now it is classical that the left kernel of the incidence matrix of a directed graph is generated by its circuits (non-necessarily oriented cycles), and that the positive cone in this left kernel is generated by its oriented cycles.
1880
\end{proof}
1881
1882
\begin{remark}
1883
Note that we made some arbitrary choices here by choosing the arc from~$\max(\order{\pi}_j)$ to~$\min(\order{\pi}_{j+1})$ between two consecutive parts~$\order{\pi}_j$ and~$\order{\pi}_{j+1}$ and a cycle inside each part~$\order{\pi}_j$ (while we said that the order in each part is irrelevant).
1884
We could instead have considered all arcs connecting two elements of two consecutive parts, or two elements inside the same part.
1885
Our choices just limit the amount of oriented cycles in~$D_{\order{\pi}}$.
1886
\end{remark}
1887
1888
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1889
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1890
1891
\clearpage
1892
\part{Diagonals of permutahedra}
1893
\label{part:diagonalsPermutahedra}
1894
1895
In this second part, we study the combinatorics of the diagonals of the permutahedra.
1896
In \cref{sec:cellularDiagonals}, we first recall the definition and some known facts about cellular diagonals of polytopes (\cref{subsec:cellularDiagonalsPolytopes}), which we immediately specialize to the classical permutahedron (\cref{sec:cellularDiagonalsPermutahedra}), and connect to \cref{part:multiBraidArrangements} to derive enumerative statements on the diagonals of permutahedra (\cref{subsec:enumerationDiagonalPermutahedra}).
1897
In \cref{sec:operadicDiagonals}, we consider two particular diagonals, the $\LA$ and $\SU$ diagonals (\cref{subsec:LASUdiagonal}), we show that these are the only two operadic diagonals which induce the weak order (\cref{subsec:operadicProperty}), and that they are isomorphic (\cref{subsec:isos-LA-SU}).
1898
Using results from \cref{sec:facePoset,sec:cellularDiagonals}, we then characterize their facets in terms of paths in $(2,n)$-partition trees (\cref{subsec:facets-operadic-diags}), and their vertices as pattern-avoiding pairs of permutations (\cref{subsec:vertices-operadic-diags}).
1899
Finally, in \cref{sec:shifts}, we show that the geometric $\SU$ diagonal $\SUD$ is a topological enhancement of the original Saneblidze-Umble diagonal~\cite{SaneblidzeUmble} (\cref{subsec:topological-SU}).
1900
In order to prove this result, we define different types of shifts that can be performed on the facets of the $\SU$ diagonal, and give several new equivalent definitions of it.
1901
These descriptions are directly translated to the $\LA$ diagonal via isomorphism (\cref{subsec:shifts-under-iso}).
1902
Moreover, we observe that the shifts define a natural lattice structure on the set of facets of operadic diagonals, that we call the \emph{shift lattice} (\cref{sec:Shift-lattice}).
1903
Finally, we present the alternative matrix (\cref{subsec:matrix}) and cubical (\cref{sec:Cubical}) descriptions of the $\SU$ diagonal from \cite{SaneblidzeUmble,SaneblidzeUmble-comparingDiagonals}, provide proofs of their equivalence with the other descriptions, and give their $\LA$ counterparts.
1904
1905
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1906
1907
\section{Cellular diagonals}
1908
\label{sec:cellularDiagonals}
1909
1910
%%%%%%%%%%%%%%%
1911
1912
\subsection{Cellular diagonals for polytopes}
1913
\label{subsec:cellularDiagonalsPolytopes}
1914
1915
As discussed in the introduction, cellular approximations of the thin diagonal for families of polytopes are of fundamental importance in algebraic topology and geometry.
1916
They allow one to define the cup product and thus define the ring structure on the cohomology groups of a topological space, and combinatorially on the Chow groups of a toric variety.
1917
We now proceed to define thin, cellular, and geometric diagonals.
1918
1919
\begin{definition}
1920
\label{def:thinDiagonal}
1921
The \defn{thin diagonal} of a set $X$ is the map~$\delta : X \to X \times X$ defined by $\delta(x) \eqdef (x,x)$ for all $x \in X$.
1922
See \cref{fig:examplesDiagonals1}\,(left).
1923
\end{definition}
1924
1925
\begin{definition}
1926
\label{def:cellularDiagonal}
1927
A \defn{cellular diagonal} of a $d$-dimensional polytope $P$ is a continuous map ${\Delta : P \to P \times P}$ such that
1928
\begin{enumerate}
1929
\item its image is a union of $d$-dimensional faces of $P\times P$ (\ie it is \defn{cellular}),
1930
\item it agrees with the thin diagonal of~$P$ on the vertices of $P$, and
1931
\item it is homotopic to the thin diagonal of~$P$, relative to the image of the vertices of~$P$.
1932
\end{enumerate}
1933
See \cref{fig:examplesDiagonals1}\,(middle left).
1934
A cellular diagonal is said to be \defn{face coherent} if its restriction to a face of $P$ is itself a cellular diagonal for that face.
1935
\end{definition}
1936
1937
A powerful geometric technique to define face coherent cellular diagonals on polytopes first appeared in~\cite{FultonSturmfels}, was presented in~\cite{MasudaThomasTonksVallette}, and was fully developed in~\cite{LaplanteAnfossi}.
1938
We provide in \cref{thm:diagonal} the precise (but slightly technical) definition of these diagonals, even though we will only use the characterization of the faces in their image provided in \cref{thm:universalFormula}.
1939
1940
The key idea is that any vector $\b{v}$ in generic position with respect to $P$ defines a cellular diagonal of $P$.
1941
For $\b{z}$ a point of $P$, we denote by $\rho_{\b{z}} P \eqdef 2\b{z}-P$ the reflection of $P$ with respect to the point~$\b{z}$.
1942
1943
\begin{definition}
1944
The \defn{fundamental hyperplane arrangement}~$\mathcal{H}_P$ of a polytope~$P \subset \R^d$ is the set of all linear hyperplanes of~$\R^d$ orthogonal to the edges of $P\cap \rho_{\b{z}} P$ for all $\b{z} \in P$.
1945
See \cref{fig:examplesHyperplanes}.
1946
\end{definition}
1947
1948
\begin{figure}[p]
1949
\centerline{
1950
\includegraphics[scale=.38]{HypSimplex.png} \hspace{-.6cm}
1951
\includegraphics[scale=.44]{HypCube.png} \hspace{-.3cm}
1952
\includegraphics[scale=.38]{HypPermuto.png}
1953
}
1954
\caption{The fundamental hyperplane arrangements of the $3$-dimensional simplex (left), cube (middle), and permutahedron (right). The hyperplanes perpendicular
1955
to edges of some intersection $P\cap \rho_z P$, which are \emph{not} edges of the polytope~$P$, are colored in blue. Left and rightmost illustrations from~\cite[Fig.~12]{LaplanteAnfossi}.}
1956
\label{fig:examplesHyperplanes}
1957
\end{figure}
1958
1959
A vector is \defn{generic with respect to~$P$} if it does not belong to the union of the hyperplanes of the fundamental hyperplane arrangement~$\mathcal{H}_P$.
1960
In particular, such a vector is not perpendicular to any edge of $P$, and we denote by~$\min_{\b{v}}(P)$ (\resp $\max_{\b{v}}(P)$) the unique vertex of~$P$ which minimizes (\resp maximizes) the scalar product with~$\b{v}$.
1961
Note that the datum of a polytope~$P$ together with a vector~$\b{v}$ generic with respect to~$P$ was called \defn{positively oriented polytope} in~\cite{MasudaThomasTonksVallette,LaplanteAnfossi,LaplanteAnfossiMazuir}.
1962
1963
\pagebreak
1964
1965
\begin{theorem}
1966
\label{thm:diagonal}
1967
For any vector $\b{v} \in \R^d$ generic with respect to $P$, the tight coherent section~$\triangle_{(P,\b{v})}$ of the projection $P \times P \to P, (\b{x}, \b{y}) \mapsto (\b{x}+\b{y})/2$ selected by the vector~$(-\b{v}, \b{v})$ defines a cellular diagonal of~$P$.
1968
More precisely, $\triangle_{(P,\b{v})}$ is given by the formula
1969
\begin{align*}
1970
\begin{array}{rlcl}
1971
\triangle_{(P,\b{v})}\ : & P & \to & P\times P \\
1972
& \b{z} & \mapsto & \bigl( \min_{\b{v}}(P\cap \rho_{\b{z}} P), \, \max_{\b{v}}(P\cap \rho_{\b{z}} P) \bigr) .
1973
\end{array}
1974
\end{align*}
1975
\end{theorem}
1976
1977
\begin{definition}
1978
\label{def:geometricDiagonal}
1979
A \defn{geometric diagonal} of a polytope~$P$ is a diagonal of the form~$\triangle_{(P,\b{v})}$ for some vector~$\b{v} \in \R^d$ generic with respect to~$P$.
1980
\end{definition}
1981
1982
Note that the geometric diagonal~$\triangle_{(P,\b{v})}$ only depends on the region of~$\mathcal{H}_P$ containing~$\b{v}$, see~\cite[Prop.~1.23]{LaplanteAnfossi}.
1983
1984
\begin{figure}[p]
1985
\centerline{\scalebox{1.25}{
1986
\begin{tabular}{c@{\hspace{-.2cm}}c@{\hspace{-.2cm}}c@{\hspace{-.2cm}}c}
1987
\includegraphics[scale=1.2]{diagonalSegment1} &
1988
\includegraphics[scale=1.2]{diagonalSegment2} &
1989
\raisebox{-.2cm}{\includegraphics[scale=1]{diagonalSegment3}} &
1990
\raisebox{-.2cm}{\includegraphics[scale=1]{diagonalSegment4}}
1991
\\
1992
\includegraphics[scale=.9]{diagonalTriangle1} &
1993
\includegraphics[scale=.9]{diagonalTriangle2} &
1994
\includegraphics[scale=.6]{diagonalTriangle3} &
1995
\includegraphics[scale=.6]{diagonalTriangle4}
1996
\\
1997
\includegraphics[scale=.9]{diagonalSquarre1} &
1998
\includegraphics[scale=.9]{diagonalSquarre2} &
1999
\includegraphics[scale=.6]{diagonalSquarre3} &
2000
\includegraphics[scale=.6]{diagonalSquarre4}
2001
\end{tabular}
2002
}}
2003
\caption{Cellular diagonals of the segment (top), the triangle (middle) and the square (bottom). For each of them, we have represented the thin diagonal of~$P$ (left, in blue), a cellular diagonal of~$P$ (middle left, in red) both in~$P \times P$, the associated polytopal subdivision of~$P$ (middle right) and the common refinement of the two copies of the normal fan of~$P$ (right) both in~$P$.}
2004
\label{fig:examplesDiagonals1}
2005
\end{figure}
2006
2007
\begin{figure}
2008
\centerline{
2009
\includegraphics[scale=.3]{diagonalSimplexGuillaume.png}
2010
\includegraphics[scale=.35]{diagonalCubeGuillaume.png}
2011
\includegraphics[scale=.3]{diagonalPermutahedronGuillaume.png}
2012
}
2013
\caption{The subdivisions induced by cellular diagonals of the $3$-dimensional simplex (left), cube (middle), and permutahedron (right). Right illustration from~\cite[Fig.~13]{LaplanteAnfossi}.}
2014
\label{fig:examplesDiagonals2}
2015
\end{figure}
2016
2017
Now the following \defn{universal formula}~\cite[Thm.~1.26]{LaplanteAnfossi} expresses combinatorially the faces in the image of the geometric diagonal~$\triangle_{(P,\b{v})}$.
2018
Recall that the \defn{normal cone} of a face~$F$ of a polytope~$P$ in~$\R^d$ is the cone of directions~$\b{c} \in \R^d$ such that the maximum of the scalar product~$\dotprod{\b{c}}{\b{x}}$ over~$P$ is attained for some~$\b{x}$ in~$F$.
2019
2020
\begin{theorem}[{\cite[Thm.~1.26]{LaplanteAnfossi}}]
2021
\label{thm:universalFormula}
2022
Fix a vector $\b{v} \in \R^d$ generic with respect to $P$.
2023
For each hyperplane~$H$ of the fundamental hyperplane arrangement~$\mathcal{H}_P$, denote by~$H^{\b{v}}$ the open half space defined by~$H$ and containing~$\b{v}$.
2024
The faces of~$P \times P$ in the image of the geometric diagonal~$\triangle_{(P,\b{v})}$ are the faces~$F \times G$ where~$F$ and~$G$ are faces of~$P$ such that either the normal cone of~$F$ intersects~$H^{-\b{v}}$ or the normal cone of~$G$ intersects~$H^{\b{v}}$, for each~$H \in \mathcal{H}_P$.
2025
\end{theorem}
2026
2027
The image of $\triangle_{(P,\b{v})}$ is a union of pairs of faces $F \times G$ of the Cartesian product~$P \times P$.
2028
By drawing the polytopes~${(F+G)/2}$ for all pairs of faces $(F,G) \in \Ima \triangle_{(P,\b{v})}$, we can visualize~$\triangle_{(P,\b{v})}$ as a polytopal subdivision of $P$.
2029
See \cref{fig:examplesDiagonals1}\,(middle right) and \cref{fig:examplesDiagonals2}.
2030
2031
It turns out that the dual of this complex is just the common refinement of two translated copies of the normal fan of~$P$.
2032
See \cref{fig:examplesDiagonals1}\,(right).
2033
Recall that the \defn{normal fan} of~$P$ is the fan formed by the normal cones of all faces of~$P$.
2034
We thus obtain the following statement.
2035
2036
\begin{proposition}[{\cite[Coro.~1.4]{LaplanteAnfossi}}]
2037
\label{prop:diagonalCommonRefinement}
2038
The inclusion poset on the faces in the image of the diagonal~$\triangle_{(P,\b{v})}$ is isomorphic to the reverse inclusion poset on the faces of the common refinement of two copies of the normal fan of~$P$, translated from each other by the vector~$\b{v}$.
2039
\end{proposition}
2040
2041
Finally, the following statement relates the image of the diagonal~$\triangle_{(P, \b{v})}$ to the intervals of the poset obtained by orienting the skeleton of~$P$ in direction~$\b{v}$.
2042
2043
\begin{proposition}[{\cite[Prop. 1.17]{LaplanteAnfossi}}]
2044
\label{prop:magicalFormula}
2045
For any polytope $P$ and any generic vector $\b{v}$, we have
2046
\begin{equation}
2047
\label{eq:magique}
2048
\Ima\triangle_{(P, \b{v})} \quad \subseteq \bigcup_{\substack{F,G \text{ faces of } P \\ \max_{\b{v}}(F) \, \le \, \min_{\b{v}}(G)}} F \times G .
2049
\end{equation}
2050
\end{proposition}
2051
2052
\begin{remark}
2053
\label{rem:magicalFormula}
2054
For some polytopes such as the simplices~\cite{EilenbergMacLane}, the cubes~\cite{Serre}, the freehedra~\cite{Saneblidze-freeLoopFibration}, and the associahedra~\cite{MasudaThomasTonksVallette}, the reverse inclusion also holds (in the case of the simplices and the cubes, the diagonals are known as the \emph{Alexander--Whitney} map~\cite{EilenbergMacLane} and \emph{Serre} map~\cite{Serre}).
2055
According to~\cite{MasudaThomasTonksVallette}, the resulting equality enhancing~(\ref{eq:magique}) was called \defn{magical formula} by J.-L. Loday.
2056
This equality simplifies the computation of the $f$-vectors of the diagonals.
2057
For instance, the number of $k$-dimensional faces in the diagonal of the $(n-1)$-dimensional simplex, cube, and associahedron are respectively given by
2058
\begin{alignat*}{2}
2059
f_k(\triangle_{\Simplex}) & = (k+1) \binom{n+1}{k+2} && \text{\OEIS{A127717}},
2060
% choose k+2 points of [n] where 2 consecutive are distinguished and might be equal, and the rest is distinct
2061
% this is the same as choosing k+2 points of [n+1], and the position of the consecutive pair of distinguished points among them
2062
% hence (k+1) \binom{n+1}{k+2}
2063
\\
2064
f_k(\triangle_{\Cube}) & = \binom{n-1}{k} 2^k 3^{n-1-k} && \text{\OEIS{A038220},}
2065
% choose a < b <= c < d in the boolean lattice such that |b-a| + |d-c| = k.
2066
% choose the positions of the ones in (b-a) + (d-c) => \binom{n}{k}
2067
% choose whether each of this ones is in (b-a) or in (d-c) => 2^k
2068
% choose the values of b and c on the remaining n-k positions to be either 00, 01 or 11 => 3^(n-k)
2069
\\
2070
f_k(\triangle_{\Asso}) & = \frac{2}{(3n+1)(3n+2)} \binom{n-1}{k} \binom{4n+1-k}{n+1} \qquad && \text{\cite{BostanChyzakPilaud}.}
2071
\end{alignat*}
2072
Polytopes of greater complexity such as the multiplihedra~\cite{LaplanteAnfossiMazuir} or the operahedra~\cite{LaplanteAnfossi}, which include the permutahedra, do not possess this exceptional property, and the $f$-vectors of their diagonals are harder to compute.
2073
\end{remark}
2074
2075
\begin{remark}
2076
\label{rem:Fulton--Sturmfels}
2077
This is in fact precisely the content of the \emph{Fulton--Sturmfels formula} for the intersection product on toric varieties~\cite[Thm.~4.2]{FultonSturmfels}, where the definition of cellular diagonals as tight coherent sections first appeared.
2078
\end{remark}
2079
2080
To conclude, we define the opposite of a geometric diagonal.
2081
2082
\begin{definition}
2083
\label{def:oppositeDiagonal}
2084
The \defn{opposite} of the geometric diagonal~$\triangle \eqdef \triangle_{(P, \b{v})}$ for a vector~$\b{v} \in \R^d$ generic with respect to $P$ is the geometric diagonal~$\triangle^{\op} \eqdef \triangle_{(P, -\b{v})}$ for the vector~$-\b{v}$.
2085
Observe that
2086
\[{
2087
F \times G \in \Ima \triangle} \qquad\iff\qquad G \times F \in \Ima \triangle^{\op}.
2088
\]
2089
\end{definition}
2090
2091
%%%%%%%%%%%%%%%
2092
2093
\subsection{Cellular diagonals for the permutahedra}
2094
\label{sec:cellularDiagonalsPermutahedra}
2095
2096
We now specialize the statements of \cref{subsec:cellularDiagonalsPolytopes} to the standard permutahedron.
2097
We first recall its definition.
2098
2099
\begin{definition}
2100
\label{def:permutahedron}
2101
The \defn{permutahedron}~$\Perm$ is the polytope in~$\R^n$ defined equivalently as
2102
\begin{itemize}
2103
\item the convex hull of the points~$\sum_{i \in [n]} i \, \b{e}_{\sigma(i)}$ for all permutations~$\sigma$ of~$[n]$, see~\cite{Schoute},
2104
\item the intersection of the hyperplane~$\smash{\bigset{\b{x} \in \R^n}{\sum_{i \in I} x_i = \binom{n+1}{2}}}$ with the affine halfspaces $\smash{\bigset{\b{x} \in \R^n}{\sum_{i \in I} x_i \ge \binom{\card{I}+1}{2}}}$ for all~${\varnothing \ne I \subsetneq [n]}$, see~\cite{Rado}.
2105
\end{itemize}
2106
The normal fan of the permutahedron~$\Perm$ is the fan defined by the braid arrangement~$\braidArrangement$.
2107
In particular, the faces of~$\Perm$ correspond to the ordered partitions of~$[n]$.
2108
Moreover, when oriented in a generic direction, the skeleton of the permutahedron~$\Perm$ is isomorphic to the Hasse diagram of the classical weak order on permutations of~$[n]$.
2109
% when oriented in the direction~${\b{\omega} \eqdef (n,\dots,1) - (1,\dots,n) = \sum_{i \in [n]} (n+1-2i) \, \b{e}_i}$
2110
See \cref{fig:permutahedron}.
2111
%
2112
\begin{figure}
2113
\centerline{
2114
\includegraphics[scale=.8]{permutahedronWeakOrder}
2115
\qquad
2116
\includegraphics[scale=.7]{weakOrder}
2117
}
2118
\caption{The permutahedron~$\Perm[4]$ (left) and the weak order on permutations of~$[4]$ (right).}
2119
\label{fig:permutahedron}
2120
\end{figure}
2121
\end{definition}
2122
2123
The fundamental hyperplane arrangement of the permutahedron~$\Perm$ was described in~\cite[Sect.~3.1]{LaplanteAnfossi}.
2124
As we will use the following set throughout the paper, we embed it in a definition.
2125
2126
\begin{definition}
2127
\label{def:Un}
2128
For~$n \in \N$, we define
2129
\[
2130
\Un(n) \eqdef \bigset{\{I,J\}}{I,J \subset [n] \text{ with } \card{I}=\card{J} \text{ and } I \cap J = \varnothing}.
2131
\]
2132
An \defn{ordering} of~$\Un(n)$ is a set containing exactly one of the two ordered pairs $(I,J)$ or $(J,I)$ for each $\{I,J\} \in \Un(n)$.
2133
\end{definition}
2134
2135
\begin{proposition}[{\cite[Sect.~3.1]{LaplanteAnfossi}}]
2136
The fundamental hyperplane arrangement of the permutahedron~$\Perm$ is given by the hyperplanes~$\bigset{\b{x} \in \R^n}{\sum\limits_{i \in I} x_i = \sum\limits_{j \in J} x_j}$ for all~$\{I,J\} \in \Un(n)$.
2137
\end{proposition}
2138
2139
For a vector~$\b{v}$ generic with respect to~$\Perm$, we denote by~$\Or(\b{v})$ the ordering of~$\Un(n)$ such that~$\sum_{i \in I} v_i > \sum_{j \in J} v_j$ for all~$(I,J) \in \Or(\b{v})$.
2140
Applying \cref{thm:universalFormula}, we next describe the faces in the image of the geometric diagonal~$\triangle_{(\Perm, \b{v})}$.
2141
For this, the following definition will be convenient.
2142
2143
\begin{definition}
2144
\label{def:domination}
2145
For $I,J \subseteq [n]$ and an ordered partition $\sigma$ of~$[n]$ into $k$ blocks, we say that $I$~\defn{dominates} $J$ in~$\sigma$ if for all $\ell \in [k]$, the first~$\ell$ blocks of~$\sigma$ contain at least as many elements of~$I$ than of~$J$.
2146
\end{definition}
2147
2148
\begin{theorem}[{\cite[Thm.~3.16]{LaplanteAnfossi}}]
2149
\label{thm:IJ-description}
2150
A pair~$(\sigma, \tau)$ of ordered partitions of~$[n]$ corresponds to a face in the image of the geometric diagonal~$\triangle_{(\Perm, \b{v})}$ if and only if, for all~$(I,J) \in \Or(\b{v})$, $J$ does not dominate~$I$ in~$\sigma$ or $I$ does not dominate~$J$ in~$\tau$.
2151
\end{theorem}
2152
2153
As the normal fan of the permutahedron~$\Perm$ is the fan defined by the braid arrangement~$\braidArrangement$, we obtain by \cref{prop:diagonalCommonRefinement} the following connection between the diagonal of~$\Perm$ and the $(2,n)$-braid arrangement~$\multiBraidArrangement[2][n]$ studied in \cref{part:multiBraidArrangements}.
2154
This connection is illustrated in \cref{fig:diagonalPermutahedron1}.
2155
%
2156
\begin{figure}
2157
\centerline{\includegraphics[scale=.7]{diagonalPermutahedron1}}
2158
\caption{The duality between the $(2,3)$-braid arrangement~$\multiBraidArrangement[3][2]$ (left) and the diagonal of the permutahedron~$\Perm[3]$ (right).}
2159
\label{fig:diagonalPermutahedron1}
2160
\end{figure}
2161
2162
\begin{proposition}
2163
\label{prop:diagonalPermutahedraMultiBraidArrangements}
2164
The inclusion poset on the faces in the image of the diagonal~$\triangle_{(\Perm, \b{v})}$ is isomorphic to the reverse inclusion poset on the faces of the $(2,n)$-braid arrangement~$\multiBraidArrangement[2][n]$.
2165
\end{proposition}
2166
2167
\begin{remark}
2168
\label{rem:translationMatrix}
2169
To be more precise, the diagonal~$\triangle_{(\Perm, \b{v})}$ is dual to the $\b{a}$-braid arrangement~$\multiBraidArrangement[2][n](\b{a})$ where the translation matrix~$\b{a}$ has two rows, with first row~${\b{a}_{1,j} = 0}$ and second row~$\b{a}_{2,j} = v_j-v_{j+1}$ for all~$j \in [n-1]$.
2170
Since~$\b{v}$ is generic with respect to~$\Perm$, we have~$\sum_{i \in I} v_i \ne \sum_{j \in J} v_j$ for all~$(I,J) \in \Un(n)$, so that~$\b{a}$ is indeed generic.
2171
\end{remark}
2172
2173
\begin{remark}
2174
Similarly, the combinatorics of the $(\ell-1)$\ordinalst{} iteration of a diagonal of the permutahedron~$\Perm$ is given by the combinatorics of the $(\ell,n)$-braid arrangement.
2175
\end{remark}
2176
2177
Finally, the permutahedron~$\Perm$ is not magical in the sense of \cref{prop:magicalFormula}.
2178
See also \cref{exm:intervalsWeakOrderNotDiagonals}.
2179
2180
\begin{proposition}[{\cite[Sect.~3]{LaplanteAnfossi}}]
2181
For any~$n > 3$ and any generic vector~$\b{v}$, the diagonal~$\triangle_{(\Perm, \b{v})}$ of the permutahedron~$\Perm$ does \emph{not} satisfy the magical formula.
2182
Namely, we have the strict inclusion
2183
\begin{equation*}
2184
\Ima\triangle_{(\Perm, \b{v})} \quad \subsetneq \bigcup_{\substack{F,G \text{ faces of } \Perm \\ \max_{\b{v}}(F) \, \le \, \min_{\b{v}}(G)}} F \times G .
2185
\end{equation*}
2186
\end{proposition}
2187
2188
\begin{remark}
2189
\label{rem:oppositeDiagonals}
2190
Note that opposite diagonals have opposite orderings.
2191
Namely, $\Or(-\b{v}) = \Or(\b{v})^{\op}$ where~$\Or^{\op} \eqdef \set{(J,I)}{(I,J) \in \Or}$.
2192
\end{remark}
2193
2194
%%%%%%%%%%%%%%%
2195
2196
\subsection{Enumerative results on cellular diagonals of the permutahedra}
2197
\label{subsec:enumerationDiagonalPermutahedra}
2198
2199
Relying on \cref{prop:diagonalPermutahedraMultiBraidArrangements}, we now specialize the results of \cref{part:multiBraidArrangements} to the case~$\ell = 2$ to derive enumerative results on the diagonals of the permutahedra.
2200
2201
Observe that one can easily compute the full M\"obius polynomials of the $(2,n)$-braid arrangements~$\multiBraidArrangement[n][2]$ from \cref{thm:MobiusPolynomialMultiBraidArrangement}:
2202
\begin{align*}
2203
\allowdisplaybreaks
2204
\mobPol[{\multiBraidArrangement[1][2]}] & = 1 , \\
2205
\mobPol[{\multiBraidArrangement[2][2]}] & = x y - 2 x + 2 , \\
2206
\mobPol[{\multiBraidArrangement[3][2]}] & = x^2 y^2 - 6 x^2 y + 10 x^2 + 6 x y - 18 x + 8 , \\
2207
\mobPol[{\multiBraidArrangement[4][2]}] & = x^3 y^3 - 12 x^3 y^2 + 52 x^3 y - 84 x^3 + 12 x^2 y^2 - 96 x^2 y + 216 x^2 + 44 x y - 182 x + 50 , \\
2208
\mobPol[{\multiBraidArrangement[5][2]}] & = x^4 y^4 - 20 x^4 y^3 + 160 x^4 y^2 - 620 x^4 y + 1008 x^4 \\ & \quad+ 20 x^3 y^3 - 300 x^3 y^2 + 1640 x^3 y - 3360 x^3 \\ & \quad+ 140 x^2 y^2 - 1430 x^2 y + 4130 x^2 + 410 x y - 2210 x + 432 .
2209
\end{align*}
2210
%which can be encoded in matrices as
2211
%{\small
2212
%\[
2213
%\left(\begin{array}{r}
2214
%1
2215
%\end{array}\right)
2216
%\left(\begin{array}{rr}
2217
%2 & -2 \\
2218
%0 & 1
2219
%\end{array}\right)
2220
%\left(\begin{array}{rrr}
2221
%8 & -18 & 10 \\
2222
%0 & 6 & -6 \\
2223
%0 & 0 & 1
2224
%\end{array}\right)
2225
%\left(\begin{array}{rrrr}
2226
%50 & -182 & 216 & -84 \\
2227
%0 & 44 & -96 & 52 \\
2228
%0 & 0 & 12 & -12 \\
2229
%0 & 0 & 0 & 1
2230
%\end{array}\right)
2231
%\left(\begin{array}{rrrrr}
2232
%432 & -2210 & 4130 & -3360 & 1008 \\
2233
%0 & 410 & -1430 & 1640 & -620 \\
2234
%0 & 0 & 140 & -300 & 160 \\
2235
%0 & 0 & 0 & 20 & -20 \\
2236
%0 & 0 & 0 & 0 & 1
2237
%\end{array}\right)
2238
%%\left(\begin{array}{r}
2239
%%1
2240
%%\end{array}\right)
2241
%%\left(\begin{array}{rr}
2242
%%2 & 0 \\
2243
%%-2 & 1
2244
%%\end{array}\right)
2245
%%\left(\begin{array}{rrr}
2246
%%8 & 0 & 0 \\
2247
%%-18 & 6 & 0 \\
2248
%%10 & -6 & 1
2249
%%\end{array}\right)
2250
%%\left(\begin{array}{rrrr}
2251
%%50 & 0 & 0 & 0 \\
2252
%%-182 & 44 & 0 & 0 \\
2253
%%216 & -96 & 12 & 0 \\
2254
%%-84 & 52 & -12 & 1
2255
%%\end{array}\right)
2256
%%\left(\begin{array}{rrrrr}
2257
%%432 & 0 & 0 & 0 & 0 \\
2258
%%-2210 & 410 & 0 & 0 & 0 \\
2259
%%4130 & -1430 & 140 & 0 & 0 \\
2260
%%-3360 & 1640 & -300 & 20 & 0 \\
2261
%%1008 & -620 & 160 & -20 & 1
2262
%%\end{array}\right)
2263
%%\left(\begin{array}{rrrrrr}
2264
%%4802 & 0 & 0 & 0 & 0 & 0 \\
2265
%%-31922 & 4732 & 0 & 0 & 0 & 0 \\
2266
%%82560 & -23100 & 1830 & 0 & 0 & 0 \\
2267
%%-104400 & 41560 & -6210 & 340 & 0 & 0 \\
2268
%%64800 & -32760 & 6960 & -720 & 30 & 0 \\
2269
%%-15840 & 9568 & -2580 & 380 & -30 & 1
2270
%%\end{array}\right)
2271
%\]
2272
%}
2273
2274
We now focus on the number of vertices, regions, and bounded regions of the $(2,n)$-braid arrangement~$\multiBraidArrangement[n][2]$, to obtain the number of facets, vertices, and internal vertices of the diagonal of the permutahedron~$\Perm$.
2275
The first few values are gathered in \cref{table:enumerationDiagonalPermutahedra1,table:enumerationDiagonalPermutahedra2}.
2276
2277
\afterpage{
2278
\begin{table}
2279
\centerline{\scalebox{1}{
2280
\begin{tabular}[t]{c|ccccccccccc}
2281
$n$ & $1$ & $2$ & $3$ & $4$ & $5$ & $6$ & $7$ & $8$ & $9$ & $\dots$ & OEIS \\
2282
\hline
2283
facets & $1$ & $2$ & $8$ & $50$ & $432$ & $4802$ & $65536$ & $1062882$ & $20000000$ & $\dots$ & \OEIS{A007334} \\
2284
vertices & $1$ & $3$ & $17$ & $149$ & $1809$ & $28399$ & $550297$ & $12732873$ & $343231361$ & $\dots$ & \OEIS{A213507} \\
2285
int. verts. & $1$ & $1$ & $5$ & $43$ & $529$ & $8501$ & $169021$ & $4010455$ & $110676833$ & $\dots$ & \OEIS{A251568}
2286
\end{tabular}
2287
}}
2288
% \vspace{.3cm}
2289
\caption{The numbers of facets, vertices, and internal vertices of the diagonal of the permutahedron~$\Perm$ for~$n \in [9]$.}
2290
\label{table:enumerationDiagonalPermutahedra1}
2291
\end{table}
2292
}
2293
2294
\afterpage{
2295
\begin{table}
2296
\centerline{
2297
\begin{tabular}{c@{\quad}c@{\quad}c@{\quad}c}
2298
$n = 1$ & $n = 2$ & $n = 3$ & $n = 4$
2299
\\
2300
\begin{tabular}[t]{r|c}
2301
\textbf{dim} & \textbf{0} \\
2302
\hline
2303
\textbf{0} & 1
2304
\end{tabular}
2305
&
2306
\begin{tabular}[t]{r|cc}
2307
\textbf{dim} & \textbf{0} & \textbf{1} \\
2308
\hline
2309
\textbf{0} & 3 & 1 \\
2310
\textbf{1} & 1 &
2311
\end{tabular}
2312
&
2313
\begin{tabular}[t]{r|ccc}
2314
\textbf{dim} & \textbf{0} & \textbf{1} & \textbf{2} \\
2315
\hline
2316
\textbf{0} & 17 & 12 & 1 \\
2317
\textbf{1} & 12 & 6 & \\
2318
\textbf{2} & 1 & &
2319
\end{tabular}
2320
&
2321
\begin{tabular}[t]{r|cccc}
2322
\textbf{dim} & \textbf{0} & \textbf{1} & \textbf{2} & \textbf{3} \\
2323
\hline
2324
\textbf{0} & 149 & 162 & 38 & 1 \\
2325
\textbf{1} & 162 & 150 & 24 & \\
2326
\textbf{2} & 38 & 24 & & \\
2327
\textbf{3} & 1 & & &
2328
\end{tabular}
2329
\end{tabular}
2330
}
2331
\vspace{.3cm}
2332
\centerline{
2333
\begin{tabular}{c@{\quad}c}
2334
$n = 5$ & $n = 6$
2335
\\
2336
\begin{tabular}[t]{r|ccccc}
2337
\textbf{dim} & \textbf{0} & \textbf{1} & \textbf{2} & \textbf{3} & \textbf{4} \\
2338
\hline
2339
\textbf{0} & 1809 & 2660 & 1080 & 110 & 1 \\
2340
\textbf{1} & 2660 & 3540 & 1200 & 80 & \\
2341
\textbf{2} & 1080 & 1200 & 270 & & \\
2342
\textbf{3} & 110 & 80 & && \\
2343
\textbf{4} & 1 & & & &
2344
\end{tabular}
2345
&
2346
\begin{tabular}[t]{r|cccccc}
2347
\textbf{dim} & \textbf{0} & \textbf{1} & \textbf{2} & \textbf{3} & \textbf{4} & \textbf{5} \\
2348
\hline
2349
\textbf{0} & 28399 & 52635 & 30820 & 6165 & 302 & 1 \\
2350
\textbf{1} & 52635 & 90870 & 67580 & 7785 & 240 & \\
2351
\textbf{2} & 30820 & 47580 & 20480 & 2160 & & \\
2352
\textbf{3} & 6165 & 7785 & 2160 & && \\
2353
\textbf{4} & 302 & 240 & & &&\\
2354
\textbf{5} & 1 & & & &&
2355
\end{tabular}
2356
\end{tabular}
2357
}
2358
\caption{Number of pairs of faces in the cellular image of the diagonal of the permutahedron~$\Perm$ for~$n \in [6]$.}
2359
\label{table:enumerationDiagonalPermutahedra2}
2360
\end{table}
2361
}
2362
2363
\begin{corollary}
2364
\label{coro:enumerationDiagonalPermutahedra}
2365
The diagonal of the permutahedron~$\Perm$ has
2366
\begin{itemize}
2367
\item $2 (n + 1)^{n-2}$ facets,
2368
\item $n \binom{n-1}{k_1} (n-k_1)^{k_1-1} (n-k_2)^{k_2-1}$ facets corresponding to pairs~$(F_1, F_2)$ of faces of the permutahedron~$\Perm$ with~$\dim(F_1) = k_1$ and~$\dim(F_2) = k_2$ (thus~$k_1 + k_2 = n-1$),
2369
\item $\displaystyle n! \, [z^n] \exp \bigg( \sum_{m \ge 1} \frac{C_m \, z^m}{m} \bigg)$ vertices,
2370
\item $\displaystyle (n-1)! \, [z^{n-1}] \exp \bigg( \sum_{m \ge 1} C_m \, z^m \bigg)$ internal vertices,
2371
\end{itemize}
2372
where~$\displaystyle C_m \eqdef \frac{1}{m+1} \binom{2m}{m}$ denotes the $m$\ordinal{} Catalan number.
2373
\end{corollary}
2374
2375
\begin{proof}
2376
Use the duality between the $(2,n)$-braid arrangement~$\multiBraidArrangement[n][2]$ and the diagonal of the permutahedron~$\Perm$ (see \cref{prop:diagonalPermutahedraMultiBraidArrangements,fig:diagonalPermutahedron1}), and specialize \cref{thm:verticesMultiBraidArrangement,thm:verticesRefinedMultiBraidArrangement,thm:regionsMultiBraidArrangement} to the case~$\ell = 2$.
2377
\end{proof}
2378
2379
\begin{remark}
2380
For completeness, we provide an alternative simpler proof of the first point of \cref{coro:enumerationDiagonalPermutahedra}.
2381
By \cref{prop:flatPosetMultiBraidArrangement}, we just need to count the $(2,n)$-partition trees.
2382
Consider a $(2,n)$-partition tree~$\b{F} \eqdef (F_1,F_2)$ (hence~$\card{F_1} + \card{F_2} = n + 1$).
2383
Consider the intersection tree~$T$ of~$\b{F}$ with vertices labeled by the parts of~$F_1$ and of~$F_2$ and edges labeled by~$[n]$, root~$T$ at the part of~$F_1$ containing vertex~$1$, forget the vertex labels of~$T$, and send each edge label of~$T$ to the next vertex away from the root, and label the root by~$0$.
2384
See \cref{fig:tree}.
2385
The result is a spanning tree of the complete graph~$K_{n+1}$ on~$\{0, \dots, n\}$ which must contain the edge~$(0,1)$ (because we have chosen the root to be the part of~$F_1$ containing~$1$).
2386
%
2387
\afterpage{
2388
\begin{figure}
2389
\centerline{\includegraphics[scale=.9]{tree}}
2390
\caption{The bijection from rooted $(\ell,n)$-partition trees (left) to spanning trees of~$K_{n+1}$ containing the edge~$(0,1)$ (right).}
2391
\label{fig:tree}
2392
\end{figure}
2393
}
2394
%
2395
Finally, by double counting the pairs~$(T,e)$ where $T$ is a spanning tree of~$K_{n+1}$ and $e$ is an edge of~$T$, we see that $n$ times the number of spanning trees of~$K_{n+1}$ equals $\binom{n+1}{2}$ times the number of spanning trees of~$K_{n+1}$ containing~$(0,1)$.
2396
Hence, by Cayley's formula for spanning trees of~$K_{n+1}$, we obtain that the number of $(2,n)$-partition trees~is
2397
\[
2398
\frac{2n}{n(n+1)} (n+1)^{n-1} = 2 (n + 1)^{n-2}.
2399
\]
2400
\end{remark}
2401
2402
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2403
2404
\pagebreak
2405
\section{Operadic diagonals}
2406
\label{sec:operadicDiagonals}
2407
2408
This section is devoted to the combinatorics of two specific diagonals of the permutahedra, the $\LA$ and $\SU$ diagonals, which are shown to be the only operadic geometric diagonals of the permutahedra.
2409
2410
%%%%%%%%%%%%%%%
2411
2412
\subsection{The $\LA$ and $\SU$ diagonals}
2413
\label{subsec:LASUdiagonal}
2414
2415
Recall from \cref{sec:cellularDiagonalsPermutahedra} that a geometric diagonal of the permutahedron~$\Perm$ gives a choice of ordering of the sets~$\Un(n)$ (see \cref{def:Un}).
2416
As we consider in this section diagonals for all permutahedra, we now consider families of orderings.
2417
2418
\begin{definition}
2419
An \defn{ordering} of $\Un \eqdef \{\Un(n)\}_{n \ge 1}$ is a family~$\Or \eqdef \{\Or(n)\}_{n \ge 1}$ where~$\Or(n)$ is an ordering of~$\Un(n)$ for each~$n \ge 1$.
2420
\end{definition}
2421
2422
We will be focusing on the following two orderings and their corresponding diagonals.
2423
2424
\begin{definition}
2425
The \defn{$\LA$} and~\defn{$\SU$ orderings} are defined by
2426
\begin{itemize}
2427
\item $\LA(n) \eqdef \set{(I,J)}{\{I,J\} \in \Un(n) \text{ and } \min(I\cup J) = \min I}$ and
2428
\item $\SU(n) \eqdef \set{(I,J)}{\{I,J\} \in \Un(n) \text{ and } \max(I\cup J) = \max J}$.
2429
\end{itemize}
2430
\end{definition}
2431
2432
\begin{definition}
2433
\label{def:LA-and-SU}
2434
The \defn{$\LA$ diagonal} $\LAD$ (\resp \defn{$\SU$ diagonal} $\SUD$) of the permutahedron~$\Perm$ is the geometric diagonal~$\triangle_{(\Perm,\b{v})}$ given by any vector $\b{v} \in \R^n$ satisfying
2435
\[
2436
\sum_{i \in I} v_i > \sum_{j \in J} v_j,
2437
\]
2438
for all~$(I,J) \in \LA(n)$ (\resp $(I,J) \in \SU(n)$).
2439
\end{definition}
2440
2441
First note that this definition makes sense, since vectors $\b{v}=(v_1,\ldots,v_n)$ such that $\Or(\b{v})$ is the $\LA$ or $\SU$ order do exist: take for instance $v_i\eqdef 2^{-i+1}$ for $\LAD$ and $v_i \eqdef 2^n - 2^{i-1}$ for $\SUD$.
2442
Second, note that the $\LA$ and $\SU$ diagonals coincide up to dimension $2$, but differ in dimension~$\ge 3$.
2443
The former is illustrated in \cref{fig:LUSAdiagonals}, with the faces labeled by ordered $(2,3)$-partition forests.
2444
2445
\begin{figure}
2446
\centerline{\includegraphics[scale=.85]{diagonalPermutahedron3}}
2447
\caption{The $\LA$ (and~$\SU$) diagonal of~$\Perm[3]$ with faces labeled by ordered $(2,3)$-partition forests. (See also \cref{fig:B23a} for the dual hyperplane arrangement.)}
2448
\label{fig:LUSAdiagonals}
2449
\end{figure}
2450
2451
We will prove in \cref{thm:recover-SU} that $\SUD$ is a topological enhancement of the Saneblidze--Umble diagonal from~\cite{SaneblidzeUmble}.
2452
The faces of the $\LA$ and $\SU$ diagonals are described by the following specialization of \cref{thm:IJ-description}.
2453
2454
\begin{theorem}
2455
\label{thm:minimal}
2456
A pair $(\sigma,\tau)$ of ordered partitions of $[n]$ is not a face of the $\LA$ diagonal~$\LAD$ if and only if there exists~$(I,J) \in \LA(n)$ such that~$J$ dominates~$I$ in~$\sigma$ and~$I$ dominates~$J$ in~$\tau$.
2457
%\[
2458
%(\sigma,\tau) \notin \LAD \iff \text{there is } (I,J) \in \LA(n) \text{ such that } J \text{ dominates } I \text{ in } \sigma \text{ and } I \text{ dominates } J \text{ in } \tau.
2459
%\]
2460
The same holds for the $\SU$ diagonal by replacing $\LAD$ by $\SUD$ and $\LA(n)$ by $\SU(n)$.
2461
\end{theorem}
2462
2463
%%%%%%%%%%%%%%%
2464
2465
\subsection{The operadic property}
2466
\label{subsec:operadicProperty}
2467
2468
The goal of this section is to prove that the $\LA$ and $\SU$ diagonals are the only two operadic diagonals which induce the weak order on the vertices of the permutahedra (\cref{thm:unique-operadic}).
2469
We start by properly defining operadic diagonals.
2470
2471
Let~$A \sqcup B = [n]$ be a partition of~$[n]$ where~$A \eqdef \{a_1,\dots,a_p\}$ and~$B \eqdef \{b_1,\dots,b_q\}$.
2472
Recall that the ordered partition $B | A$ corresponds to a facet of the permutahedron~$\Perm$ defined by the inequality~$\sum_{b \in B} x_b \ge \binom{\card{B}+1}{2}$.
2473
This facet is isomorphic to the Cartesian product
2474
\[
2475
(\Perm[p] +q \one_{[p]}) \times \Perm[q]
2476
\]
2477
of lower dimensional permutahedra, where the first factor is translated by~$q \one_{[p]} \eqdef q \sum_{i \in [p]} \b{e}_i$, via the permutation of coordinates
2478
\begin{equation*}
2479
\begin{matrix}
2480
\Theta & : & \R^{p} \times \R^{q} & \overset{\cong}{\longrightarrow} & \R^{n} \\
2481
& & (x_1,\ldots,x_p) \times (x_{p+1}, \ldots, x_{n}) & \longmapsto & (x_{\sigma^{-1}(1)},\ldots,x_{\sigma^{-1}(n)}) \ ,
2482
\end{matrix}
2483
\end{equation*}
2484
where $\sigma$ is the $(p,q)$-shuffle defined by $\sigma(i) \eqdef a_i$ for~$i \in [p]$, and $\sigma(p+j)\eqdef b_j$ for~$j \in [q]$.
2485
Note that this map is a particular instance of the eponym map introduced in Point (5) of~\cite[Prop.~2.3]{LaplanteAnfossi}.
2486
2487
\begin{definition}
2488
\label{def:operadicDiagonal}
2489
A family of geometric diagonals $\triangle \eqdef \{\triangle_n : \Perm \to \Perm \times \Perm\}_{n\geq 1}$ of the permutahedra is \defn{operadic} if for every face $A_1 | \ldots | A_k$ of the permutahedron $\Perm[\card{A_1}+\cdots + \card{A_k}]$, the map $\Theta$ induces a topological cellular isomorphism
2490
\[
2491
\triangle_{\card{A_1}} \times \ldots \times \triangle_{\card{A_k}} \cong \triangle_{\card{A_1} + \ldots + \card{A_k}} .
2492
\]
2493
\end{definition}
2494
2495
In other words, we require that the diagonal $\triangle$ commutes with the map $\Theta$, see~\cite[Sect.~4.2]{LaplanteAnfossi}.
2496
At the algebraic level, this property is called ``comultiplicativity" in~\cite{SaneblidzeUmble}.
2497
Note that in particular, such an isomorphism respects the poset structures.
2498
2499
We will now translate this operadic property for geometric diagonals to the corresponding orderings~$\Or$ of~$\Un$.
2500
We need the following standardization map (this map is classical for permutations or words, but we use it here for pairs of sets).
2501
2502
\begin{definition}
2503
The \defn{standardization} of a pair~$(I,J)$ of disjoint subsets of~$[n]$ is the only partition~$\std(I,J)$ of~$[\card{I}+\card{J}]$ where the relative order of the elements is the same as in~$(I,J)$.
2504
More precisely, it is defined recursively by
2505
\begin{itemize}
2506
\item $\std(\varnothing, \varnothing) \eqdef (\varnothing, \varnothing)$, and
2507
\item if~$k \eqdef \card{I}+\card{J}$ and~$\ell \eqdef \max(I \cup J)$ belongs to~$I$ (\resp $J$), then~$\std(I,J) = (U \cup \{k\}, V)$ (\resp $\std(I,J) \! = \! (U, V \cup \{k\})$) where~$(U,V) \! = \! \std(I \ssm \{\ell\}, J)$ (\resp ${(U,V) \! = \! \std(I, J \ssm \{\ell\})}$).
2508
\end{itemize}
2509
\end{definition}
2510
2511
For example, $\std(\{5,9,10\},\{6,8,12\}) = (\{1,4,5\},\{2,3,6\})$.
2512
2513
\begin{definition}
2514
\label{def:operadicOrdering}
2515
An ordering $\Or$ of $\Un$ is \defn{operadic} if every $\{I,J\} \in \Un$ satisfies the following two conditions
2516
\begin{enumerate}
2517
\item $\std(I,J) \in \Or$ implies~$(I,J) \in \Or$,
2518
\item if there exist~$I'\subset I, J'\subset J$ such that $(I',J') \in \Or$ and $(I\ssm I',J\ssm J') \in \Or$, then~$(I,J) \in \Or$.
2519
\end{enumerate}
2520
We call \defn{indecomposables} of $\Or$ the pairs $(I,J) \in \Or$ for which the only subpair $(I',J')$ satisfying Condition (2) above is the pair $(I,J)$ itself.
2521
\end{definition}
2522
2523
\begin{proposition}
2524
\label{prop:equiv-operadic}
2525
A geometric diagonal~$\triangle \eqdef \bigl( \triangle_{(\Perm, \b{v}_n)} \bigr)_{n \in \N}$ of the permutahedra is operadic if and only if its associated ordering~$\Or \eqdef \bigl( \Or(\b{v}_n) \bigr)_{n \in \N}$ is operadic.
2526
\end{proposition}
2527
2528
\begin{proof}
2529
Every operadic diagonal satisfies~\cite[Prop.~4.14]{LaplanteAnfossi}, which amounts precisely to an operadic ordering of $\Un$ in the sense of \cref{def:operadicOrdering}.
2530
\end{proof}
2531
2532
We now turn to the study of the $\LA$ and $\SU$ orderings.
2533
2534
\begin{lemma}
2535
\label{lem:operadic-ordering}
2536
The $\LA$ and $\SU$ orderings are operadic, and their indecomposables are the pairs whose standardization are $(\{1\},\{2\})$ and
2537
\begin{align}
2538
(\{1,k+2,k+3,\dots,2k-1,2k\}, \{2,3,\dots,k+1\}) \tag{$\LA$} \label{eq:std-LA} \\
2539
(\{k,k+1,\dots,2k-1\},\{1,2,3,\dots,k-1,2k\}) \tag{$\SU$}
2540
\end{align}
2541
for all~$k\geq 1$.
2542
The opposite orderings $\LA^{op}$ and $\SU^{op}$ are also operadic, and generated by the opposite pairs.
2543
\end{lemma}
2544
2545
\begin{proof}
2546
We present the proof for the $\LA$ ordering, the proofs for the $\SU$ and opposite orders are similar.
2547
First, we verify that $\LA$ is operadic.
2548
Condition (1) follows from the fact that standardizing a pair preserves its minimal element.
2549
Condition (2) also holds, since whenever $(I',J')$ and its complement are in $\LA$, we have $\min(I)=\min\{\min(I'),\min(I\ssm I')\}$, and thus $(I,J)$ itself is in~$\LA$.
2550
2551
Second, we compute the indecomposables.
2552
Let $(I,J)$ be a pair with standardization (\ref{eq:std-LA}).
2553
If we try to decompose $(I,J)$ as a non-trivial union, there is always one pair $(I',J')$ in this union for which~$\min (I\cup J) \notin I'$, so we have $\min ( I' \cup J') = \min J'$, which implies that $(I',J') \notin \LA$.
2554
Thus, the pair~$(I,J)$ is indecomposable.
2555
2556
It remains to show that any pair $(I,J)$ in $\LA$ whose standardization is not of the form (\ref{eq:std-LA}) can be decomposed as a union of such pairs.
2557
Let us denote by $(I_k,J_k)$ the standard form (\ref{eq:std-LA}) and let $(I,J)\in \LA$ be such that $\std(I,J) \neq (I_k,J_k)$.
2558
Then there exists $i_2 \in I\ssm \min I$ such that $i_2 < \max J$.
2559
This means that $(I,J)$ can be decomposed as a union: if we write it as $(\{i_1,\dots,i_k\},\{j_1,\dots,j_k\})$, where each set ordered smallest to largest, then we must have $\min (I\cup J)=i_1<i_2<j_k$, in which case $(\{i_2\},\{j_k\})$ and $(\{i_1,i_3,\dots,i_k\},\{j_1,\dots,j_{k-1}\})$ are both smaller $\LA$ pairs.
2560
Then it must be the case that $\std((\{i_2\},\{j_k\})) = (\{1\},\{2\})$, and $\std((\{i_1,i_3,\dots,i_k\},\{j_1,\dots,j_{k-1}\}))$ is either $(I_{k-1},J_{k-1})$, or we can repeat this decomposition.
2561
%This process must eventually terminate with the right-hand side reducing to $(I_l,J_l)$ for some $1 \leq l \leq k-1$.
2562
%In other words, any ${(I,J) = (\{i_1,\dots,i_k\}, \{j_1,\dots,j_k\})}$ decomposes as
2563
%\begin{align*}
2564
% (I,J) = (\{i_2\},\{j_k\}) \sqcup (\{i_3\},\{j_{k-1}\}) \sqcup \dots \sqcup (\{i_{l+1} \},\{j_{k-l-1} \}) \sqcup (I',J')
2565
%\end{align*}
2566
%where $\std((I',J')) = (I_l,J_l)$, and $1\leq l \leq k$.
2567
\end{proof}
2568
2569
\begin{remark}
2570
We note that the decomposition in \cref{lem:operadic-ordering} is one of potentially many different decompositions of the pair $(I,J)$.
2571
However, by definition of the $\LA$ order, for any decomposition $(I,J) = (\bigsqcup_{a\in A} I_a, \bigsqcup_{a \in A} J_a)$, we have $\std(I_a, J_a) \in \LA$ for all~$a \in A$.
2572
As such, all decompositions of a pair $(I,J)$, order it the same way.
2573
\end{remark}
2574
2575
\begin{proposition}
2576
\label{prop:operadicOrdering}
2577
The only operadic orderings of $\Un=\{\Un(n)\}_{n\geq 1}$ are the $\LA,\SU,\LA^{\op}$ and $\SU^{\op}$ orderings.
2578
\end{proposition}
2579
2580
\begin{proof}
2581
We build operadic orderings inductively, showing that the choices for $\Un(n)$, $n\leq 4$ determine higher ones.
2582
We prove the statement for the $\LA$ order, the $\SU$ and opposite orders are similar.
2583
First, we decide to order the unique pair of $\Un(2)$ as $(\{1\},\{2\})$.
2584
The operadic property then determines the orders of all the pairs of $\Un(3)$ and $\Un(4)$, except the pair $\{\{1,4\},\{2,3\}\}$, for which we choose the order $(\{1,4\},\{2,3\})$.
2585
Now, we claim that all the higher choices are forced by the operadic property, and lead to the $\LA$ diagonal.
2586
Starting instead with the orders $(\{1\},\{2\})$ and $(\{2,3\},\{1,4\})$ would give the $\SU$ diagonal, and reversing the pairs would give the opposite orders.
2587
2588
Let $\ell \geq 2$ and suppose that for all $k\leq \ell$, we have given the pair \[\{I_k,J_k\} \eqdef \{\{1,k+2,{k+3},\dots,{2k-1},2k\},\{2,3,\dots,k+1\}\}\] the $\LA$ ordering $(I_k,J_k)$.
2589
Then from \cref{lem:operadic-ordering}, we know that the only $\{I,J\}$ pair of order $\ell+1$ that will not decompose, and hence be specified by the already chosen conditions, is~$\{I_{\ell+1},J_{\ell+1}\}$.
2590
As such, the only way we can vary from $\LA$ is to order this element in the opposite direction $(J_{\ell+1},I_{\ell+1})$.
2591
We now consider a particular decomposable pair $\{I'_m,J'_m\}$ where ${I'_m \eqdef I_m \sqcup \{3\} \ssm \{m+2\}}$ and~$J'_m \eqdef J_m \sqcup \{m+2\} \ssm \{3\}$, of order $m \eqdef \ell+2$, that will lead to a contradiction (see \cref{ex:Non-coherent order contradiction}).
2592
On the one side, we can decompose $\{I'_m,J'_m\} = \{I_a \cup I_b, J_a \cup J_b\}$ with $I_a \eqdef \{1,m+3,\dots,2m\}, J_a \eqdef \{4,5,\dots,m+2\}, I_b \eqdef \{3\}$ and $J_b \eqdef \{2\}$.
2593
By hypothesis, we have the orders $(J_a,I_a)$ and $(J_b,I_b)$ which imply the order $(J'_m,I'_m)$ since our ordering is operadic.
2594
On the other side, we can decompose $\{I'_m,J'_m\} = \{I_c \cup I_d, J_c \cup J_d\}$, where $I_c \eqdef \{1,m+3,\dots, 2m-1\},$ $J_c \eqdef \{2,5,\dots,m+1\}$, $I_d \eqdef \{3, 2m\}$ and $J_d \eqdef \{4, m+2\}$.
2595
By hypothesis, we have the orders $(I_c,J_c)$ and $(I_d,J_d)$, which imply that $(I'_m,J'_m)$ since our ordering is operadic.
2596
We arrived at a contradiction.
2597
Thus, the only possible operadic choice of ordering for $\{I_{\ell+1},J_{\ell+1}\}$ is the $\LA$ ordering, which finishes the proof.
2598
\end{proof}
2599
2600
\begin{example}
2601
\label{ex:Non-coherent order contradiction}
2602
To illustrate our proof of \cref{prop:operadicOrdering}, consider an operadic ordering $\Or$ for which the $\LA$ ordering holds for pairs of order $1$ and $2$, but is reversed for pairs of order $3$, \ie
2603
\begin{align*}
2604
(\{1\}, \{2\}) \in \Or, \quad (\{1,4\}, \{2,3\}) \in \Or, \text{ and } (\{2, 3, 4\}, \{1, 5, 6 \}) \in \Or.
2605
\end{align*}
2606
Then, the pair $\{I'_4,J'_4\}=\{\{1, 3, 7, 8\}, \{2, 4, 5, 6\}\}$ admits two different orientations.
2607
In particular,
2608
\begin{align*}
2609
(\{ 4, 5, 6 \} , \{1, 7, 8\}) \in \Or \text{ and } (\{2\}, \{3\}) \in \Or & \Longrightarrow (\{2, 4, 5, 6\}, \{1, 3, 7, 8\} ) \in \Or \ \text{and} \\
2610
(\{1, 7\}, \{2, 5\}) \in \Or \text{ and } (\{3, 8\}, \{4, 6\}) \in \Or & \Longrightarrow (\{1, 3, 7, 8\}, \{2, 4, 5, 6\}) \in \Or.
2611
\end{align*}
2612
\end{example}
2613
2614
Combining \cref{prop:equiv-operadic} with \cref{prop:operadicOrdering}, we get the main result of this section.
2615
Recall that a vector induces the weak order on the vertices of the standard permutahedron if and only if it has strictly decreasing coordinates (see \cref{def:permutahedron}).
2616
2617
\begin{theorem}
2618
\label{thm:unique-operadic}
2619
There are exactly four operadic geometric diagonals of the permutahedra, given by the $\LA$ and $\SU$ diagonals, and their opposites~$\LA^{\op}$ and $\SU^{\op}$.
2620
The $\LA$ and $\SU$ diagonals are the only operadic geometric diagonals which induce the weak order on the vertices of the permutahedron.
2621
\end{theorem}
2622
2623
\begin{remark}
2624
This answers by the negative a conjecture regarding unicity of diagonals on the permutahedra, raised at the beginning of~\cite[Sect.~3]{SaneblidzeUmble}, and could be seen as an alternative statement.
2625
See \cref{sec:shifts} where we prove that the geometric $\SU$ diagonal is a topological enhancement of the $\SU$ diagonal from~\cite{SaneblidzeUmble}.
2626
\end{remark}
2627
2628
%%%%%%%%%%%%%%%
2629
2630
\subsection{Isomorphisms between operadic diagonals}
2631
\label{subsec:isos-LA-SU}
2632
2633
Let $P$ be a centrally symmetric polytope, and let $s : P \to P$ be its involution, given by taking a point $\b{p} \in P$ to its reflection $s(\b{p})$ with respect to the center of symmetry of $P$.
2634
Note that this map is cellular, and induces an involution on the face lattice of $P$.
2635
For the permutahedron~$\Perm$, this face poset involution is given in terms of ordered partitions of~$[n]$ by the map~$A_1 | \cdots | A_k \mapsto A_k | \cdots | A_1$.
2636
2637
The permutahedron~$\Perm$ possesses another interesting symmetry, namely, the cellular involution ${r : \Perm \to \Perm}$ which sends a point $p \in \Perm$ to its reflection with respect to the hyperplane of equation \[ \sum_{i=1}^{\lfloor n/2 \rfloor}x_i = \sum_{i=1}^{\lfloor n/2 \rfloor}x_{n-i+1} . \]
2638
This involution also respects the face poset structure.
2639
In terms of ordered partitions, it replaces each block $A_j$ in $A_1 | \cdots | A_k$ by the block $r(A_j) \eqdef \set{r(i)}{i \in A_j}$ where~$r(i) \eqdef n-i+1$.
2640
2641
The involution $t : P \times P \to P \times P$, sending $(x,y)$ to~$(y,x)$, takes a cellular diagonal to another cellular diagonal.
2642
As we have already remarked in \cref{def:oppositeDiagonal}, this involution sends a geometric diagonal~$\triangle$ to its opposite~$\triangle^{\op}$.
2643
%As we have already remarked in \cref{subsec:LASUdiagonal}, in the case of the permutahedron it sends $\LAD$ to $(\LAD)^\op$ and $\SUD$ to $(\SUD)^\op$.
2644
2645
\pagebreak
2646
\begin{theorem}
2647
\label{thm:bijection-operadic-diagonals}
2648
For the permutahedron~$\Perm$, the involutions $t$ and $rs \times rs$ are cellular isomorphisms between the four operadic diagonals, through the following commutative diagram
2649
\begin{center}
2650
\begin{tikzcd}
2651
\LAD \arrow[r,"t"] \arrow[d,"rs \times rs"']&
2652
(\LAD)^{\op}\arrow[d,"rs \times rs"]\\
2653
\SUD \arrow[r,"t"'] &
2654
(\SUD)^\op
2655
\end{tikzcd}
2656
\end{center}
2657
Moreover, they induce poset isomorphisms at the level of the face lattices.
2658
\end{theorem}
2659
2660
\begin{proof}
2661
The result for the transpositions $t$ and the commutativity of the diagram are straightforward, so we prove that $rs\times rs$ is a cellular isomorphism respecting the poset structure.
2662
First, since $r(\min(I\cup J))=\max(r(I) \cup r(J))$, we observe that the map $(I,J) \mapsto (r(J),r(I))$ defines a bijection between $\LA(n)$ and $\SU(n)$.
2663
Then, if $\sigma$ is an ordered partition such that $I$ dominates $J$ in $\sigma$ (\cref{def:domination}), it must be the case that $J$ dominates $I$ in $s\sigma$, and consequently $rJ$ dominates $rI$ in $rs\sigma$.
2664
As such, the domination characterization of the diagonals (\cref{thm:minimal}), tells us~$(\sigma,\tau) \in \LAD$ if and only if $(rs(\sigma),rs(\tau)) \in \SUD$.
2665
Finally, since both $t,r$ and $s$ preserve the poset structures, so does their composition, which finishes the proof.
2666
\end{proof}
2667
2668
\begin{remark} \label{rem:Alternate Isomorphism}
2669
There is a second, distinct isomorphism given by $t(r\times r)$.
2670
This follows from the fact that $s\times s:\LAD \to (\LAD)^{\op}$ is an isomorphism (and also for $s\times s:\SUD \to (\SUD)^{\op}$ ), as such the composite
2671
\begin{center}
2672
\begin{tikzcd}
2673
t(r\times r):\LAD \arrow[r,"s\times s"] &
2674
(\LAD)^{\op}\arrow[r,"rs \times rs"] &
2675
(\SUD)^\op \arrow[r,"t"] & \SUD
2676
\end{tikzcd}
2677
\end{center}
2678
is also an isomorphism.
2679
This second isomorphism has the conceptual benefit of sending left shift operators to left shift operators (and right to right), see \cref{prop:trr is an isomorphism of shifts}.
2680
\end{remark}
2681
2682
%%%%%%%%%%%%%%%
2683
2684
\subsection{Facets of operadic diagonals}
2685
\label{subsec:facets-operadic-diags}
2686
2687
We now aim at describing the facets of the $\LA$ and $\SU$ diagonals.
2688
We have seen in \cref{subsec:multiBraidArrangement} that facets of any diagonal of the permutahedron~$\Perm$ are in bijection with $(2,n)$-partition trees, that is, pairs of unordered partitions of $[n]$ whose intersection graph is a tree.
2689
These pairs of partitions were first studied and called \defn{essential complementary partitions} in~\cite{Chen, ChenGoyal, KajitaniUenoChen}.
2690
Specializing \cref{prop:PFtoOPF1}, we now explain how to order these pairs of partitions to get the facets of $\LAD$ and $\SUD$.
2691
2692
\begin{theorem}
2693
\label{thm:facet-ordering}
2694
Let $(\sigma,\tau)$ be a pair of ordered partitions of $[n]$ whose underlying intersection graph is a $(2,n)$-partition tree.
2695
The pair~$(\sigma,\tau)$ is a facet of the $\LA$ (\resp $\SU$) diagonal if and only if the minimum (\resp maximum) of every directed path between two consecutive blocks of $\sigma$ or $\tau$ is oriented from $\sigma$ to $\tau$ (\resp from $\tau$ to $\sigma$).
2696
\end{theorem}
2697
2698
\begin{proof}
2699
We specialize \cref{prop:PFtoOPF1} to the $\LA$ diagonal, the proof for the $\SU$ diagonal is similar.
2700
Let $\b{v}$ be a vector inducing the $\LA$ diagonal as in \cref{def:LA-and-SU}.
2701
Without loss of generality, we place the first copy of the braid arrangement centered at $0$, and the second copy centered at $\b{v}$.
2702
From \cref{def:multiBraidArrangementPrecise,rem:translationMatrix}, we get that $A_{1,s,t}=0$ and $A_{2,s,t}=v_s-v_t$ for any edges $s,t$.
2703
We treat the case of two consecutive blocks $A$ and $B$ of $\sigma$, the analysis for $\tau$ is similar.
2704
The directed path between $A$ and $B$ is a sequence of edges $r_0,r_1,\dots,r_q$.
2705
Let us denote by $I \eqdef \{r_0,r_2,\dots\}$ de set of edges directed from $\sigma$ to $\tau$, and by $J \eqdef \{r_1,r_3,\dots\}$ its complement.
2706
According to \cref{prop:PFtoOPF1}, the order between $A$ and $B$ is determined by the sign of $A_{1,r_0,r_q}- \sum_{p \in [q]} A_{i_p,r_{p-1},r_p}$, where $i_p$ is the copy of the block adjacent to both edges $r_{p-1}$ and $r_p$.
2707
Thus, the order between $A$ and $B$ is determined by the sign of $\sum_{i \in I} v_i - \sum_{j \in J} v_j$, which according to the definition of $\LAD$ is positive whenever $\min(I\cup J) \in I$.
2708
Thus, the pair $(\sigma,\tau) \in \LAD$ if and only if the minimum of every directed path between two consecutive blocks of $\sigma$ or $\tau$ is oriented from $\sigma$ to $\tau$.
2709
\end{proof}
2710
2711
In the rest of the paper, we shall represent ordered $(2,n)$-partition trees $(\sigma,\tau)$ of facets by drawing $\sigma$ on the left, and $\tau$ on the right, with their blocks from top to bottom.
2712
The conditions ``oriented from $\sigma$ to $\tau$" in the preceding Theorem then reads as ``oriented from left to right", an expression we shall adopt from now on.
2713
2714
\begin{example}
2715
\label{ex:ECbijection}
2716
Below are the two orderings of the $(2,7)$-partition tree $\{15,234,6,7\} \times \{13,2,46,57\}$ giving facets of the $\LA$ (left) and $\SU$ (right) diagonals, obtained via \cref{thm:facet-ordering}.
2717
Note that ordered partitions should be read from top to bottom.
2718
\begin{center}
2719
\begin{tikzpicture}[scale=.7]
2720
\node[anchor=east] (1) at (-1.5, -1) {$15$};
2721
\node[anchor=east] (2) at (-1.5, -2) {$7$};
2722
\node[anchor=east] (3) at (-1.5, -3) {$234$};
2723
\node[anchor=east] (4) at (-1.5, -4) {$6$};
2724
%
2725
\node[anchor=west] (5) at (1.5, -1) {$57$};
2726
\node[anchor=west] (6) at (1.5, -2) {$46$};
2727
\node[anchor=west] (7) at (1.5, -3) {$13$};
2728
\node[anchor=west] (8) at (1.5, -4) {$2$};
2729
%
2730
\draw[thick] (1.east) -- (5.west);
2731
\draw[thick] (1.east) -- (7.west);
2732
\draw[thick] (2.east) -- (5.west);
2733
\draw[thick] (3.east) -- (6.west);
2734
\draw[thick] (3.east) -- (7.west);
2735
\draw[thick] (3.east) -- (8.west);
2736
\draw[thick] (4.east) -- (6.west);
2737
\end{tikzpicture}
2738
$\quad$
2739
\begin{tikzpicture}[scale=.7]
2740
\node[anchor=east] (7) at (-1.5, -4) {$7$};
2741
\node[anchor=east] (6) at (-1.5, -3) {$6$};
2742
\node[anchor=east] (234) at (-1.5, -2) {$234$};
2743
\node[anchor=east] (15) at (-1.5, -1) {$15$};
2744
%
2745
\node[anchor=west] (2) at (1.5, -4) {$2$};
2746
\node[anchor=west] (13) at (1.5, -3) {$13$};
2747
\node[anchor=west] (46) at (1.5, -2) {$46$};
2748
\node[anchor=west] (57) at (1.5, -1) {$57$};
2749
%
2750
\draw[thick] (2.west) -- (234.east);
2751
\draw[thick] (13.west) -- (15.east);
2752
\draw[thick] (13.west) -- (234.east);
2753
\draw[thick] (46.west) -- (234.east);
2754
\draw[thick] (46.west) -- (6.east);
2755
\draw[thick] (57.west) -- (15.east);
2756
\draw[thick] (57.west) -- (7.east);
2757
\end{tikzpicture}
2758
\end{center}
2759
In the $\LA$ facet (left), we have $7 < 234$, since in the path between the two vertices $7 \xrightarrow{7} 57 \xrightarrow{5} 15 \xrightarrow{1} 13 \xrightarrow{3} 234$, the minimum $1$ is oriented from left to right.
2760
In this case, we have $(I,J)=(\{1,7\},\{3,5\})$.
2761
Similarly, the path $57 \xrightarrow{5} 15 \xrightarrow{1} 13 \xrightarrow{3} 234 \xrightarrow{4} 46$ imposes that $57 < 46$, for $(I,J)=(\{1,4\},\{3,5\})$.
2762
\end{example}
2763
2764
\begin{remark}
2765
Note that forgetting the order in a facet of $\LAD$, and then ordering again the $(2,n)$-partition tree to obtain a facet of $\SUD$, provides a bijection between facets that was not considered in \cref{subsec:isos-LA-SU}.
2766
However, this map is not defined on the other faces.
2767
\end{remark}
2768
2769
\begin{example}
2770
The isomorphism $rs\times rs$ from \cref{thm:bijection-operadic-diagonals} sends the $\LA$ facet from \cref{ex:ECbijection}\,(left) to the following $\SU$ facet\,(right).
2771
\begin{center}
2772
\begin{tikzpicture}[scale=.7]
2773
\node[anchor=east] (1) at (-1.5, -1) {$15$};
2774
\node[anchor=east] (2) at (-1.5, -2) {$7$};
2775
\node[anchor=east] (3) at (-1.5, -3) {$234$};
2776
\node[anchor=east] (4) at (-1.5, -4) {$6$};
2777
%
2778
\node[anchor=west] (5) at (1.5, -1) {$57$};
2779
\node[anchor=west] (6) at (1.5, -2) {$46$};
2780
\node[anchor=west] (7) at (1.5, -3) {$13$};
2781
\node[anchor=west] (8) at (1.5, -4) {$2$};
2782
%
2783
\draw[thick] (1.east) -- (5.west);
2784
\draw[thick] (1.east) -- (7.west);
2785
\draw[thick] (2.east) -- (5.west);
2786
\draw[thick] (3.east) -- (6.west);
2787
\draw[thick] (3.east) -- (7.west);
2788
\draw[thick] (3.east) -- (8.west);
2789
\draw[thick] (4.east) -- (6.west);
2790
%
2791
\end{tikzpicture}
2792
\raisebox{3.4em}{$\xrightarrow{rs\times rs}$}
2793
\begin{tikzpicture}[scale=.7]
2794
\node[anchor=east] (5) at (-1.5, -4) {$37$};
2795
\node[anchor=east] (6) at (-1.5, -3) {$1$};
2796
\node[anchor=east] (7) at (-1.5, -2) {$456$};
2797
\node[anchor=east] (8) at (-1.5, -1) {$2$};
2798
%
2799
\node[anchor=west] (1) at (1.5, -4) {$13$};
2800
\node[anchor=west] (2) at (1.5, -3) {$24$};
2801
\node[anchor=west] (3) at (1.5, -2) {$57$};
2802
\node[anchor=west] (4) at (1.5, -1) {$6$};
2803
%
2804
\draw[thick] (1.west) -- (5.east);
2805
\draw[thick] (1.west) -- (6.east);
2806
\draw[thick] (2.west) -- (7.east);
2807
\draw[thick] (2.west) -- (8.east);
2808
\draw[thick] (3.west) -- (5.east);
2809
\draw[thick] (3.west) -- (7.east);
2810
\draw[thick] (4.west) -- (7.east);
2811
\end{tikzpicture}
2812
\end{center}
2813
Moreover, it sends the path $57 \xrightarrow{5} 15 \xrightarrow{1} 13 \xrightarrow{3} 234 \xrightarrow{4} 46 $ to the path $24 \xrightarrow{4} 456 \xrightarrow{5} 57 \xrightarrow{7} 37 \xrightarrow{3} 13$, where the maximum $7$ is oriented from right to left, witnessing the fact that $24 < 13$.
2814
The associated $(I,J)=(\{1,4 \},\{3,5\})$ becomes $(r(J),r(I))=(\{3,5 \},\{4,7\})$.
2815
\end{example}
2816
2817
2818
%%%%%%%%%%%%%%%
2819
2820
\subsection{Vertices of operadic diagonals}
2821
\label{subsec:vertices-operadic-diags}
2822
2823
We are now interested in characterizing the vertices that occur in an operadic diagonal as pattern-avoiding pairs of partitions of $[n]$.
2824
These pairs form a strict subset of the weak order intervals.
2825
We first need the following Lemma, which follows directly from the definition of domination (\cref{def:domination}).
2826
2827
\begin{lemma}
2828
\label{lem:Coherent Domination}
2829
Let $\sigma$ be an ordered partition of $[n]$ and let $I,J \subseteq [n]$ be such that $I$ dominates $J$ in~$\sigma$.
2830
For an element $i$ in $I$ or $J$, we denote by $\sigma^{-1}(i)$ the index of the block containing it in $\sigma$.
2831
Then, for any $i \in I,j \in J$, we have $I\ssm i$ dominates $J\ssm j$ in $\sigma$ if and only if
2832
\begin{enumerate}
2833
\item either~$\sigma^{-1}(j) < \sigma^{-1}(i)$ (meaning that $j$ comes strictly before $i$ in $\sigma$),
2834
\item or for all $k$ between $\sigma^{-1}(i)$ and $\sigma^{-1}(j)$, the first $k$ blocks of $\sigma$ contain strictly more elements of $I$ than of $J$.
2835
\end{enumerate}
2836
\end{lemma}
2837
2838
% As before, we represent a permutation $\sigma: [n] \to [n]$ by the non-commutative word~$\sigma(1)\cdots \sigma(n)$.
2839
2840
\begin{definition}
2841
For $k\leq n$, a permutation $\tau$ of $[k]$ is a \defn{pattern} of a permutation $\sigma$ of $[n]$ if there exists a subset $I \eqdef \{i_1 < \dots < i_k\} \subset [n]$ so that the permutation~$\tau$ gives the relative order of the entries of~$\sigma$ at positions in~$I$, \ie $\tau_u < \tau_v \eqdef \sigma_{i_u} < \sigma_{i_v}$.
2842
We say that $\sigma$ \defn{avoids} $\tau$ if $\tau$ is not a pattern~of~$\sigma$.
2843
\end{definition}
2844
2845
\begin{example}
2846
The pairs of permutations $(\sigma,\tau)$ avoiding the patterns $(21,12)$ are precisely the intervals of the weak order.
2847
\end{example}
2848
2849
\begin{theorem}
2850
\label{thm:patterns}
2851
A pair of permutations of $[n]$ is a vertex of the $\LA$ (\resp $\SU$) diagonal if and only if for any~$k\geq 1$ and for any $(I,J) \in \LA(k)$ (\resp $\SU(k)$) of size $\card{I}=\card{J}=k$ it avoids~the~patterns
2852
\begin{align}
2853
(j_1 i_1 j_2 i_2 \cdots j_k i_k,\ i_2 j_1 i_3 j_2 \cdots i_k j_{k-1} i_1 j_k), \tag{LA} \label{eq:pattern} \\
2854
\text{ \resp } (j_1 i_1 j_2 i_2 \cdots j_k i_k, \ i_1 j_k i_2 j_1 \cdots i_{k-1} j_{k-2}i_k j_{k-1}), \tag{SU}
2855
\end{align}
2856
where $I=\{i_1,\dots,i_k\}$, $J=\{j_1,\dots,j_k\}$ and $i_1=1$ (\resp $j_k=2k$).
2857
\end{theorem}
2858
2859
\begin{example}\label{exm:intervalsWeakOrderNotDiagonals}
2860
\enlargethispage{.5cm}
2861
For each $k \ge 1$, there are $\binom{2k-1}{k-1,k}(k-1)!k!$ avoided standardized patterns.
2862
For~$k=1$, both diagonals avoid $(21,12)$, so the vertices are intervals of the weak order.
2863
For~$k=2$,
2864
\begin{itemize}
2865
\item $\LA$ avoids
2866
$(3142,2314), (4132,2413),
2867
(2143,3214), (4123,3412),
2868
(2134,4213), (3124,4312)$.
2869
\item $\SU$ avoids
2870
$(1243,2431),(1342,3421),
2871
(2143,1432),(2341,3412),
2872
(3142,1423),(3241,2413)$.
2873
\end{itemize}
2874
As such, the vertices of $\LA$ and $\SU$ are a strict subset of all intervals of the weak order.
2875
Here is an illustration of the patterns avoided for $k=4$.
2876
The $\LA$ pattern is drawn on the left, the $\SU$ pattern is drawn on the right, and they are in bijection under the isomorphism $t(r\times r)$ (\cref{subsec:isos-LA-SU}), where the bijection between elements is $(i_1,i_2,i_3,i_4)\mapsto (j'_4,j'_1,j'_2,j'_3)$ and $(j_1,j_2,j_3,j_4)\mapsto (i'_1,i'_2,i'_3,i'_4)$.
2877
\begin{center}
2878
\begin{tikzpicture}[scale=.65]
2879
\node[anchor=east] (0) at (-1.5, 0) {$j_1$};
2880
\node[anchor=east] (1) at (-1.5, -1) {$i_1$};
2881
\node[anchor=east] (2) at (-1.5, -2) {$j_2$};
2882
\node[anchor=east] (3) at (-1.5, -3) {$i_2$};
2883
\node[anchor=east] (4) at (-1.5, -4) {$j_3$};
2884
\node[anchor=east] (5) at (-1.5, -5) {$i_3$};
2885
\node[anchor=east] (6) at (-1.5, -6) {$j_4$};
2886
\node[anchor=east] (7) at (-1.5, -7) {$i_4$};
2887
%
2888
\node[anchor=west] (8) at (1.5, 0) {$i_2$};
2889
\node[anchor=west] (9) at (1.5, -1) {$j_1$};
2890
\node[anchor=west] (10) at (1.5, -2) {$i_3$};
2891
\node[anchor=west] (11) at (1.5, -3) {$j_2$};
2892
\node[anchor=west] (12) at (1.5, -4) {$i_4$};
2893
\node[anchor=west] (13) at (1.5, -5) {$j_3$};
2894
\node[anchor=west] (14) at (1.5, -6) {$i_1$};
2895
\node[anchor=west] (15) at (1.5, -7) {$j_4$};
2896
%
2897
\draw[thick, green] (0.east) -- (9.west);
2898
\draw[thick, blue, dashed] (1.east) -- (14.west);
2899
\draw[thick, green] (2.east) -- (11.west);
2900
\draw[thick, blue] (3.east) -- (8.west);
2901
\draw[thick, green] (4.east) -- (13.west);
2902
\draw[thick, blue] (5.east) -- (10.west);
2903
\draw[thick, green] (6.east) -- (15.west);
2904
\draw[thick, blue] (7.east) -- (12.west);
2905
%
2906
\end{tikzpicture}
2907
\raisebox{7em}{$\xrightarrow{t(r\times r)}$}
2908
\begin{tikzpicture}[scale=.65]
2909
\node[anchor=east] (0) at (-1.5, 0) {$j'_1$};
2910
\node[anchor=east] (1) at (-1.5, -1) {$i'_1$};
2911
\node[anchor=east] (2) at (-1.5, -2) {$j'_2$};
2912
\node[anchor=east] (3) at (-1.5, -3) {$i'_2$};
2913
\node[anchor=east] (4) at (-1.5, -4) {$j'_3$};
2914
\node[anchor=east] (5) at (-1.5, -5) {$i'_3$};
2915
\node[anchor=east] (6) at (-1.5, -6) {$j'_4$};
2916
\node[anchor=east] (7) at (-1.5, -7) {$i'_4$};
2917
%
2918
\node[anchor=west] (8) at (1.5, 0) {$i'_1$};
2919
\node[anchor=west] (9) at (1.5, -1) {$j'_4$};
2920
\node[anchor=west] (10) at (1.5, -2) {$i'_2$};
2921
\node[anchor=west] (11) at (1.5, -3) {$j'_1$};
2922
\node[anchor=west] (12) at (1.5, -4) {$i'_3$};
2923
\node[anchor=west] (13) at (1.5, -5) {$j'_2$};
2924
\node[anchor=west] (14) at (1.5, -6) {$i'_4$};
2925
\node[anchor=west] (15) at (1.5, -7) {$j'_3$};
2926
%
2927
\draw[thick, green] (0.east) -- (11.west);
2928
\draw[thick, blue] (1.east) -- (8.west);
2929
\draw[thick, green] (2.east) -- (13.west);
2930
\draw[thick, blue] (3.east) -- (10.west);
2931
\draw[thick, green] (4.east) -- (15.west);
2932
\draw[thick, blue] (5.east) -- (12.west);
2933
\draw[thick, green, dashed] (6.east) -- (9.west);
2934
\draw[thick, blue] (7.east) -- (14.west);
2935
%
2936
\end{tikzpicture}
2937
\end{center}
2938
The alternate isomorphism $t(s\times s)$, provides an alternate way to establish a bijection between these two patterns.
2939
The avoided patterns for higher $k$ extend this crisscrossing shape.
2940
2941
\end{example}
2942
2943
\begin{proof}[Proof of \cref{thm:patterns}]
2944
We give the proof for the $\LA$ diagonal, the one for the $\SU$ diagonal is similar, and can be obtained by applying either of the two isomorphisms of \cref{subsec:isos-LA-SU}.
2945
According to \cref{thm:IJ-description}, we have $(\sigma,\tau) \notin \LAD$ if and only if there is an $(I,J)$ such that~$J$ dominates~$I$ in $\sigma$ and~$I$ dominates $J$ in $\tau$.
2946
It is clear that if a pair of permutations $(\sigma,\tau)$ contains a pattern of the form~(\ref{eq:pattern}), then the associated $(I,J)$ satisfies the domination condition.
2947
Thus, we just need to show the reverse implication.
2948
We proceed by induction on the cardinality $\card{I} = \card{J}$.
2949
The case of cardinality $1$ is clear.
2950
Now suppose that for all $(I,J)$ of size $\card{I} = \card{J} \le m-1$, we have if~$J$ dominates $I$ in $\sigma$ and $I$ dominates $J$ in $\tau$, then $(\sigma,\tau)$ contains a pattern of the form~(\ref{eq:pattern}).
2951
2952
We need to show that this is still true for the pairs $(I,J)$ of size $\card{I} = \card{J} = m$.
2953
Firstly, we need only consider standardized $I,J$ conditions, and pairs of permutations of $[2m]$.
2954
Indeed, if we define $(\sigma',\tau') \eqdef \std(\sigma \cap (I\cup J),\tau \cap (I\cup J))$, and $(I',J')\eqdef\std(I',J')$,
2955
then if $(\sigma,\tau)$ satisfies the $(I,J)$ domination condition, this implies $(\sigma',\tau')$ satisfies $(I',J')$.
2956
Conversely, if $(\sigma',\tau')$ has a pattern, then this implies $(\sigma,\tau)$ has the same pattern, which yields the indicated simplification.
2957
2958
Let $(\sigma,\tau)$ be such a pair of permutations.
2959
Suppose the leftmost element $i_1$ of $I$ in $\sigma$ is not $1$, and let us write $j_1$ for the leftmost element of $J$ in $\tau$.
2960
Consider the pair ${(I',J') \eqdef (I\ssm \{i_1\}, J \ssm \{j_1\})}$.
2961
Using both cases of \cref{lem:Coherent Domination}, we have $J'$ dominates $I'$ in $\sigma$, and $I'$ dominates $J'$ in $\tau$, and we can thus conclude by induction that $(\sigma,\tau)$ contains a smaller pattern.
2962
2963
So, we assume the leftmost element of $I$ in $\sigma$ is $i_1=1$, and for $n\geq1$ we prove that,
2964
\begin{enumerate}[(a)]
2965
\item If $(\sigma,\tau)=(j_1 i_1 j_2 i_2 \dots j_{n-1} i_{n-1} w i_{n} w', \ i_2 j_1 i_3 j_2 \dots i_{n}j_{n-1}w'') $, and $j_n$ is the leftmost element of $J\ssm \{j_1, \dots, j_{n-1}\}$ in $\tau$, then either $w = j_n$, or it matches a smaller pattern.
2966
\item If $(\sigma,\tau)=(j_1 i_1 j_2 i_2 \dots j_{n-1} i_{n-1} w'', \ i_2 j_1 i_3 j_2 \dots i_{n-1}j_{n-2} w j_{n-1} w'')$, and $i_n$ is the leftmost element of ${I\ssm \{i_1, \dots, i_{n-1}\}}$ in $\sigma$, then either $w = i_n$, or it matches a smaller pattern.
2967
\end{enumerate}
2968
We prove (a), the proof of (b) proceeds similarly.
2969
Let $j_n$ be the leftmost element of $J\ssm \{j_1, \dots, j_{n-1} \}$ in $\tau$.
2970
If $w\neq j_n$, then either (i) $w$ consists of multiple elements of $J$ including $j_n$, or (ii) $j_n$ comes after $i_n$ in $\sigma$.
2971
Now consider the pair $(I',J') \eqdef (I\ssm \{i_n\}, J \ssm \{j_n\})$.
2972
As was the case for the proof that $i_1=1$, we have $I'$ dominates $J'$ in $\tau$.
2973
To prove that $J'$ dominates $I'$ in $\sigma$, we split by the cases.
2974
In case (i), we may apply condition $(2)$ of \cref{lem:Coherent Domination}.
2975
In case (ii), either $i_n$ comes before~$j_n$, in which case we meet condition $(1)$ of \cref{lem:Coherent Domination}, or $j_n$ comes before $i_n$.
2976
In this last situation, we have condition $(2)$ of \cref{lem:Coherent Domination} holds as $i_n$ is the leftmost element of~${I\ssm \{i_1, \dots, i_{n-1}\}}$ in $\sigma$.
2977
Thus, if $w \neq j_n$, by the inductive hypothesis, we match a smaller pattern.
2978
2979
Finally, using statements (a) and (b) above, we can inductively generate $(\sigma,\tau)$, determining $j_1$ via (a), then $i_2$ via (b), then $j_2$ via (a), and so on.
2980
This inductive process fully generates $\sigma$, and places all elements of $\tau$ except $i_1$, yielding $\tau=i_2 j_1 i_3 j_2 \cdots i_k j_{k-1} w j_k w''$.
2981
However, as $j_k$ must be dominated by an element of $I$, this forces $w = i_1$ and $w'' =\varnothing$, completing the proof.
2982
\end{proof}
2983
2984
\subsection{Relation to the facial weak order}
2985
\label{sec:facial-weak-order}
2986
2987
There is a natural lattice structure on all ordered partitions of~$[n]$ which extends the weak order on permutations of~$[n]$.
2988
This lattice was introduced in~\cite{KrobLatapyNovelliPhanSchwer}, where it is called \emph{pseudo-permutahedron} and defined on packed words rather than ordered partitions.
2989
It was later generalized to arbitrary finite Coxeter groups in~\cite{PalaciosRonco, DermenjianHohlwegPilaud}, where it is called \emph{facial weak order} and expressed in more geometric terms.
2990
We now recall a definition of the facial weak order on ordered partitions, and use the vertex characterization of the preceding section to show all faces of the $\LA$ and $\SU$ diagonal are intervals of this order.
2991
2992
\begin{definition}[\cite{KrobLatapyNovelliPhanSchwer,PalaciosRonco,DermenjianHohlwegPilaud}]
2993
The \defn{facial weak order} on ordered partitions is the transitive closure of the relations
2994
\begin{align}
2995
\sigma_1|\dots|\sigma_k < \sigma_1|\cdots|\sigma_i \sqcup \sigma_{i+1}|\cdots|\sigma_k \quad & \text{for any } \sigma_1|\dots|\sigma_k \text{ with } \max \sigma_i < \min \sigma_{i+1}, \label{eq:facial weak 1}\\
2996
\sigma_1|\cdots|\sigma_i \sqcup \sigma_{i+1}|\cdots|\sigma_k < \sigma_1|\dots|\sigma_k \quad & \text{for any } \sigma_1|\dots|\sigma_k \text{ with } \min \sigma_i > \max \sigma_{i+1}. \label{eq:facial weak 2}
2997
\end{align}
2998
\end{definition}
2999
3000
The facial weak order recovers the weak order on permutations as illustrated in \cref{fig:Hasse diagram Perm3}.
3001
\begin{figure}
3002
\begin{tikzpicture}[xscale=1.3, yscale=1.5]
3003
\node (1) at (-1, -.5) {$3|12$};
3004
\node (2) at (-2, -1) {$3|1|2$};
3005
\node (3) at (-2, -2) {$13|2$};
3006
\node (4) at (-2, -3) {$1|3|2$};
3007
\node (5) at (-1, -3.5) {$1|23$};
3008
%
3009
\node (6) at (1, -.5) {$23|1$};
3010
\node (7) at (2, -1) {$2|3|1$};
3011
\node (8) at (2, -2) {$2|13$};
3012
\node (9) at (2, -3) {$2|1|3$};
3013
\node (10) at (1, -3.5) {$12|3$};
3014
%
3015
\node (11) at (0, 0) {$3|2|1$};
3016
\node (12) at (0, -2) {$123$};
3017
\node (13) at (0, -4) {$1|2|3$};
3018
%
3019
\draw[thick] (1) -- (2);
3020
\draw[thick] (2) -- (3);
3021
\draw[thick] (3) -- (4);
3022
\draw[thick] (4) -- (5);
3023
%
3024
\draw[thick] (6) -- (7);
3025
\draw[thick] (7) -- (8);
3026
\draw[thick] (8) -- (9);
3027
\draw[thick] (9) -- (10);
3028
%
3029
\draw[thick] (1) -- (11);
3030
\draw[thick] (6) -- (11);
3031
%
3032
\draw[thick] (1) -- (12);
3033
\draw[thick] (5) -- (12);
3034
\draw[thick] (6) -- (12);
3035
\draw[thick] (10) -- (12);
3036
%
3037
\draw[thick] (5) -- (13);
3038
\draw[thick] (10) -- (13);
3039
\end{tikzpicture}
3040
\caption{The Hasse diagram of the facial weak order for $\Perm[3]$.}
3041
\label{fig:Hasse diagram Perm3}
3042
\end{figure}
3043
3044
\begin{proposition}
3045
If $(\sigma,\tau)\in \LAD$, or $(\sigma,\tau)\in \SUD$, then $\sigma \leq \tau$ under the facial weak order.
3046
\end{proposition}
3047
3048
\begin{proof}
3049
By \cref{prop:magicalFormula}, the faces $(\sigma,\tau)$ satisfy $\max_{\mathbf{v}} \sigma \leq \min_{\mathbf{v}} \tau$ under the weak order.
3050
Thus, if we can show that $\sigma\leq \max_{\mathbf{v}} \sigma$ and $\min_{\mathbf{v}} \tau \leq \tau$ under the facial weak order, then the result immediately follows.
3051
If $\sigma$ is a face of the permutahedra, then under both the $\LA$, and $\SU$ orientation vectors, the vertex $\max_{\mathbf{v}} \sigma$ is given by writing out each block of $\sigma$ in decreasing order, and the vertex $\min_{\mathbf{v}} \sigma$ is given by writing out each block of $\sigma$ in increasing order.
3052
Then under the facial weak order
3053
$\sigma \leq \max_{\mathbf{v}} \sigma$, as repeated applications of \cref{eq:facial weak 2} shows that a block of elements is smaller than those same elements arranged in decreasing order.
3054
Similarly $\min_{\mathbf{v}} \sigma \leq \sigma$, as repeated applications of \cref{eq:facial weak 1} shows that a sequence of increasing elements is smaller than those same elements in a block.
3055
\end{proof}
3056
3057
\begin{example}
3058
The facet $13|24|57|6 \times 3|17|456|2 \in \SUD$, satisfies the inequality through the vertices
3059
\begin{align*}
3060
13|24|57|6 < 3|1|4|2|7|5|6 < 3|1|7|4|5|6|2 < 3|17|456|2 \, .
3061
\end{align*}
3062
\end{example}
3063
3064
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3065
3066
\newpage
3067
\section{Shift lattices}
3068
\label{sec:shifts}
3069
3070
In this section, we prove that the geometric $\SU$ diagonal $\SUD$ agrees at the cellular level with the original Saneblidze-Umble diagonal defined in~\cite{SaneblidzeUmble}.
3071
This involves some shift operations on facets of the diagonal, which are interesting on their own right, and lead to lattice and cubic structures.
3072
The proof is technical and proceeds in several steps: we introduce two additional combinatorial descriptions of the diagonal, that we call the $1$-shift and $m$-shift $\SU$ diagonals, and show the sequence of equivalences
3073
\begin{center}
3074
\begin{tikzcd}
3075
\text{original $\SUD$} \arrow[r,leftrightarrow,"\ref{prop:iso-original-shift-diagonals}"]&
3076
\text{$1$-shift $\SUD$} \arrow[r,leftrightarrow,"\ref{prop:iso-1-to-m-shift}"]&
3077
\text{$m$-shift $\SUD$} \arrow[r,leftrightarrow,"\ref{prop:iso-shift-IJ-diagonals}"]&
3078
\text{geometric $\SUD$} .
3079
\end{tikzcd}
3080
\end{center}
3081
Throughout this section, we borrow notation from~\cite{SaneblidzeUmble-comparingDiagonals}.
3082
3083
%%%%%%%%%%%%%%%
3084
3085
\subsection{Topological enhancement of the original $\SU$ diagonal}
3086
\label{subsec:topological-SU}
3087
3088
We proceed to introduce different versions of the $\SU$ diagonal, and to prove that all these notions coincide.
3089
3090
%%%%%%%%%%%%%%%
3091
3092
\subsubsection{Strong complementary pairs}
3093
3094
We start by the following definition.
3095
3096
\begin{definition}
3097
\label{def:strong-complementary-pairs}
3098
A \defn{strong complementary pair}, or~\defn{$\SCP$} for short, is a pair $(\sigma,\tau)$ of ordered partitions of $[n]$ where $\sigma$ is obtained from a permutation of $[n]$ by merging the adjacent elements which are decreasing, and $\tau$ is obtained from the same permutation by merging the adjacent elements which are increasing.
3099
\end{definition}
3100
3101
We denote by $\SCP(n)$ the set of $\SCP$s of $[n]$, which is in bijection with the set of permutations of $[n]$.
3102
As is clear from the definition, the intersection graph of a $\SCP$ is a $(2,n)$-partition tree.
3103
3104
\begin{example}
3105
\label{ex:strong-complementary}
3106
The $\SCP$ associated to the permutation $3|1|7|4|2|5|6$ is
3107
\begin{center}
3108
\begin{tikzpicture}[scale=.7]
3109
\node (p) at (0, 0) {$13|247|5|6 \times 3|17|4|256$};
3110
\node[anchor=east] (1) at (-1.5, -1) {$13$};
3111
\node[anchor=east] (2) at (-1.5, -2) {$247$};
3112
\node[anchor=east] (3) at (-1.5, -3) {$5$};
3113
\node[anchor=east] (4) at (-1.5, -4) {$6$};
3114
%
3115
\node[anchor=west] (5) at (1.5, -1) {$3$};
3116
\node[anchor=west] (6) at (1.5, -2) {$17$};
3117
\node[anchor=west] (7) at (1.5, -3) {$4$};
3118
\node[anchor=west] (8) at (1.5, -4) {$256$};
3119
%
3120
\draw[thick] (1.east) -- (5.west);
3121
\draw[thick] (1.east) -- (6.west);
3122
\draw[thick] (2.east) -- (6.west);
3123
\draw[thick] (2.east) -- (7.west);
3124
\draw[thick] (2.east) -- (8.west);
3125
\draw[thick] (3.east) -- (8.west);
3126
\draw[thick] (4.east) -- (8.west);
3127
%
3128
\end{tikzpicture}
3129
\end{center}
3130
Observe that the permutation can be read off the intersection graph of the SCP by a vertical down slice through the edges.
3131
\end{example}
3132
3133
\begin{notation}
3134
For a $(2,n)$-partition tree $(\sigma,\tau)$, we denote by $\sigma_{i,j}$ (\resp $\tau_{i,j}$) the unique oriented path between blocks $\sigma_{i}$ and $\sigma_j$ (\resp $\tau_{i}$ and $\tau_j$).
3135
Note that we make a slight abuse in notation, as the path $\sigma_{i,j}$ also depends on $\tau$.
3136
\end{notation}
3137
3138
We can immediately characterize the paths between adjacent blocks in $\SCP$s.
3139
3140
\begin{proposition}
3141
\label{lem:SCP-path-desc}
3142
For any $\SCP$ $(\sigma,\tau)$, we have
3143
\begin{enumerate}
3144
\item $ \sigma_{i,i+1} = ( \min \sigma_i, \max \sigma_{i+1} )$ and $\min \sigma_i< \max \sigma_{i+1}$, and
3145
\item $ \tau_{i,i+1} = ( \max \tau_i, \min \tau_{i+1} )$ and $\min \tau_{i+1}< \max \tau_{i}$.
3146
\end{enumerate}
3147
As a consequence, all $\SCP$s are in both $\LAD$ and $\SUD$, and constitute precisely the set of facets~$(\sigma,\tau)$ of these diagonals such that $\max_{\b{v}}(\sigma) = \min_{\b{v}}(\tau)$.
3148
\end{proposition}
3149
\begin{proof}
3150
First, the path description of $(\sigma,\tau)$ is a straightforward observation.
3151
Second, since the minima of these paths are traversed from left to right, and the maxima from right to left, \cref{thm:facet-ordering} implies that $\SCP$s are in both geometric operadic diagonals~$\LAD$ and~$\SUD$.
3152
Third, the fact that these constitute the facets $(\sigma,\tau) \in \LAD$ or $\SUD$ satisfying $\max_{\b{v}}(\sigma) = \min_{\b{v}}(\tau)$ can be seen as follows.
3153
The maximal (\resp minimal) vertex of a face $\sigma$ of the permutahedron with respect to the weak order, is obtained by ordering the elements of each block of $\sigma$ in decreasing (\resp increasing) order.
3154
Thus, it is clear that the original permutation giving rise to a $\SCP$ $(\sigma,\tau)$ is the permutation $\max_{\b{v}}(\sigma)=\min_{\b{v}}(\tau)$, for any vector $\b{v}$ inducing the weak order.
3155
Since both diagonals agree with this order on the vertices, we have that $\SCP$s are indeed facets of $\LAD$ and $\SUD$ with the desired property.
3156
The fact that these are \emph{all} the facets with this property follows from the bijection between $\SCP(n)$ and the permutations of~$[n]$.
3157
\end{proof}
3158
3159
3160
%%%%%%%%%%%%%%%
3161
3162
\subsubsection{Shifts and the $\SU$ diagonals}
3163
\label{subsec:SU-shifts}
3164
3165
We recall the definition of the original $\SU$ diagonal of~\cite{SaneblidzeUmble}, based on the exposition given in~\cite{SaneblidzeUmble-comparingDiagonals}.
3166
We then introduce two variants of this definition, the $1$-shift and $m$-shift $\SU$ diagonals, which will be shown to be equivalent to the original one.
3167
3168
\begin{definition}
3169
\label{def:subset shifts}
3170
Let $\sigma=\sigma_1| \cdots |\sigma_k$ be an ordered partition, and let~$M \subsetneq \sigma_{i}$ be a non-empty subset of the block $\sigma_i$.
3171
For $m\geq 1$, the \defn{right $m$-shift} $R_M$, moves the subset $M$, $m$ blocks to the right, \ie
3172
\begin{align*}
3173
R_M(\sigma) \eqdef \sigma_1 | \cdots | \sigma_i \ssm M | \cdots | \sigma_{i+m} \cup M | \cdots | \sigma_k
3174
\end{align*}
3175
while the \defn{left $m$-shift} $L_M$, moves the subset $M$, $m$ blocks to the left, \ie
3176
\begin{align*}
3177
L_M(\sigma) \eqdef \sigma_1 | \cdots | \sigma_{i-m} \cup M | \cdots | \sigma_{i} \ssm M | \cdots | \sigma_k .
3178
\end{align*}
3179
\end{definition}
3180
3181
\begin{definition}
3182
\label{def:movable-subsets}
3183
Let $\sigma$ denote either one of the two ordered partitions of $[n]$ in an ordered \mbox{$(2,n)$-partition} tree, and let $M \subsetneq \sigma_i$.
3184
The right $m$-shift $R_M$ (\resp the left $m$-shift $L_M$) is
3185
\begin{enumerate}
3186
\item \defn{block-admissible} if $\min \sigma_i \notin M$ and $\min M > \max \sigma_{i+m}$ (\resp $\min M > \max \sigma_{i-m}$),
3187
\item \defn{path-admissible} if $\min M > \max \sigma_{i,i+m}$ (\resp $\min M > \max \sigma_{i,i-m}$).
3188
\end{enumerate}
3189
\end{definition}
3190
3191
\begin{remark}
3192
\label{rem:inverses}
3193
Observe that for a given subset $M$, an inverse to the right $m$-shift $R_M$ (\resp left $m$-shift $L_M$) is given by the left $m$-shift $L_M$ (\resp right $m$-shift $R_M$).
3194
Moreover, one $m$-shift is path-admissible if and only if its inverse is.
3195
\end{remark}
3196
3197
\begin{example}\label{ex:1-shift example}
3198
Performing the $1$-shifts $R_7$ and $L_{56}$ (they are both block and path admissible) of the $\SCP$~$(\sigma,\tau)$ of \cref{ex:strong-complementary}, one obtains the pair~$R_{7}(\sigma) \times L_{56}(\tau)$, as illustrated below.
3199
\begin{center}
3200
\begin{tikzpicture}[scale=.7]
3201
\node (p) at (0, 0) {$13|247|5|6 \times 3|17|4|256$};
3202
\node[anchor=east] (1) at (-1.5, -1) {$13$};
3203
\node[anchor=east] (2) at (-1.5, -2) {$247$};
3204
\node[anchor=east] (3) at (-1.5, -3) {$5$};
3205
\node[anchor=east] (4) at (-1.5, -4) {$6$};
3206
%
3207
\node[anchor=west] (5) at (1.5, -1) {$3$};
3208
\node[anchor=west] (6) at (1.5, -2) {$17$};
3209
\node[anchor=west] (7) at (1.5, -3) {$4$};
3210
\node[anchor=west] (8) at (1.5, -4) {$256$};
3211
%
3212
\draw[thick] (1.east) -- (5.west);
3213
\draw[thick] (1.east) -- (6.west);
3214
\draw[thick] (2.east) -- (6.west);
3215
\draw[thick] (2.east) -- (7.west);
3216
\draw[thick] (2.east) -- (8.west);
3217
\draw[thick] (3.east) -- (8.west);
3218
\draw[thick] (4.east) -- (8.west);
3219
\end{tikzpicture}
3220
\quad
3221
\raisebox{3.4em}{$\xrightarrow{R_{7} \times L_{56}}$}
3222
\quad
3223
\begin{tikzpicture}[scale=.7]
3224
\node (p) at (0, 0) {$13|24|57|6 \times 3|17|456|2$};
3225
\node[anchor=east] (1) at (-1.5, -1) {$13$};
3226
\node[anchor=east] (2) at (-1.5, -2) {$24$};
3227
\node[anchor=east] (3) at (-1.5, -3) {$57$};
3228
\node[anchor=east] (4) at (-1.5, -4) {$6$};
3229
%
3230
\node[anchor=west] (5) at (1.5, -1) {$3$};
3231
\node[anchor=west] (6) at (1.5, -2) {$17$};
3232
\node[anchor=west] (7) at (1.5, -3) {$456$};
3233
\node[anchor=west] (8) at (1.5, -4) {$2$};
3234
%
3235
\draw[thick] (1.east) -- (5.west);
3236
\draw[thick] (1.east) -- (6.west);
3237
\draw[thick] (3.east) -- (6.west);
3238
\draw[thick] (2.east) -- (7.west);
3239
\draw[thick] (2.east) -- (8.west);
3240
\draw[thick] (3.east) -- (7.west);
3241
\draw[thick] (4.east) -- (7.west);
3242
\end{tikzpicture}
3243
\end{center}
3244
\end{example}
3245
3246
We shall concentrate on three families of shifts: block-admissible $1$-shifts of subsets of various sizes, path-admissible $1$-shifts of singletons, and path-admissible $m$-shifts of singletons, for various $m\geq 1$,
3247
and show that specific sequences generate the same diagonal.
3248
3249
\begin{definition}
3250
\label{def:SU-admissible}
3251
Let $\sigma$ denote either one of the two ordered partitions of $[n]$ in an ordered \mbox{$(2,n)$-parti}\-tion tree, and let~$\b{M} = (M_1,\dots,M_p)$ with $M_1 \subsetneq \sigma_{i_1}, \dots, M_p \subsetneq \sigma_{i_p}$ for some~$p \ge 1$.
3252
Then the sequence of right shifts~$R_\mathbf{M}(\sigma) \eqdef R_{M_p} \dots R_{M_1}(\sigma)$ is
3253
\begin{enumerate}
3254
\item \defn{block-admissible} if we have $1\leq i_1 < i_2 < \dots < i_p \leq k-1$, and each $R_{M_j}$ is a block-admissible $1$-shift,
3255
\item \defn{path-admissible} if each $R_{M_j}$ is a path-admissible $m_j$-shift, for some $m_j \geq 1$.
3256
\end{enumerate}
3257
Admissible sequences of left shifts are defined similarly and are denoted by $L_\mathbf{M}(\tau)$.
3258
\end{definition}
3259
3260
By convention, we declare the empty sequence of shift operators to be admissible, and to act by the identity, \ie we have $R_{\mathbf{\varnothing}}(\sigma) \eqdef \sigma$ and $L_{\mathbf{\varnothing}} (\tau) \eqdef \tau$.
3261
3262
\begin{definition}
3263
\label{def:classical-SU}
3264
The facets of the \defn{original $\SU$ diagonal}, the \defn{$1$-shift $\SU$ diagonal} and the \defn{$m$-shift $\SU$ diagonal} are defined by the formula
3265
\begin{align*}
3266
\SUD([n]) = \bigcup_{(\sigma,\tau)} \bigcup_{\mathbf{M}, \mathbf{N}} R_\mathbf{M}(\sigma)\times L_\mathbf{N}(\tau)
3267
\end{align*}
3268
where the unions are taken over all $\SCP$s $(\sigma, \tau)$ of $[n]$, and respectively over all block-admissible sequences of subset $1$-shifts $\b{M},\b{N}$, over all path-admissible sequences of singleton $1$-shifts, and over all path-admissible sequences of singleton $m_j$-shifts, for various $m_j \ge 1$.
3269
\end{definition}
3270
3271
\begin{remark}
3272
Observe that the left (\resp right) shifts acts on the right (\resp left) ordered partition.
3273
In the analogous description for the $\LA$ diagonal $\LAD$ obtained at in \cref{subsec:shifts-under-iso}, left and right shifts act on the left and right ordered partitions, respectively.
3274
\end{remark}
3275
3276
%%%%%%%%%%%%%%%
3277
3278
\subsubsection{First isomorphism between $\SU$ diagonals}
3279
3280
We start by analyzing the original $\SU$ diagonal.
3281
3282
\begin{proposition}
3283
\label{prop:SU-preserves-max}
3284
Block-admissible sequences of subset $1$-shifts, defining the original $\SU$ diagonal, conserve
3285
\begin{enumerate}
3286
\item the maximal element of any path between two blocks of the same ordered partition,
3287
\item the direction in which this element is traversed.
3288
\end{enumerate}
3289
In particular, for a pair of ordered partitions $(R_{\mathbf{M}}(\sigma),L_{\mathbf{N}}(\tau))$ obtained via a block-admissible sequence of $1$-shifts from a $\SCP$~$(\sigma,\tau)$, we have
3290
\begin{align}
3291
\label{eq:max-1}
3292
\max R_{\mathbf{M}}(\sigma)_{i,j} = \max \sigma_{i,j} \qquad \text{and} \qquad \max L_{\mathbf{N}}(\tau)_{i,j} = \max \tau_{i,j} \tag{P}
3293
\end{align}
3294
and consequently
3295
\begin{align}
3296
\label{eq:max-2}
3297
\max R_{\mathbf{M}}(\sigma)_{i,i+1} = \max \sigma_{i+1} \qquad \text{and}\qquad \max L_{\mathbf{N}}(\tau)_{i,i+1} = \max \tau_{i} . \tag{B}
3298
\end{align}
3299
\end{proposition}
3300
3301
Note that in \cref{eq:max-1}, the maxima $\max \sigma_{i,j}$ and $\max \tau_{i,j}$ are maxima of \emph{paths}, while in \cref{eq:max-2} the maxima $\max \sigma_{i+1}$ and $\max \tau_{i}$ are maxima of \emph{blocks}.
3302
3303
\begin{proof}
3304
We consider the right shift operator; the left shift operator proceeds similarly.
3305
Let us start with Point (1).
3306
As $\SCP$s trivially meet the above conditions, we will prove the result inductively by assuming that the result holds for $(R_{\mathbf{M}}(\sigma),L_{\mathbf{N}}(\tau))$, and then showing that applying a block-admissible operator $R_{M_k}$, for $M_k\subsetneq \sigma_k$,
3307
conserves the maximal elements of paths.
3308
By the inductive hypothesis, we know that $\max R_{\mathbf{M}}(\sigma)_{k,k+1}= \max \sigma_{k+1}$.
3309
As $R_{M_k}$ is an admissible operator, we know two things: firstly that $\min M_k > \max R_{\mathbf{M}}(\sigma)_{k+1}$, and secondly that $\max \sigma_{k+1} = \max R_{\mathbf{M}}(\sigma)_{k+1}$, as $k$ is greater than the maximal index used by $\mathbf{M}$.
3310
So combining these, we know that
3311
\begin{align}
3312
\label{eq:shift-subset-dominates}
3313
\min M_k > \max R_{\mathbf{M}}(\sigma)_{k+1} = \max \sigma_{k+1} = \max R_{\mathbf{M}}(\sigma)_{k,k+1} . \tag{W}
3314
\end{align}
3315
A key consequence of this inequality is that the intersection graph of $(R_{M_k}R_{\mathbf{M}}(\sigma),L_{\mathbf{N}}(\tau))$ is a bipartite tree conditional on $(R_{\mathbf{M}}(\sigma),L_{\mathbf{N}}(\tau))$ being a bipartite tree: the shift will not disconnect the graph as none of the shifted elements are in the path $R_{\mathbf{M}}(\sigma)_{k,k+1}$.
3316
So, it is legitimate to speak of unique paths between blocks.
3317
3318
We now explicitly explore how the shift operator $R_{M_k}$ alters paths.
3319
Throughout the rest of this proof, we use the following shorthand: we denote by $\delta_{k,k+1} \eqdef R_{\mathbf{M}}(\sigma)_{k,k+1}$ the path between the $k$\ordinal{} and $(k+1)$\ordinalst{} blocks in $(R_{\mathbf{M}}(\sigma),L_{\mathbf{N}}(\tau))$, and by $\delta_{k+1,k}$ the same path reversed.
3320
Let~$\gamma$ be any path between two blocks of $(R_{\mathbf{M}}(\sigma),L_{\mathbf{N}}(\tau))$, and let~$\gamma'$ be the path between the same blocks, by indices, in $(R_{M_k}R_{\mathbf{M}}(\sigma),L_{\mathbf{N}}(\tau))$.
3321
There are four cases to consider.
3322
\begin{enumerate}[i)]
3323
\item the path $\gamma$ does not contain an element of $M_k$.
3324
In this case, it is unaffected by the shift, so $\gamma' = \gamma$.
3325
We note that in light of \cref{eq:shift-subset-dominates}, the path $\delta_{k,k+1}$ meets this case.
3326
3327
\item the path $\gamma$ contains one element $m\in M_k$, \ie it is of the form $\gamma = \alpha m \beta$, with $\alpha$ or $\beta$ possibly empty.
3328
We assume that $m$ is incoming to $R_\mathbf{M}(\sigma)_{k}$, the case when it is outgoing is similar.
3329
We must have $\alpha \cap \delta_{k,k+1}=\varnothing$, since otherwise there would be an oriented loop from $R_\mathbf{M}(\sigma)_{k}$ to itself.
3330
For the same reason, $\beta \cap \delta_{k,k+1}$ must be connected and starting at $R_\mathbf{M}(\sigma)_{k}$.
3331
Then there are two cases to consider:
3332
\begin{enumerate}
3333
\item the path $\beta$ does not use any steps of $\delta_{k,k+1}$, in which case $\gamma' = \alpha m \delta_{k+1,k} \beta$.
3334
This is a path in the tree of $(R_{M_k}R_{\mathbf{M}}(\sigma),L_{\mathbf{N}}(\tau))$ with no repeated steps, as such it must be the unique minimal path.
3335
3336
\item the path $\beta$ uses steps of $\delta_{k,k+1}$, in which case $\gamma' = \alpha m (\delta_{k+1,k}\ssm \beta)(\beta \ssm \delta_{k+1,k})$.
3337
This follows, as we know that $\beta$ must follow the path $\delta_{k,k+1}$ for some time before diverging ($\beta$ could also be a subset of $\delta_{k,k+1}$, in which case it will never diverge).
3338
As such, the path $(\delta_{k+1,k}\ssm \beta)$ reaches the point of divergence from $R_{\mathbf{M}}(\sigma)_{k+1}$ instead of $R_{\mathbf{M}}(\sigma)_{k}$, and the path $(\beta \ssm \delta_{k+1,k})$ completes the rest of the route unchanged.
3339
\end{enumerate}
3340
3341
\item the path $\gamma$ contains two elements of $M_k$.
3342
In this case we still have $\gamma'=\gamma$ (in path elements) but $\gamma'$ will step through the $(k+1)$\ordinalst{} block instead of $\gamma$ stepping through the $k$\ordinal{} block.
3343
3344
\item the path $\gamma$ contains more than two elements of $M_k$.
3345
This is impossible, as $\gamma$ would not be a minimal path on a tree.
3346
3347
\end{enumerate}
3348
Observe that all (non-trivial or non-contradictory) paths $\gamma'$ contain $m \geq \min M_k$ and either some addition or deletion by $\delta_{k,k+1}$.
3349
It thus follows from \cref{eq:shift-subset-dominates} that $\max \gamma' = \max \gamma$, since in each case, the maximal element will either be $m$, or in $\alpha$, or in $\beta$.
3350
This finishes the proof of Point~(1).
3351
3352
For Point (2), we need to see that the maximal element $\max \gamma = \max \gamma'$ is traversed in the same direction.
3353
It is immediate for cases (i), (ii.a) and (iii); the condition is empty in the case~(iv).
3354
For the case (ii.b) it follows from the observation that the number of steps of $\beta \cap \delta_{k,k+1}$ and~$\delta_{k,k+1} \ssm \beta$ have the same parity: since~$\delta_{k,k+1}$ has an even number of steps, either they both have an even number of steps, or an odd number, which completes the proof.
3355
\end{proof}
3356
3357
\begin{corollary}
3358
\label{cor:SU-1-shift-preserves-max}
3359
Path-admissible sequences of singleton $1$-shifts, defining the $1$-shift $\SU$ diagonal, conserve
3360
\begin{enumerate}
3361
\item the maximal element of any path between two blocks of the same ordered partition,
3362
\item the direction in which this element is traversed.
3363
\end{enumerate}
3364
\end{corollary}
3365
3366
\begin{proof}
3367
By the definition of path-admissible $1$-shifts (\cref{def:SU-admissible}), the outer inequality of \cref{eq:shift-subset-dominates} holds by assumption.
3368
As such, can run the proof of \cref{prop:SU-preserves-max} \emph{mutatis mutandis}.
3369
\end{proof}
3370
3371
We are now in position to show that the $1$-shift and $m$-shift descriptions are equivalent.
3372
3373
%Combining \cref{cor:SU-1-shift-preserves-max} and \cref{prop:iso-1-to-m-shift}
3374
3375
\begin{proposition}
3376
\label{prop:iso-1-to-m-shift}
3377
The $1$-shift and $m$-shift $\SU$ diagonals coincide.
3378
\end{proposition}
3379
3380
\begin{proof}
3381
It is clear that any path-admissible sequence of $1$-shifts is a path-admissible sequence of $m$-shifts, and thus that the facets of the $1$-shift $\SU$ diagonal are facets of the $m$-shift $\SU$ diagonal.
3382
For the reverse inclusion, we need to show that any $m$-shift can be re-expressed as a path-admissible sequence of $1$-shifts.
3383
We proceed by induction, and consider only the case of right shifts, the case of left shifts is similar.
3384
For right $1$-shifts, there is nothing to prove.
3385
Let $(\sigma,\tau)$ be a pair of ordered partitions which has been generated only by $k$-shifts, for~$k<m$.
3386
We wish to show that any path-admissible right $m$-shift $R_\rho^m$ on $\sigma$ can be decomposed as a path-admissible $1$-shift $R_\rho^{1}$ followed by a path-admissible $(m-1)$-shift $R_\rho^{m-1}$, which yields the result by induction.
3387
As $R_\rho^{m}$ is path-admissible, we know that $\rho > \max \sigma_{i,i+m}$, and we want to show that
3388
$\rho>\max \sigma_{i,i+m}\geq \max \sigma_{i,i+1}, \max R_\rho^{1}(\sigma)_{i+1,i+m}$.
3389
\begin{center}
3390
\begin{tikzpicture}[scale=0.7]
3391
\node[anchor=east] (1) at (-0.5,0) {$\sigma_i$};
3392
\coordinate (2) at (1,0);
3393
\node[anchor=south] (3) at (1,1.5) {$\sigma_{i+1}$};
3394
\node[anchor=west] (4) at (2.5,0) {$\sigma_{i+m}$};
3395
\draw[thick,->] (1.east) -- node[below] {$\alpha$} (2);
3396
\draw[thick,->] (2) -- node[left] {$\beta$} (3.south);
3397
\draw[thick,->] (2) -- node[below] {$\gamma$} (4.west);
3398
\end{tikzpicture}
3399
\end{center}
3400
We define the oriented paths $\alpha \eqdef \sigma_{i,i+1}\ssm \sigma_{i+1,i+m}$, $\beta \eqdef \sigma_{i,i+1}\ssm \sigma_{i,i+m}$ and $\gamma \eqdef \sigma_{i,i+m}\ssm \sigma_{i,i+1}$, as illustrated above.
3401
Suppose that $\beta$ is not empty, and moreover that $\max\sigma_{i+1,i+m}>\max\sigma_{i,i+m}$ or $\max \sigma_{i,i+1}> \max \sigma_{i,i+m}$.
3402
Then we must have that $\max \beta> \max \alpha, \max \gamma$.
3403
However, $\max\beta$ cannot be the maximum of both $\sigma_{i,i+1}$ and $\sigma_{i+1,i+m}$.
3404
Indeed, in this case, it would be traversed in two opposite directions, which is impossible since by the induction hypothesis $(\sigma,\tau)$ can be generated by $1$-shifts, and by \cref{cor:SU-1-shift-preserves-max} these conserve the maximal elements of paths and their direction.
3405
We thus have $\max \sigma_{i,i+m}\geq \max\sigma_{i,i+1}, \max \sigma_{i+1,i+m}$, and applying \cref{cor:SU-1-shift-preserves-max} again yields $\max \sigma_{i,i+m}\geq \max R_\rho^{1}(\sigma)_{i+1,i+m}$ as required.
3406
\end{proof}
3407
3408
We stress that \cref{eq:shift-subset-dominates} holds for $m$-shifts without us needing to perform shifts in increasing order, or requiring $\min M_k > \max R_{\mathbf{M}}(\sigma)_{k+1}$.
3409
We are now ready to prove that the $m$-shift description is equivalent to the original one.
3410
3411
\begin{theorem}
3412
\label{prop:iso-original-shift-diagonals}
3413
The original and $m$-shift $\SU$ diagonals coincide.
3414
\end{theorem}
3415
3416
\begin{proof}
3417
Since $m$-shift and $1$-shift diagonals are equivalent (\cref{prop:iso-1-to-m-shift}), it suffices to show that the $1$-shift and original $\SU$ diagonals coincide.
3418
We analyze the right shift operator, the case of the left shift is similar.
3419
First, we observe that any block-admissible right shift $R_{M}(\sigma)$, for~$M\subsetneq\sigma_k$, can be decomposed into a series of singleton right $1$-shifts; since $\min M > \max \sigma_{k,k+1}$ by the proof of \cref{prop:SU-preserves-max}, we can shift the elements of $M$ to the right, one after the other (in any order!).
3420
This shows that any facet of the original $\SU$ diagonal is also a facet in the $1$-shift $\SU$ diagonal.
3421
3422
For the reverse inclusion, we proceed by induction.
3423
We are required to show that if we apply a right $1$-shift to $(R_{\mathbf{M}}(\sigma),L_{\mathbf{N}}(\tau))$, say $(R_{\rho}R_{\mathbf{M}}(\sigma),L_{\mathbf{N}}(\tau))$, then this can be re-expressed as a well-defined subset shift operation $(R_{\mathbf{M'}}(\sigma),L_{\mathbf{N}}(\tau))$.
3424
Suppose that prior to the $1$-shift, the element $\rho$ lives in block $\ell$, then we must have
3425
\begin{align*}
3426
1 \leq i_1 < \cdots < i_j \leq \ell < i_{j+1} <\cdots< i_p \leq k-1
3427
\end{align*}
3428
for some $j$.
3429
If $i_j < \ell$, then we have $R_{\rho}R_{\mathbf{M}}(\sigma) = R_{M_{p}}\cdots R_{\{\rho\}}R_{M_{j}}\cdots R_{M_{1}}(\sigma)$, and we are done.
3430
Otherwise, if $i_j = \ell$, we set $M'_{j} \eqdef M_{j} \cup \{\rho \}$.
3431
It is clear that ${R_{\rho}R_{\mathbf{M}}(\sigma) = R_{M_{p}}\cdots R_{M'_{j}}\cdots R_{M_{1}}(\sigma)}$, however, we need to check that $R_{M'_{j}}$ is block-admissible, \ie that $\min M_{j}' > \max (R_{M_{j-1}}\cdots R_{M_{1}}(\sigma))_{i_j+1}$.
3432
If we have~$\rho > \min M_{j}$, then we are done since in this case $\min M_{j}'=\min M_{j}$ and $R_{M_{j}}$ is block-admissible.
3433
Otherwise, we have $\rho=\min M_{j}'$.
3434
Since by definition block-admissible shift operators do not move the minimal element of a block, we have $\rho > \min R_{\mathbf{M}}(\sigma)_{i_j}= \min (R_{M_{j-1}}\cdots R_{M_{1}}(\sigma))_{i_j}$.
3435
Then, by induction, \cref{prop:SU-preserves-max} shows that $\rho>\max \sigma_{i_j+1} = \max (R_{M_{j-1}}\cdots R_{M_{1}}(\sigma))_{i_j + 1}$, where the equality follows as $i_1<\cdots<i_j$.
3436
This proves that $R_{M'_{j}}$ is block-admissible, completing the inductive proof.
3437
\end{proof}
3438
3439
%%%%%%%%%%%%%%%
3440
3441
\subsubsection{Inversions}
3442
\label{subsec:inversions}
3443
3444
Our next goal is to prove the equivalence between the $m$-shift and geometric $\SU$ diagonals (\cref{prop:iso-shift-IJ-diagonals}).
3445
As a tool for this proof, we now study \emph{inversions}, or crossings in the partition trees of the geometric $\SU$ diagonal.
3446
3447
\begin{definition}\label{def:inverions}
3448
Let $\sigma$ be an ordered partition.
3449
\begin{itemize}
3450
\item The \defn{inversions} of an ordered partition are $I(\sigma):= \{(i,j): i<j \land \sigma^{-1}(j)<\sigma^{-1}(i) \}$.
3451
\item The \defn{anti-inversions} of an ordered partition are $J(\sigma):= \{(i,j): i<j \land \sigma^{-1}(i)< \sigma^{-1}(j) \}$.
3452
\end{itemize}
3453
We then define the \defn{inversions of an ordered partition pair} $I((\sigma,\tau)):=I(\tau)\cap J(\sigma)$.
3454
\end{definition}
3455
In words, the inversions of an ordered partition pair are those $i<j$ pairs in which $j$ comes in an earlier block than $i$ in $\tau$, and $i$ comes in an earlier block than $j$ in $\sigma$.
3456
3457
\begin{proposition}
3458
\label{p:crossings}
3459
The set of inversions of a facet of the geometric $\SU$ diagonal is in bijection with its set of edge crossings.
3460
Moreover, under this bijection, strong complementary pairs correspond to facets with no crossings.
3461
\end{proposition}
3462
3463
\begin{proof}
3464
For the first part of the statement,
3465
we note that crossings are clearly produced by both $I(\tau)\cap J(\sigma)$ and $I(\sigma)\cap J(\tau)$ (i.e. in the second case $j$ appears before $i$ in $\sigma$ and $i$ before $j$ in $\tau$).
3466
However, the later cannot occur in a facet of the geometric $\SUD$; this follows immediately from the $(I,J)$-conditions for $\card{I} = \card{J} = 1$.
3467
The second part of the statement follows from the fact that facets of the diagonal $\SUD$ with no crossings are in bijection with permutations.
3468
By definition (\cref{def:strong-complementary-pairs}), from a partition one obtains a $\SCP$, which is in the geometric $\SUD$ (\cref{lem:SCP-path-desc}).
3469
In the other way around, given a $\SCP$, one can read-off the partition in the associated tree, which has no crossings, by a vertical down-slice of edges (\cref{ex:strong-complementary}).
3470
\end{proof}
3471
3472
See \cref{fig: Inversion and lattice counter example} for an example of this bijection.
3473
3474
\begin{definition}
3475
We say that an edge crossing is an \defn{adjacent crossing} if the two crossing elements are in adjacent blocks of the partition tree (\ie they are in blocks of the form $\sigma_i|\sigma_{i+1}$ or $\tau_i|\tau_{i+1}$).
3476
\end{definition}
3477
3478
\begin{lemma}
3479
\label{lem:adjacent-crossing}
3480
A facet $(\sigma,\tau)$ of the geometric $\SU$ diagonal has a crossing if and only if it has an adjacent crossing.
3481
\end{lemma}
3482
3483
\begin{proof}
3484
An adjacent crossing is clearly a crossing.
3485
In the other direction, suppose there is a crossing between an element of $\sigma_i$ and an element of $\sigma_j$.
3486
If $\sigma_i$ and $\sigma_j$ are not adjacent, then the ``triangle" produced by the crossing elements encloses another $\sigma_k$ such that $i<k<j$, and this produces other crossings. We may repeat this process until an adjacent crossing is found.
3487
\end{proof}
3488
3489
3490
3491
%%%%%%%%%%%%%%%
3492
3493
\subsubsection{Second isomorphism between $\SU$ diagonals}
3494
\label{sec:Iso m-shifts to IJ}
3495
3496
We now aim at showing that the $m$-shift and the geometric $\SU$ diagonal coincide (\cref{thm:unique-operadic}).
3497
Recall from \cref{rem:inverses} that left and right path-admissible $m$-shifts are inverses to one another.
3498
3499
\begin{proposition}
3500
\label{lem:IJ-closed-under-shifts}
3501
Let $(\sigma,\tau)$ be a facet of the geometric $\SU$ diagonal $\SUD$.
3502
Then, any pair of ordered partitions obtained by applying a path-admissible $m$-shift to $(\sigma,\tau)$ is also in the geometric $\SU$ diagonal.
3503
\end{proposition}
3504
3505
\begin{proof}
3506
We consider a right path-admissible shift $R_\rho$, the left shift and dual result proceeds similarly.
3507
Combining \cref{cor:SU-1-shift-preserves-max} and \cref{prop:iso-1-to-m-shift}, we have that the maxima of paths between consecutive vertices in $(R_{\rho}(\sigma),\tau)$ are the same as the ones in $(\sigma,\tau)$, and are moreover traversed in the same direction.
3508
Thus, all maxima of paths in $(R_{\rho}(\sigma),\tau)$ are traversed from right to left, and hence by \cref{thm:facet-ordering}, we have that $(R_{\rho}(\sigma),\tau)$ is in the geometric $\SU$ diagonal.
3509
\end{proof}
3510
3511
\begin{lemma}
3512
\label{lem:inverse-to-SCP}
3513
Any facet $(\sigma,\tau)$ of the geometric $\SU$ diagonal $\SUD$ is mapped to a $\SCP$ by a path-admissible sequence of inverse $m$-shifts.
3514
\end{lemma}
3515
3516
\begin{proof}
3517
We show that any facet $(\sigma,\tau)$ which has a crossing, and is hence not a $\SCP$, admits an inverse shift operation.
3518
This shows that a finite number of inverse shift operations converts any facet to a $\SCP$ (if $\sigma$ is an ordered partition of $n$ with $k$ blocks, then clearly less than $n^k$ inverse shifts are possible).
3519
We consider the following partition of the set of facets with crossings, illustrated by example in \cref{ex:proof-inverse-shift}.
3520
\begin{enumerate}
3521
\item All adjacent blocks are connected by paths of length $2$.
3522
\item There exist adjacent blocks which are connected by a path of length $2k$ for $k>1$.
3523
\begin{enumerate}
3524
\item The maximal step of this path is not the last step.
3525
\item The maximal step of this path is the last step.
3526
\begin{enumerate}
3527
\item The $\tau$ block containing the maximal step is not the greatest block.
3528
\item The $\tau$ block containing the maximal step is the greatest block.
3529
\end{enumerate}
3530
\end{enumerate}
3531
\end{enumerate}
3532
In Case $(1)$, as $(\sigma,\tau)$ has a crossing, it has an adjacent crossing by \cref{lem:adjacent-crossing}.
3533
This adjacent crossing is of the following form (we illustrate the case where the crossing happens on the left, the case when it happens on the right is similar):
3534
\begin{center}
3535
\begin{tikzpicture}[scale=.6]
3536
\node[anchor=east] (1) at (-1.5, 0) {$\sigma_i$};
3537
\node[anchor=east] (2) at (-1.5, -2) {$\sigma_{i+1}$};
3538
%
3539
\node[anchor=west] (4) at (1.5, 0) {$\tau_j$};
3540
\node[anchor=west] (5) at (1.5, -2) {$\tau_{k}$};
3541
%
3542
\draw[thick] (1.east) -- (5.west);
3543
\draw[thick] (2.east) -- (5.west);
3544
\draw[thick] (2.east) -- (4.west);
3545
\node (6) at (1.1, -0.7) {$\rho$};
3546
\node (7) at (0, -1.6) {$b$};
3547
\node (8) at (-1.1, -0.7) {$c$};
3548
\end{tikzpicture}
3549
\quad
3550
\raisebox{1.9em}{$\xrightarrow{R^{-1}_{\rho}}$}
3551
\quad
3552
\begin{tikzpicture}[scale=.6]
3553
\node[anchor=east] (1) at (-1.5, 0) {$\sigma_i \sqcup \rho$};
3554
\node[anchor=east] (2) at (-1.5, -2) {$\sigma_{i+1}\ssm \rho$};
3555
%
3556
\node[anchor=west] (4) at (1.5, 0) {$\tau_j$};
3557
\node[anchor=west] (5) at (1.5, -2) {$\tau_{k}$};
3558
%
3559
\draw[thick] (1.east) -- (4.west);
3560
\draw[thick] (1.east) -- (5.west);
3561
\draw[thick] (2.east) -- (5.west);
3562
\node (6) at (0.9, -0.4) {$\rho$};
3563
\node (7) at (0, -1.6) {$b$};
3564
\node (8) at (-1.1, -0.7) {$c$};
3565
\end{tikzpicture}
3566
\end{center}
3567
By the path characterization of the geometric $\SU$ diagonal (\cref{thm:facet-ordering}), the fact that $\tau_j < \tau_k$ implies that $\rho > b$, and the fact that $\sigma_i < \sigma_{i+1}$ implies that $b>c$.
3568
Thus, we have $\rho > \max \sigma_{i,i+1}$, which implies that a path-admissible left shift $R_\rho^{-1}$ can be performed.
3569
3570
In Case $(2.a)$, consider the two adjacent blocks say $\sigma_i|\sigma_{i+1}$ (the $\tau$ case is similar), let $\rho \eqdef \max \sigma_{i,i+1}$ denote the maximum of the path between them, and let $m\geq 1$ be such that $\rho$ steps into $\sigma_{i+1+m}$.
3571
Since $\max \sigma_{i,i+1+m} = \max \sigma_{i,i+1}= \rho$ the definition of the geometric $\SU$ diagonal implies that the block $\sigma_{i+1+m}$ comes after the block $\sigma_{i}$, and by assumption after the block $\sigma_{i+1}$.
3572
Thus, we know that $m > 1$, and using dashed lines to denote paths of length $\geq 1$ we have the following picture
3573
\begin{center}
3574
\begin{tikzpicture}[scale=.6]
3575
\node[anchor=east] (1) at (-1.5, 0) {$\sigma_i$};
3576
\node[anchor=east] (2) at (-1.5, -1.5) {$\sigma_{i+1}$};
3577
\node[anchor=east] (3) at (-1.5, -3) {$\sigma_{i+1+m}$};
3578
%
3579
\node[anchor=west](4) at (1.5, 0) {$\hphantom{.}$};
3580
\node[anchor=west](5) at (1.5, -1.5) {$\hphantom{.}$};
3581
%
3582
\draw[thick,dashed] (1.east) -- (4.west);
3583
\draw[thick,dashed] (2.east) -- (5.west);
3584
\draw[thick] (3.east) -- (4.west);
3585
\draw[thick,dashed] (3.east) -- (5.west);
3586
%
3587
\node (6) at (1.1, -0.9) {$\rho$};
3588
\end{tikzpicture}
3589
\raisebox{2.15em}{$\xrightarrow{R^{-1}_{\rho}}$}
3590
\begin{tikzpicture}[scale=.6]
3591
\node[anchor=east] (1) at (-1.5, 0) {$\sigma_i$};
3592
\node[anchor=east] (2) at (-1.5, -1.5) {$\sigma_{i+1}\sqcup \rho$};
3593
\node[anchor=east] (3) at (-1.5, -3) {$\sigma_{i+1+m}\ssm \rho$};
3594
%
3595
\node[anchor=west] (4) at (1.5, 0) {$\hphantom{.}$};
3596
\node[anchor=west] (5) at (1.5, -1.5) {$\hphantom{.}$};
3597
%
3598
\draw[thick,dashed] (1.east) -- (4.west);
3599
\draw[thick,dashed] (2.east) -- (5.west);
3600
\draw[thick] (2.east) -- (4.west);
3601
\draw[thick,dashed] (3.east) -- (5.west);
3602
%
3603
\node (6) at (1.1, -0.7) {$\rho$};
3604
\end{tikzpicture}
3605
\end{center}
3606
As we have $\rho=\max \sigma_{i,i+1} > \max\sigma_{i+1+m,i+1}$, an inverse $m$-shift operation can be performed.
3607
3608
In the Case $(2.b.i)$, there exists $j,m>0$ such that $\tau_{j-m}$ is the block of $\tau$ which contains the maximal element $\rho$, and $\tau_j$ is any greater block on the path from $\sigma_i$ to $\sigma_{i+1}$.
3609
Then we may apply the following inverse $m$-shift operator.
3610
\begin{center}
3611
\begin{tikzpicture}[scale=.6]
3612
\node[anchor=east] (1) at (-1.5, 0) {$\sigma_i$};
3613
\node[anchor=east] (2) at (-1.5, -2) {$\sigma_{i+1}$};
3614
%
3615
\node[anchor=west] (3) at (1.5, 0) {$\tau_{j-m}$};
3616
\node[anchor=west] (4) at (1.5, -2) {$\tau_{j}$};
3617
%
3618
\draw[thick, dashed] (1.east) -- (4.west);
3619
\draw[thick] (2.east) -- (3.west);
3620
\draw[thick, dashed] (3.west) -- (4.west);
3621
\node (6) at (-0.6, -1.8) {$\rho$};
3622
\end{tikzpicture}
3623
\raisebox{1.8em}{$\xrightarrow{L_\rho^{-1}}$}
3624
\begin{tikzpicture}[scale=.6]
3625
\node[anchor=east] (1) at (-1.5, 0) {$\sigma_i$};
3626
\node[anchor=east] (2) at (-1.5, -2) {$\sigma_{i+1}$};
3627
%
3628
\node[anchor=west] (3) at (1.5, 0) {$\tau_{j-m} \ssm \rho$};
3629
\node[anchor=west] (4) at (1.5, -2) {$\tau_{j} \sqcup \rho$};
3630
%
3631
\draw[thick, dashed] (1.east) -- (4.west);
3632
\draw[thick] (2.east) -- (4.west);
3633
\draw[thick, dashed] (3.west) -- (4.west);
3634
\node (6) at (-0.9, -1.6) {$\rho$};
3635
\end{tikzpicture}
3636
\end{center}
3637
3638
There remains to be treated Case $(2.b.ii)$.
3639
We consider the path $\sigma_{i,i+1} \eqdef (i_1,j_1,i_2,\dots,i_{k},\rho)$ to be of length $2k$, $k>1$.
3640
We denote by $\tau_j$ the block of $\tau$ containing the maximal step $\rho$ of this path, which by hypothesis is the last block of $\tau$.
3641
Let $\sigma_{i-n}$ be the last block of $\sigma$ which is attained by the path $\sigma_{i,i+1}$ before the block $\sigma_{i+1}$.
3642
We must have $n>1$, since $\rho$ is the last step and $\sigma_i|\sigma_{i+1}$ are adjacent blocks.
3643
The situation can be pictured as on the left.
3644
\begin{center}
3645
\begin{tikzpicture}[scale=.6]
3646
\node[anchor=east] (1) at (-1.5, 0) {$\sigma_{i-n}$};
3647
\node[anchor=east] (2) at (-1.5, -1.5) {$\sigma_{i}$};
3648
\node[anchor=east] (3) at (-1.5, -3) {$\sigma_{i+1}$};
3649
%
3650
\node[anchor=west] (4) at (1.5, 0) {$\tau_{j-m}$};
3651
\node[anchor=west] (5) at (1.5, -1.5) {$\tau_j$};
3652
%
3653
\draw[thick,dashed] (1.east) -- (4.west);
3654
\draw[thick] (1.east) -- (5.west);
3655
\draw[thick,dashed] (2.east) -- (4.west);
3656
\draw[thick] (3.east) -- (5.west);
3657
%
3658
\node (6) at (1.1, -2.1) {$\rho$};
3659
\node (7) at (1.1, -0.8) {$i_k$};
3660
\end{tikzpicture}
3661
\raisebox{2.15em}{$\xrightarrow{L_{\rho'}^{-1}}$}
3662
\begin{tikzpicture}[scale=.6]
3663
\node[anchor=east] (1) at (-1.5, 0) {$\sigma_{i-n}$};
3664
\node[anchor=east] (2) at (-1.5, -1.5) {$\sigma_{i}$};
3665
\node[anchor=east] (3) at (-1.5, -3) {$\sigma_{i+1}$};
3666
%
3667
\node[anchor=west] (4) at (1.5, 0) {$\tau_{j-m}\ssm \rho'$};
3668
\node[anchor=west] (5) at (1.5, -1.5) {$\tau_j \sqcup \rho'$};
3669
%
3670
\draw[thick,dashed] (1.east) -- (4.west);
3671
\draw[thick] (1.east) -- (5.west);
3672
\draw[thick,dashed] (2.east) -- (5.west);
3673
\draw[thick] (3.east) -- (5.west);
3674
%
3675
\node (6) at (1.1, -2.1) {$\rho$};
3676
\node (7) at (1.1, -0.8) {$i_k$};
3677
\end{tikzpicture}
3678
\end{center}
3679
Now let $\rho' \eqdef \max\sigma_{i,i-n}$ be the maximum of the path $\sigma_{i,i-n}=(i_1,j_1,i_2,\dots,i_{k-1},j_{k-1})$, and let~$\tau_{j-m}$, $m\geq 1$ be the block of $\tau$ containing $\rho'$.
3680
We want to show that an inverse left $m$-shift can bring $\rho'$ back to $\tau_j$, \ie we need $\rho' > \max \tau_{j-m,j}$.
3681
As $\rho'\eqdef \max \sigma_{i,i-n}$, and apart from the step $i_k$ we have $\tau_{j-m,j}\subset \sigma_{i,i-n}$, we just need to show that $\rho' > i_k$.
3682
To see that this is indeed the case, it suffices to look at the last steps $$\tau_l \overset{j_{k-1}}{\longrightarrow} \sigma_{i-n} \overset{i_{k}}{\longrightarrow} \tau_j \overset{\rho}{\longrightarrow} \sigma_{i+1}$$ of the path $\sigma_{i,i+1}$.
3683
By the definition of the geometric $\SU$ diagonal, the fact that $\tau_l < \tau_j$ implies that $j_{k-1} > i_k$, and thus that $\rho' > i_k$, which finishes the proof.
3684
\end{proof}
3685
3686
\begin{example}
3687
\label{ex:proof-inverse-shift}
3688
Here are examples of each of the cases in the proof of \cref{lem:inverse-to-SCP}.
3689
We display cases $(1)$, $(2.a)$, $(2.b.i)$ and, $(2.b.ii)$ respectively, reading the diagrams from top-left to bottom-right.
3690
\begin{center}
3691
\begin{tikzpicture}[xscale=.6,yscale=0.9]
3692
\node[anchor=east] (1) at (-1.5, -1) {$1$};
3693
\node[anchor=east] (2) at (-1.5, -2) {$23$};
3694
%
3695
\node[anchor=west] (4) at (1.5, -1) {$3$};
3696
\node[anchor=west] (5) at (1.5, -2) {$12$};
3697
%
3698
\draw[thick] (1.east) -- (5.west);
3699
\draw[thick] (2.east) -- (5.west);
3700
\draw[thick] (2.east) -- (4.west);
3701
\end{tikzpicture}
3702
\raisebox{1.3em}{$\xrightarrow{R^{-1}_{3}}$}
3703
\begin{tikzpicture}[xscale=.6,yscale=0.9]
3704
\node[anchor=east] (1) at (-1.5, -1) {$13$};
3705
\node[anchor=east] (2) at (-1.5, -2) {$2$};
3706
%
3707
\node[anchor=west] (4) at (1.5, -1) {$3$};
3708
\node[anchor=west] (5) at (1.5, -2) {$12$};
3709
%
3710
\draw[thick] (1.east) -- (4.west);
3711
\draw[thick] (1.east) -- (5.west);
3712
\draw[thick] (2.east) -- (5.west);
3713
\end{tikzpicture}
3714
,
3715
\begin{tikzpicture}[xscale=.6,yscale=0.9]
3716
\node[anchor=east] (1) at (-1.5, -1) {$1$};
3717
\node[anchor=east] (2) at (-1.5, -2) {$2$};
3718
\node[anchor=east] (3) at (-1.5, -3) {$34$};
3719
%
3720
\node[anchor=west] (4) at (1.5, -1) {$14$};
3721
\node[anchor=west] (5) at (1.5, -2) {$23$};
3722
%
3723
\draw[thick] (1.east) -- (4.west);
3724
\draw[thick] (2.east) -- (5.west);
3725
\draw[thick] (3.east) -- (4.west);
3726
\draw[thick] (3.east) -- (5.west);
3727
\end{tikzpicture}
3728
\raisebox{2.15em}{$\xrightarrow{R^{-1}_{4}}$}
3729
\begin{tikzpicture}[xscale=.6,yscale=0.9]
3730
\node[anchor=east] (1) at (-1.5, -1) {$1$};
3731
\node[anchor=east] (2) at (-1.5, -2) {$24$};
3732
\node[anchor=east] (3) at (-1.5, -3) {$3$};
3733
%
3734
\node[anchor=west] (4) at (1.5, -1) {$14$};
3735
\node[anchor=west] (5) at (1.5, -2) {$23$};
3736
%
3737
\draw[thick] (1.east) -- (4.west);
3738
\draw[thick] (2.east) -- (4.west);
3739
\draw[thick] (2.east) -- (5.west);
3740
\draw[thick] (3.east) -- (5.west);
3741
\end{tikzpicture}
3742
\end{center}
3743
\begin{center}
3744
\begin{tikzpicture}[xscale=.6,yscale=0.9]
3745
\node[anchor=east] (1) at (-1.5, -1) {$13$};
3746
\node[anchor=east] (2) at (-1.5, -2) {$2$};
3747
\node[anchor=east] (3) at (-1.5, -3) {$4$};
3748
%
3749
\node[anchor=west] (4) at (1.5, -1) {$34$};
3750
\node[anchor=west] (5) at (1.5, -2) {$12$};
3751
%
3752
\draw[thick] (1.east) -- (4.west);
3753
\draw[thick] (1.east) -- (5.west);
3754
\draw[thick] (2.east) -- (5.west);
3755
\draw[thick] (3.east) -- (4.west);
3756
\end{tikzpicture}
3757
\raisebox{2.15em}{$\xrightarrow{L^{-1}_{4}}$}
3758
\begin{tikzpicture}[xscale=.6,yscale=0.9]
3759
\node[anchor=east] (1) at (-1.5, -1) {$13$};
3760
\node[anchor=east] (2) at (-1.5, -2) {$2$};
3761
\node[anchor=east] (3) at (-1.5, -3) {$4$};
3762
%
3763
\node[anchor=west] (4) at (1.5, -1) {$3$};
3764
\node[anchor=west] (5) at (1.5, -2) {$124$};
3765
%
3766
\draw[thick] (1.east) -- (4.west);
3767
\draw[thick] (1.east) -- (5.west);
3768
\draw[thick] (2.east) -- (5.west);
3769
\draw[thick] (3.east) -- (5.west);
3770
\end{tikzpicture}
3771
,
3772
\begin{tikzpicture}[xscale=.6,yscale=0.9]
3773
\node[anchor=east] (1) at (-1.5, -1) {$12$};
3774
\node[anchor=east] (2) at (-1.5, -2) {$3$};
3775
\node[anchor=east] (3) at (-1.5, -3) {$4$};
3776
%
3777
\node[anchor=west] (4) at (1.5, -1) {$23$};
3778
\node[anchor=west] (5) at (1.5, -2) {$14$};
3779
%
3780
\draw[thick] (1.east) -- (4.west);
3781
\draw[thick] (1.east) -- (5.west);
3782
\draw[thick] (2.east) -- (4.west);
3783
\draw[thick] (3.east) -- (5.west);
3784
\end{tikzpicture}
3785
\raisebox{2.15em}{$\xrightarrow{L^{-1}_{3}}$}
3786
\begin{tikzpicture}[xscale=.6,yscale=0.9]
3787
\node[anchor=east] (1) at (-1.5, -1) {$12$};
3788
\node[anchor=east] (2) at (-1.5, -2) {$3$};
3789
\node[anchor=east] (3) at (-1.5, -3) {$4$};
3790
%
3791
\node[anchor=west] (4) at (1.5, -1) {$2$};
3792
\node[anchor=west] (5) at (1.5, -2) {$134$};
3793
%
3794
\draw[thick] (1.east) -- (4.west);
3795
\draw[thick] (1.east) -- (5.west);
3796
\draw[thick] (2.east) -- (5.west);
3797
\draw[thick] (3.east) -- (5.west);
3798
\end{tikzpicture}
3799
\end{center}
3800
Note that, as we have chosen minimal illustrative examples, each inverse $m$-shift is an inverse $1$-shift, and after each shift we obtain a $\SCP$.
3801
This is not typically the case.
3802
\end{example}
3803
3804
\begin{theorem}
3805
\label{prop:iso-shift-IJ-diagonals}
3806
The $m$-shift and the geometric $\SU$ diagonals coincide.
3807
\end{theorem}
3808
3809
\begin{proof}
3810
We first note that $\SCP$s are known elements of both the $m$-shift and the geometric $\SU$ diagonals (\cref{lem:SCP-path-desc}).
3811
The proof that every facet of the $m$-shift $\SUD$ is in geometric $\SUD$ follows from the closure of $\SUD$ under the shift operators (\cref{lem:IJ-closed-under-shifts}).
3812
The proof that every facet of geometric $\SUD$ is in shift $\SUD$ follows from the closure of $\SUD$ under the inverse shift operator (\cref{lem:IJ-closed-under-shifts}) and the fact that every facet is sent to a $\SCP$ after a finite number of inverse shifts (\cref{lem:inverse-to-SCP}).
3813
In particular, for any given facet in geometric $\SUD$, this provides a $\SCP$ and a sequence of shifts to form it, showing it is a facet of $m$-shift $\SUD$.
3814
\end{proof}
3815
3816
Combining \cref{prop:iso-original-shift-diagonals} and \cref{prop:iso-shift-IJ-diagonals}, we obtain the desired equivalence between the original $\SU$ diagonal from~\cite{SaneblidzeUmble} and the geometric $\SU$ diagonal from \cref{def:LA-and-SU}.
3817
3818
\begin{theorem}
3819
\label{thm:recover-SU}
3820
The original and geometric $\SU$ diagonals coincide.
3821
\end{theorem}
3822
3823
%%%%%%%%%%%%%%%
3824
3825
\subsection{Shifts under the face poset isomorphism}
3826
\label{subsec:shifts-under-iso}
3827
3828
Having proven the equivalence of the original and geometric $\SU$ diagonals, we now use the face poset isomorphisms between the geometric $\LA$ and $\SU$ diagonals from \cref{subsec:isos-LA-SU} to translate results and combinatorial descriptions from one to the other.
3829
Under the isomorphism $t(r\times r):\LAD\to\SUD$ from \cref{rem:Alternate Isomorphism}, we get a straightforward analogue of \cref{def:SU-admissible} for the $\LA$ diagonal.
3830
Firstly, the morphism $r \times r$ exchanges $\max$ and $\min$, which yields the following ``dual" notions of admissibility.
3831
3832
\begin{definition}
3833
\label{def:LA-admissible}
3834
Let $\sigma$ denote either one of the two ordered partitions of $[n]$ in a $(2,n)$-partition tree, and let $M \subsetneq \sigma_i$.
3835
The right $m$-shift $R_{M}$ (\resp the left $m$-shift $L_{M}$) is
3836
\begin{enumerate}
3837
\item \defn{block-admissible} if $\max \sigma_i \notin M$ and $\max M < \min \sigma_{i+m}$ (\resp $\max M < \min \sigma_{i-m}$),
3838
\item \defn{path-admissible} if $\max M< \min \sigma_{i,i+m}$ (\resp $\max M < \min \sigma_{i,i-m}$).
3839
\end{enumerate}
3840
\end{definition}
3841
3842
Secondly, as the morphism $t$ switches our ordered partitions, this means that the $\LA$ lefts shifts will act on the left ordered partition, and the $\LA$ right shifts will act on the right ordered partition.
3843
Consequently, admissible sequences of $\LA$ shifts are defined similarly to \cref{def:SU-admissible} (simply replace $\sigma$ with $\tau$).
3844
Which provides an analogue of \cref{def:classical-SU} for the $\LA$ diagonal.
3845
3846
\begin{definition}
3847
\label{def:classical-LA}
3848
The facets of the \defn{subset shift, $1$-shift and $m$-shift $\LA$ diagonals} are given by the formula
3849
\begin{align*}
3850
\LAD([n]) = \bigcup_{(\sigma,\tau)} \bigcup_{\mathbf{M}, \mathbf{N}} L_\mathbf{M}(\sigma)\times R_\mathbf{N}(\tau)
3851
\end{align*}
3852
where the unions are taken over all $\SCP$s $(\sigma, \tau)$ of $[n]$, and respecitvely over all block-admissible sequences of subset $1$-shifts $\b{M},\b{N}$, over all path-admissible sequences of singleton $1$-shifts, and over all path-admissible sequences of singleton $m_j$-shifts, for various $m_j \ge 1$.
3853
\end{definition}
3854
3855
We now formally verify that the isomorphism $t(r\times r)$ relates these shift definitions as claimed.
3856
3857
\begin{proposition}
3858
\label{prop:trr is an isomorphism of shifts}
3859
Let $(\sigma,\tau)$ be a facet of $\LAD$.
3860
For each type of $\LA$ shift, let $\b{M},\b{N}$ be admissible sequences of this type, then
3861
\begin{align*}
3862
t(r,r)(L_\mathbf{N}(\sigma), R_\mathbf{M}(\tau)) = (R_{r\mathbf{M}}(r\tau), L_{r\mathbf{N}}(r\sigma))
3863
\end{align*}
3864
where $r\mathbf{M} \eqdef (rM_{1},\dots,rM_{p})$ and $r\mathbf{N}$ (defined similarly) are admissible sequences of $\SU$ shifts of the same type.
3865
\end{proposition}
3866
3867
\begin{proof}
3868
As reversing the elements then shifting them is the same as shifting the elements then reversing them,
3869
it is clear that the equality holds if $r\mathbf{M},r\mathbf{N}$ are admissible sequences of $\SU$ shifts.
3870
As such, we must simply verify the admissibility, and given the equivalence of the various shift definitions, we just do this for path-admissibility of $1$-shifts.
3871
Consider a right shift $R_{M}$, for $M \in \sigma_i$, which is path-admissible in the $\LA$ sense (\cref{def:LA-admissible}).
3872
Then, we have $\max M < \min \sigma_{i,i+1}$ which implies that $\min r M > \max r \sigma_{i,i+1}$, and thus the right shift $R_{rM}$ is path-admissible in the $\SU$ sense (\cref{def:SU-admissible}).
3873
Here, $rM \in r\sigma_i$ is interpreted as being in a block of the right partition ($\tau$ in the notation of the definition).
3874
So, if $\mathbf{M} = (M_{i_1},\dots,M_{i_p})$, define $r(\mathbf{M}) \eqdef (rM_{1},\dots,rM_{p})$, and from the prior it is clear this is a path-admissible sequence of $\SU$ shifts, finishing the proof.
3875
\end{proof}
3876
Thus, given $t(r\times r)$ is an isomorphism of the geometric diagonals (\cref{rem:Alternate Isomorphism}), and the geometric $\SU$ diagonal coincides with the shift $\SU$ diagonals (\cref{subsec:SU-shifts}), we immediately obtain the following statement.
3877
\begin{proposition}
3878
The geometric $\LA$ diagonal and all shift $\LA$ diagonals coincide.
3879
\end{proposition}
3880
3881
\begin{remark}
3882
The isomorphism $rs\times rs:\LAD \to \SUD$, identified in \cref{thm:bijection-operadic-diagonals}, also sends shifts operators to shift operators, but it sends left shift operators to right shift operators and vice versa.
3883
The $\LA^{\op}$ and $\SU^{\op}$ diagonals also admit obvious dual shift descriptions.
3884
\end{remark}
3885
3886
We now explore in example how the isomorphism $t(r\times r)$ translates the shift operators.
3887
3888
\begin{example}
3889
\label{ex:shift translation by theta}
3890
The isomorphism $t(r\times r)$ sends the $\SCP$ $(\sigma,\tau) \eqdef (5|17|4|236,57|146|3|2)$ to the $\SCP$ $(\sigma',\tau') \eqdef (13|247|5|6,3|17|4|256)$.
3891
The corresponding $(2,n)$-partition trees present a clear symmetry
3892
\begin{center}
3893
\begin{tikzpicture}[scale=.7]
3894
\node[anchor=east] (1) at (-1.5, -1) {$5$};
3895
\node[anchor=east] (2) at (-1.5, -2) {$17$};
3896
\node[anchor=east] (3) at (-1.5, -3) {$4$};
3897
\node[anchor=east] (4) at (-1.5, -4) {$236$};
3898
%
3899
\node[anchor=west] (5) at (1.5, -1) {$57$};
3900
\node[anchor=west] (6) at (1.5, -2) {$146$};
3901
\node[anchor=west] (7) at (1.5, -3) {$3$};
3902
\node[anchor=west] (8) at (1.5, -4) {$2$};
3903
%
3904
\draw[thick] (1.east) -- (5.west);
3905
\draw[thick] (2.east) -- (5.west);
3906
\draw[thick] (2.east) -- (6.west);
3907
\draw[thick] (3.east) -- (6.west);
3908
\draw[thick] (4.east) -- (6.west);
3909
\draw[thick] (4.east) -- (7.west);
3910
\draw[thick] (4.east) -- (8.west);
3911
%
3912
\end{tikzpicture}
3913
\raisebox{3.4em}{$\xrightarrow{t(r\times r)}$}
3914
\begin{tikzpicture}[scale=.7]
3915
\node[anchor=east] (13) at (-1.5, -1) {$13$};
3916
\node[anchor=east] (247) at (-1.5, -2) {$247$};
3917
\node[anchor=east] (5) at (-1.5, -3) {$5$};
3918
\node[anchor=east] (6) at (-1.5, -4) {$6$};
3919
%
3920
\node[anchor=west] (3) at (1.5, -1) {$3$};
3921
\node[anchor=west] (17) at (1.5, -2) {$17$};
3922
\node[anchor=west] (4) at (1.5, -3) {$4$};
3923
\node[anchor=west] (256) at (1.5, -4) {$256$};
3924
%
3925
\draw[thick] (13.east) -- (17.west);
3926
\draw[thick] (13.east) -- (3.west);
3927
\draw[thick] (247.east) -- (256.west);
3928
\draw[thick] (247.east) -- (17.west);
3929
\draw[thick] (247.east) -- (4.west);
3930
\draw[thick] (5.east) -- (256.west);
3931
\draw[thick] (6.east) -- (256.west);
3932
%
3933
\end{tikzpicture}
3934
\end{center}
3935
We now illustrate all possible path admissible $\LA$ $1$-shifts of $(\sigma,\tau)$ and all possible path admissible $\SU$ $1$-shifts of $(\sigma',\tau')$.
3936
We first display how the $\LA$ left shifts act on $\sigma$, and the $\SU$ left shifts act on $\tau'$.
3937
The $\LA$ shifts have been drawn so that the leftmost arrow shifts the smallest element, and the $\SU$ shifts have been drawn so that the leftmost arrow shifts the largest element.
3938
As such, the face poset isomorphism $t(r\times r)$ directly translates one diagram to the other.
3939
The specific element being shifted can be inferred by the source and target of the arrow.
3940
\begin{center}
3941
{\small
3942
\begin{tikzcd}
3943
& 5|17|4|236 \arrow[ld] \arrow[d] \arrow[rd]\\
3944
15|7|4|236 \arrow[rd] \arrow[d]&
3945
5|17|24|36 \arrow[dl] \arrow[rd]&
3946
5|17|34|26 \arrow[d] \arrow[dl]\\
3947
15|7|24|36 \arrow[dr]&
3948
15|7|34|26 \arrow[d]&
3949
5|17|234|6 \arrow[dl]\\
3950
&15|7|234|6
3951
\end{tikzcd}
3952
\qquad
3953
\begin{tikzcd}
3954
& 3|17|4|256 \arrow[ld] \arrow[d] \arrow[rd]\\
3955
37|1|4|256 \arrow[rd] \arrow[d]&
3956
3|17|46|25 \arrow[dl] \arrow[rd]&
3957
3|17|44|26 \arrow[d] \arrow[dl]\\
3958
37|1|46|25 \arrow[dr]&
3959
36|17|44|26 \arrow[d]&
3960
3|17|456|2 \arrow[dl]\\
3961
&37|1|456|2
3962
\end{tikzcd}
3963
}
3964
\end{center}
3965
We now illustrate all the possible $\LA$ right shifts acting on $\tau$
3966
\begin{center}
3967
\begin{tikzcd}
3968
\sigma \times 57|146|3|2 \arrow[r,"\rho=1"] & \sigma \times 57|46|13|2 \arrow[r,"\rho=1"] & \sigma \times 57|46|3|12
3969
\end{tikzcd}
3970
\end{center}
3971
and all possible $\SU$ right shifts acting on $\sigma'$
3972
\begin{center}
3973
\begin{tikzcd}
3974
13|247|5|6 \times \tau' \arrow[r,"\rho=7"] & 13|24|57|6 \arrow[r,"\rho=7"] \times \tau' & 13|24|5|67 \times \tau' .
3975
\end{tikzcd}
3976
\end{center}
3977
No other shifts are possible; observe for instance that we cannot perform the $\LA$ left shift $15|7|234|6 \times 57|46|13|2 \xrightarrow{\rho=2} 15|27|34|6 \times 57|46|13|2$ as the minimal path connecting $234$ and $7$ contains $1$, which is smaller than $2$ (see \cref{ex:ECbijection}).
3978
We shall see in \cref{sec:Shift-lattice}, that these diagrams are the Hasse diagrams of lattices.
3979
\end{example}
3980
3981
\begin{example}
3982
It was observed in~\cite{LaplanteAnfossi} that the $\LA$ and $\SU$ diagonals coincided up until $n=3$, however due to their dual shift structure they generate the non-$\SCP$ pairs in a dual fashion.
3983
In particular, the two center faces of the subdivided hexagon of \cref{fig:LUSAdiagonals} are generated by
3984
\begin{center}
3985
\begin{tikzpicture}[scale=.6,xscale=0.6]
3986
\node[anchor=east] (1) at (-1.5, -1) {$13$};
3987
\node[anchor=east] (2) at (-1.5, -2) {$2$};
3988
%
3989
\node[anchor=west] (4) at (1.5, -1) {$3$};
3990
\node[anchor=west] (5) at (1.5, -2) {$12$};
3991
%
3992
\draw[thick] (1.east) -- (4.west);
3993
\draw[thick] (1.east) -- (5.west);
3994
\draw[thick] (2.east) -- (5.west);
3995
\end{tikzpicture}
3996
\quad
3997
\raisebox{1.3em}{$\xrightarrow{R^{\SU}_{3}}$}
3998
\quad
3999
\begin{tikzpicture}[scale=.6,xscale=0.6]
4000
\node[anchor=east] (1) at (-1.5, -1) {$1$};
4001
\node[anchor=east] (2) at (-1.5, -2) {$23$};
4002
%
4003
\node[anchor=west] (4) at (1.5, -1) {$3$};
4004
\node[anchor=west] (5) at (1.5, -2) {$12$};
4005
%
4006
\draw[thick] (1.east) -- (5.west);
4007
\draw[thick] (2.east) -- (5.west);
4008
\draw[thick] (2.east) -- (4.west);
4009
\end{tikzpicture}
4010
\qquad \qquad
4011
\begin{tikzpicture}[scale=.6,xscale=0.6]
4012
\node[anchor=east] (1) at (-1.5, -1) {$12$};
4013
\node[anchor=east] (2) at (-1.5, -2) {$3$};
4014
%
4015
\node[anchor=west] (4) at (1.5, -1) {$2$};
4016
\node[anchor=west] (5) at (1.5, -2) {$13$};
4017
%
4018
\draw[thick] (1.east) -- (4.west);
4019
\draw[thick] (2.east) -- (5.west);
4020
\draw[thick] (1.east) -- (5.west);
4021
\end{tikzpicture}
4022
\quad
4023
\raisebox{1.3em}{$\xrightarrow{L^{\SU}_{3}}$}
4024
\quad
4025
\begin{tikzpicture}[scale=.6,xscale=0.6]
4026
\node[anchor=east] (1) at (-1.5, -1) {$12$};
4027
\node[anchor=east] (2) at (-1.5, -2) {$3$};
4028
%
4029
\node[anchor=west] (4) at (1.5, -1) {$23$};
4030
\node[anchor=west] (5) at (1.5, -2) {$1$};
4031
%
4032
\draw[thick] (1.east) -- (4.west);
4033
\draw[thick] (2.east) -- (4.west);
4034
\draw[thick] (1.east) -- (5.west);
4035
\end{tikzpicture}
4036
\end{center}
4037
and
4038
\begin{center}
4039
\begin{tikzpicture}[scale=.6,xscale=0.6]
4040
\node[anchor=east] (1) at (-1.5, -1) {$1$};
4041
\node[anchor=east] (2) at (-1.5, -2) {$23$};
4042
%
4043
\node[anchor=west] (4) at (1.5, -1) {$13$};
4044
\node[anchor=west] (5) at (1.5, -2) {$2$};
4045
%
4046
\draw[thick] (1.east) -- (4.west);
4047
\draw[thick] (2.east) -- (4.west);
4048
\draw[thick] (2.east) -- (5.west);
4049
\end{tikzpicture}
4050
\quad
4051
\raisebox{1.3em}{$\xrightarrow{R^{\LA}_{1}}$}
4052
\quad
4053
\begin{tikzpicture}[scale=.6,xscale=0.6]
4054
\node[anchor=east] (1) at (-1.5, -1) {$1$};
4055
\node[anchor=east] (2) at (-1.5, -2) {$23$};
4056
%
4057
\node[anchor=west] (4) at (1.5, -1) {$3$};
4058
\node[anchor=west] (5) at (1.5, -2) {$12$};
4059
%
4060
\draw[thick] (1.east) -- (5.west);
4061
\draw[thick] (2.east) -- (4.west);
4062
\draw[thick] (2.east) -- (5.west);
4063
\end{tikzpicture}
4064
\qquad \qquad
4065
\begin{tikzpicture}[scale=.6,xscale=0.6]
4066
\node[anchor=east] (1) at (-1.5, -1) {$2$};
4067
\node[anchor=east] (2) at (-1.5, -2) {$13$};
4068
%
4069
\node[anchor=west] (4) at (1.5, -1) {$23$};
4070
\node[anchor=west] (5) at (1.5, -2) {$1$};
4071
%
4072
\draw[thick] (1.east) -- (4.west);
4073
\draw[thick] (2.east) -- (4.west);
4074
\draw[thick] (2.east) -- (5.west);
4075
\end{tikzpicture}
4076
\quad
4077
\raisebox{1.3em}{$\xrightarrow{L^{\LA}_{1}}$}
4078
\quad
4079
\begin{tikzpicture}[scale=.6,xscale=0.6]
4080
\node[anchor=east] (1) at (-1.5, -1) {$12$};
4081
\node[anchor=east] (2) at (-1.5, -2) {$3$};
4082
%
4083
\node[anchor=west] (4) at (1.5, -1) {$23$};
4084
\node[anchor=west] (5) at (1.5, -2) {$1$};
4085
%
4086
\draw[thick] (1.east) -- (4.west);
4087
\draw[thick] (2.east) -- (4.west);
4088
\draw[thick] (1.east) -- (5.west);
4089
\end{tikzpicture}
4090
\end{center}
4091
where we have aligned each element, of each diagonal, vertically with its dual element.
4092
\end{example}
4093
4094
%%%%%%%%%%%%%%%
4095
4096
\subsection{Shift lattices}
4097
\label{sec:Shift-lattice}
4098
4099
In this section, we show that the $1$-shifts of the operadic diagonals~$\LAD$ and~$\SUD$, define the covering relations of a lattice structure on the set of facets.
4100
More precisely, we show that each $\SCP$ is the minimal element of a lattice isomorphic to a product of chains, where the partial order is given by shifts.
4101
Given the bijection between $\SCP$s and permutations, and our prior enumeration of the facets of the diagonal, this produces two new statistics on permutations.
4102
In addition, later in \cref{sec:Cubical}, we will use the lattice structure to relate the cubical and shift definitions of the $\SUD$ diagonal.
4103
4104
\begin{definition}
4105
\label{def:shift-lattice}
4106
The $\LA$ \defn{shift poset} on the set of facets $(\sigma,\tau) \in \LAD$ is defined to be the transitive closure of the relations $ (\sigma,\tau) \prec (L_\rho \sigma,\tau)$ and $ (\sigma,\tau) \prec (\sigma,R_\rho \tau)$, for all $\LA$ path-admissible $1$-shifts $L_{\rho}$ and $R_\rho$.
4107
The $\SU$ \defn{shift poset} is defined similarly.
4108
Then, for each permutation $v$ of $[n]$, we define the subposet \defn{$\hour^\LA$} (\resp \defn{$\hour^\SU$}) to be the set of all admissible $\LA$ (\resp $\SU$) shifts of the associated~$\SCP$.
4109
\end{definition}
4110
4111
Given the shift definitions of $\SUD$ and $\LAD$ (\cref{def:classical-SU,def:classical-LA}), the subposets $\hour^\LA$ and~$\hour^\SU$ are clearly connected components of their posets, with a unique minimal element corresponding to the $\SCP$.
4112
We now aim to prove they are also lattices (\cref{prop:shift lattice}).
4113
4114
\begin{lemma}
4115
\label{lem:m-shifts commute}
4116
The $m$-shift operators defining the facets of an operadic diagonal commute.
4117
That is, whenever the successive composition of two $m_1$, $m_2$-shifts is defined on a facet of the $\LA$ or $\SU$ diagonal, the reverse order of composition is also defined, and the two yield the same facet.
4118
\end{lemma}
4119
\begin{proof}
4120
A $\SUD$ (\resp $\LAD$) $m$-shift is defined when $\rho$ is greater (\resp smaller) than the maximal (\resp minimal) element of the connecting path.
4121
Combining \cref{cor:SU-1-shift-preserves-max} and \cref{prop:iso-1-to-m-shift}, we know $m$-shifts conserve the maximal (minimal) elements of paths and their directions.
4122
As such, we can commute any two shift operators.
4123
\end{proof}
4124
4125
\begin{definition}
4126
\label{def:heights}
4127
Let $v$ be a permutation of $[n]$, and $(\sigma,\tau)$ the SCP corresponding to $v$.
4128
For $\rho \in [n]$, we define the $\LA$ \defn{left height} and \defn{right height} of $\rho$ to be
4129
\begin{align*}
4130
\ell_v(\rho) \eqdef \max (\{0\}\cup \{m>0: L_{\rho}(\sigma) \text{ is a path admissible $\LA$ $m$-shift} \}), \\
4131
r_v(\rho) \eqdef \max (\{0\}\cup \{m>0: R_{\rho}(\tau) \text{ is a path admissible $\LA$ $m$-shift} \}).
4132
\end{align*}
4133
The left and right heights for the $\SU$ diagonal are defined similarly.
4134
\end{definition}
4135
4136
The height of an element $\rho$ in a $\SCP$ can be explicitly calculated as follows.
4137
4138
\begin{lemma}
4139
\label{prop:maximal m-shift formulae}
4140
Let $(\sigma,\tau)$ be the $\SCP$ corresponding to a permutation $v$.
4141
Then, the right $\LA$ height $r_v(\rho)$ (\resp the left $\LA$ height $\ell_v(\rho)$) of $\rho \in [n]$ is given by the number of consecutive blocks of $\sigma$ (\resp of~$\tau$) to the right (\resp left) of the one containing $\rho$ whose minima are larger than $\rho$.
4142
The $\SU$ heights are obtained similarly by considering blocks whose maxima are smaller than $\rho$.
4143
\end{lemma}
4144
4145
\begin{proof}
4146
We consider the right $\SU$ height, the other cases are similar.
4147
As $(\sigma,\tau)$ is a $\SCP$, we have $\max \sigma_{k,k+1} = \max \sigma_{k+1}$ for all $k\geq 1$.
4148
Moreover, from the equivalence between $1$-shifts and $m$-shifts (\cref{prop:iso-1-to-m-shift}), there exists a $m$-shift of $\rho$ from $\sigma_i$ to $\sigma_{i+m}$ if, and only if, there exists a sequence of $m$ consecutive $1$-shifts, each satisfying $\rho > \max \sigma_{j,j+1}=\max \sigma_{j+1}$.
4149
Thus, these iterated $1$-shifts will be path-admissible until the first failure at $j=r_v(\rho)+1$.
4150
\end{proof}
4151
4152
\begin{remark}
4153
The height calculations can also be reformulated directly in terms of the permutation.
4154
For instance, for the $\SU$ diagonal $r_v(\rho)$ is the number of consecutive descending runs, to the right of the descending run containing $\rho$, whose maximal element is smaller than $\rho$.
4155
\end{remark}
4156
4157
\begin{proposition}
4158
\label{prop:shift lattice}
4159
The subposets~$\hour^\LA$ and~$\hour^\SU$, are lattices isomorphic to products of chains
4160
\begin{align*}
4161
\hour^\LA \cong \prod_{\rho \in [n]} [0,\ell_v(\rho)] \times \prod_{\rho \in [n]} [0,r_v(\rho)]
4162
\quad \text{ and } \quad
4163
\hour^\SU \cong \prod_{\rho \in [n]} [0,r_v(\rho)] \times \prod_{\rho \in [n]} [0,\ell_v(\rho)] \ ,
4164
\end{align*}
4165
where $[0,k]$ is the chain lattice $0<1<\dots<k$, for $k\geq 0$.
4166
\end{proposition}
4167
4168
\begin{proof}
4169
We denote by $L_\rho^m$ (\resp $R_\rho^m$) a left (right) $m$-shift of $\rho$ for $m>0$, and we let it be the identity if~$m=0$.
4170
By the commutativity of $m$-shift operators (\cref{lem:m-shifts commute}), and the existence of unique heights for each element (\cref{prop:maximal m-shift formulae}), every element of $\hour^\LA$ admits a unique shift description $(L^{\ell_n}_{n} \dots L^{\ell_1}_{1}(\sigma),
4171
R^{r_n}_{n}\cdots R^{r_{1}}_{1}(\tau))$, where $0\leq \ell_\rho\leq \ell_v(\rho)$ and $0\leq r_\rho\leq r_v(\rho)$.
4172
Thus, we identify it with the pair of tuples $(\ell_1,\ldots,\ell_n)\times (r_1,\ldots,r_n)$.
4173
This is clearly an isomorphism of lattices.
4174
\end{proof}
4175
4176
Note that the maximal element of $\hour$ (for either diagonal) is given by shifting each element of $(\sigma,\tau)$ by its maximal shift.
4177
The joins and meets of any two elements are also thus given by isomorphism to the product of chains.
4178
For instance, in the case of meets,
4179
\begin{multline*}
4180
(\ell_1,\ldots,\ell_n,r_1,\ldots,r_n)\land (\ell'_1,\ldots,\ell'_n,r'_1,\ldots,r'_n) = \\ (\min\{\ell_1,\ell'_1\},\ldots,\min\{\ell_n,\ell'_n\},\min\{r_1,r'_1\},\ldots,\min\{r_n,r'_n\})
4181
\end{multline*}
4182
For clear examples of the Hasse diagrams corresponding to our lattices, we direct the reader to \cref{ex:shift translation by theta}, and \cref{fig: Inversion and lattice counter example}.
4183
We note that \cref{fig: Inversion and lattice counter example} also illustrates that there is no general relation between the shift lattice structure and inversions sets.
4184
In particular, the shift lattices are not sub-lattices of the facial weak order (discussed in \cref{sec:facial-weak-order}), as $24|13$ and $234|1$ are incomparable.
4185
4186
\begin{figure}
4187
{\footnotesize
4188
\begin{tikzcd}[column sep=tiny]
4189
&
4190
\begin{tikzpicture}[xscale=.4,yscale=0.7]
4191
\node[anchor=east] (1) at (-1.5, -1) {$12$};
4192
\node[anchor=east] (2) at (-1.5, -2) {$3$};
4193
\node[anchor=east] (3) at (-1.5, -3) {$4$};
4194
%
4195
\node[anchor=west](4) at (1.5, -1) {$234$};
4196
\node[anchor=west](5) at (1.5, -2) {$1$};
4197
%
4198
\draw[thick] (1.east) -- (4.west);
4199
\draw[thick] (1.east) -- (5.west);
4200
\draw[thick] (2.east) -- (4.west);
4201
\draw[thick] (3.east) -- (4.west);
4202
\end{tikzpicture}
4203
\begin{tikzpicture}[xscale=.4,yscale=0.7]
4204
\node[anchor=east, circle, draw=black, thick] (1) at (-1, -1) {$1$};
4205
\node[anchor=east, circle, draw=black, thick] (2) at (-1, -3) {$2$};
4206
\node[anchor=west, circle, draw=black, thick] (3) at (1, -1) {$3$};
4207
\node[anchor=west, circle, draw=black, thick] (4) at (1, -3) {$4$};
4208
%
4209
\draw[thick, green] (1) -- (3);
4210
\draw[thick, green] (1) -- (4);
4211
\draw[thick, green] (2) -- (3);
4212
\draw[thick, green] (2) -- (4);
4213
\draw[thick, green] (3) -- (4);
4214
%
4215
\draw[ultra thick, blue, dotted] (1) -- (2);
4216
\draw[ultra thick, blue, dotted] (1) -- (3);
4217
\draw[ultra thick, blue, dotted] (1) -- (4);
4218
\end{tikzpicture}
4219
\\
4220
\begin{tikzpicture}[xscale=.4,yscale=0.7]
4221
\node[anchor=east] (1) at (-1.5, -1) {$12$};
4222
\node[anchor=east] (2) at (-1.5, -2) {$3$};
4223
\node[anchor=east] (3) at (-1.5, -3) {$4$};
4224
%
4225
\node[anchor=west](4) at (1.5, -1) {$23$};
4226
\node[anchor=west](5) at (1.5, -2) {$14$};
4227
%
4228
\draw[thick] (1.east) -- (4.west);
4229
\draw[thick] (1.east) -- (5.west);
4230
\draw[thick] (2.east) -- (4.west);
4231
\draw[thick] (3.east) -- (5.west);
4232
\end{tikzpicture}
4233
\begin{tikzpicture}[xscale=.4,yscale=0.7]
4234
\node[anchor=east, circle, draw=black, thick] (1) at (-1, -1) {$1$};
4235
\node[anchor=east, circle, draw=black, thick] (2) at (-1, -3) {$2$};
4236
\node[anchor=west, circle, draw=black, thick] (3) at (1, -1) {$3$};
4237
\node[anchor=west, circle, draw=black, thick] (4) at (1, -3) {$4$};
4238
%
4239
\draw[thick, green] (1) -- (3);
4240
\draw[thick, green] (1) -- (4);
4241
\draw[thick, green] (2) -- (3);
4242
\draw[thick, green] (2) -- (4);
4243
\draw[thick, green] (3) -- (4);
4244
%
4245
\draw[ultra thick, blue, dotted] (1) -- (2);
4246
\draw[ultra thick, blue, dotted] (1) -- (3);
4247
\end{tikzpicture}
4248
\arrow[ur]
4249
&
4250
&
4251
\begin{tikzpicture}[xscale=.4,yscale=0.7]
4252
\node[anchor=east] (1) at (-1.5, -1) {$12$};
4253
\node[anchor=east] (2) at (-1.5, -2) {$3$};
4254
\node[anchor=east] (3) at (-1.5, -3) {$4$};
4255
%
4256
\node[anchor=west](4) at (1.5, -1) {$24$};
4257
\node[anchor=west](5) at (1.5, -2) {$13$};
4258
%
4259
\draw[thick] (1.east) -- (4.west);
4260
\draw[thick] (1.east) -- (5.west);
4261
\draw[thick] (2.east) -- (5.west);
4262
\draw[thick] (3.east) -- (4.west);
4263
\end{tikzpicture}
4264
\begin{tikzpicture}[xscale=.4,yscale=0.7]
4265
\node[anchor=east, circle, draw=black, thick] (1) at (-1, -1) {$1$};
4266
\node[anchor=east, circle, draw=black, thick] (2) at (-1, -3) {$2$};
4267
\node[anchor=west, circle, draw=black, thick] (3) at (1, -1) {$3$};
4268
\node[anchor=west, circle, draw=black, thick] (4) at (1, -3) {$4$};
4269
%
4270
\draw[thick, green] (1) -- (3);
4271
\draw[thick, green] (1) -- (4);
4272
\draw[thick, green] (2) -- (3);
4273
\draw[thick, green] (2) -- (4);
4274
\draw[thick, green] (3) -- (4);
4275
%
4276
\draw[ultra thick, blue, dotted] (1) -- (2);
4277
\draw[ultra thick, blue, dotted] (1) -- (4);
4278
\draw[ultra thick, blue, dotted] (3) -- (4);
4279
\end{tikzpicture}
4280
\arrow[ul]
4281
\\
4282
&
4283
\begin{tikzpicture}[xscale=.4,yscale=0.7]
4284
\node[anchor=east] (1) at (-1.5, -1) {$12$};
4285
\node[anchor=east] (2) at (-1.5, -2) {$3$};
4286
\node[anchor=east] (3) at (-1.5, -3) {$4$};
4287
%
4288
\node[anchor=west](4) at (1.5, -1) {$2$};
4289
\node[anchor=west](5) at (1.5, -2) {$134$};
4290
%
4291
\draw[thick] (1.east) -- (4.west);
4292
\draw[thick] (1.east) -- (5.west);
4293
\draw[thick] (2.east) -- (5.west);
4294
\draw[thick] (3.east) -- (5.west);
4295
\end{tikzpicture}
4296
\begin{tikzpicture}[xscale=.4,yscale=0.7]
4297
\node[anchor=east, circle, draw=black, thick] (1) at (-1, -1) {$1$};
4298
\node[anchor=east, circle, draw=black, thick] (2) at (-1, -3) {$2$};
4299
\node[anchor=west, circle, draw=black, thick] (3) at (1, -1) {$3$};
4300
\node[anchor=west, circle, draw=black, thick] (4) at (1, -3) {$4$};
4301
%
4302
\draw[thick, green] (1) -- (3);
4303
\draw[thick, green] (1) -- (4);
4304
\draw[thick, green] (2) -- (3);
4305
\draw[thick, green] (2) -- (4);
4306
\draw[thick, green] (3) -- (4);
4307
%
4308
\draw[ultra thick, blue, dotted] (1) -- (2);
4309
\end{tikzpicture}
4310
\arrow[ul, to path= (\tikztostart.210) -- (\tikztotarget.south)]
4311
\arrow[ur, to path= (\tikztostart.330) -- (\tikztotarget.south)]
4312
\end{tikzcd}
4313
}
4314
\caption{The shift lattice $\hour^\SU$ for $v=4|3|1|2$.
4315
Each facet is drawn next to a graph encoding its inversions (\cref{def:inverions}).
4316
If $(i,j) \in J(\sigma)$, then a green edge connects $(i,j)$, and if $(i,j) \in I(\tau)$, then a blue dotted edge connects $(i,j)$.
4317
Consequently, $I((\sigma,\tau))$ is encoded by the presence of both edges, and also the crossings, by \cref{p:crossings}.
4318
}
4319
\label{fig: Inversion and lattice counter example}
4320
\end{figure}
4321
4322
\begin{remark}
4323
As a consequence of \cref{prop:shift lattice}, the facets of the operadic diagonals are disjoint unions of lattices.
4324
However, any lattice $L$ on permutations (such as the weak order) induces a lattice on the facets as follows.
4325
For every $v \in L$, we can substitute the lattice $\hour^\LA$ (or $\hour^\SU$) into the permutation $v$.
4326
In particular, every element which was covered by $v$ is now covered by the minimal element of $\hour^\LA$, and every element which was covering $v$ now covers the maximal element of $\hour^\LA$.
4327
\end{remark}
4328
4329
Given our previously obtained formulae for the number of elements in the diagonal (\cref{subsec:enumerationDiagonalPermutahedra}), and the results of this section, we obtain the following statistics on permutations.
4330
4331
\begin{corollary}
4332
\label{cor:statistics-lattice}
4333
Using the heights of either diagonal,
4334
\begin{align*}
4335
2(n+1)^{n-2} = \sum_{v \in \mathbb{S}_n} \prod_{\rho \in [n]} (\ell_v(\rho)+1)(r_v(\rho)+1)
4336
\end{align*}
4337
Moreover, denoting by $\mathbb{S}_n^{k_1} \subseteq \mathbb{S}_n$ the set of permutations with $k_1$ ascending runs, and consequently $k_2 = n-1-k_1$ descending runs, we have
4338
\begin{align*}
4339
n \binom{n-1}{k_1} (n-k_1)^{k_1-1} (n-k_2)^{k_2-1} = \sum_{v \in \mathbb{S}_n^{k_1}} \prod_{\rho \in [n]} (\ell_v(\rho)+1)(r_v(\rho)+1).
4340
\end{align*}
4341
\end{corollary}
4342
4343
\begin{proof}
4344
This follows directly from \cref{coro:enumerationDiagonalPermutahedra}, with the observation that shifts conserve the number of blocks, and hence the dimensions of the faces.
4345
\end{proof}
4346
4347
%%%%%%%%%%%%%%%
4348
4349
\subsection{Cubical description}
4350
\label{sec:Cubical}
4351
4352
In this section, we recall the cubical definition of the $\SU$ diagonal from~\cite{SaneblidzeUmble-comparingDiagonals} and explicitly relate it to their shift description, using a new proof that exploits the lattice description of the diagonal (\cref{sec:Shift-lattice}).
4353
Then we construct an analogous cubical definition of the $\LA$ diagonal, transferring the cubical formulae via isomorphism.
4354
4355
\subsubsection{The cubical $\SU$ diagonal}
4356
4357
We define inductively a subdivision $\divcube{n-1}^\SU$ of the $(n-1)$-dimensional cube which is combinatorially isomorphic to the permutahedron $\Perm$ (\cref{prop:subdiv cube Combinatorially Isomorphic to perm}).
4358
4359
\begin{construction}
4360
\label{constr:cubicPermutahedron1}
4361
Given a $(n-k)$-dimensional face $\sigma = \sigma_1| \cdots |\sigma_k$ of the $(n-1)$-dimensional permutahedron $\Perm[n]$, we set $n_j \eqdef \card{\sigma_{k-j+1}\cup\dots\cup \sigma_k}$, and define a subdivision $I_\sigma \eqdef I_1 \cup \dots \cup I_k$ of the interval $[0,1]$ by the following formulas
4362
\[
4363
I_j \eqdef
4364
\begin{cases}
4365
\left[0,1 - 2^{-n_j}\right] & \text{ if } j=1, \\
4366
\left[1 - 2^{-n_{j-1}}, 1 - 2^{-n_{j}}\right] & \text{ if } 1 < j < k, \\
4367
\left[1 - 2^{-n_{j-1}},1\right] & \text{ if } j=k .
4368
\end{cases}
4369
\]
4370
Let $\divcube{0}^\SU$ be the $0$-dimensional cube (a point), trivially subdivided by the sole element $1$ of~$\Perm[1]$.
4371
Then, assuming we have constructed the subdivision $\divcube{n-1}^\SU$ of the $(n-1)$-cube, we construct $\divcube{n}^\SU$ as the subdivision of $\divcube{n-1}^\SU \times [0,1]$ given, for each face $\sigma$ of $\divcube{n-1}^\SU$, by the polytopal complex $\sigma \times I_\sigma$.
4372
We label the faces $\sigma \times I$ of the subdivided rectangular prism $\sigma \times I_\sigma$ by the following rule
4373
\begin{align}
4374
\label{eq:sub}
4375
\sigma \times I \eqdef
4376
\begin{cases}
4377
\sigma_1| \cdots |\sigma_k| n+1 & \text{if } I = \{0\}, \\
4378
\sigma_1| \cdots |\sigma_j| n+1 |\sigma_{j+1}| \cdots |\sigma_k & \text{if } I = I_j \cap I_{j+1} \text{ with } 1 \leq j \leq k-1 , \\
4379
n+1|\sigma_1| \cdots |\sigma_k & \text{if } I = \{1\}, \\
4380
\sigma_1| \cdots |\sigma_j \cup \{n+1\}| \cdots |\sigma_k & \text{if } I = I_j \text{ with } 1\leq j \leq k.
4381
\end{cases}
4382
\tag{I}
4383
\end{align}
4384
This defines a subdivision $\divcube{n}^\SU$ of the $n$-cube.
4385
\end{construction}
4386
4387
\cref{fig:cubicPermutahedron1} illustrates this subdivision for the first few dimensions.
4388
We indicate, in bold, the embedding $\divcube{n-1}^\SU\hookrightarrow \divcube{n}^\SU$ induced by the natural embedding $\R^{n-1} \hookrightarrow \R^n$.
4389
Only the vertices of~$\divcube{3}^\SU$ are labelled, but its edges, facets and outer face are all identified with the expected elements of $\Perm[4]$.
4390
%
4391
\begin{figure}[h!]
4392
\begin{center}
4393
{\small
4394
\begin{tikzcd}[sep=0.3cm]
4395
\mathbf{1|2} \arrow[r,dash, "12"] & 2|1
4396
\end{tikzcd}
4397
}
4398
$\hookrightarrow$
4399
{\small
4400
\begin{tikzcd}[column sep=0.3cm]
4401
3|1|2 \arrow[r,dash, "3|12"] \arrow[d,dash, "13|2"]& 3|2|1 \arrow[d,dash, "13|2"]\\
4402
1|3|2 \arrow[d,dash, "1|23"] & 2|3|1 \arrow[d,dash, "2|13"]\\
4403
\mathbf{1|2|3} \arrow[r,dash,thick, "12|3"] & \mathbf{2|1|3}
4404
\end{tikzcd}
4405
}
4406
$\hookrightarrow$
4407
{\small
4408
\begin{tikzcd}[sep=0.15cm,scale=0.8]
4409
& 4|2|1|3 \arrow[rr,dash] \arrow[dd,dash,dotted] \arrow[dl,dash] & & 4|2|3|1 \arrow [dd, dash,dotted] \arrow[rr,dash] & & 4|3|2|1 \arrow[dd,dash] \arrow[dl,dash]\\
4410
4|1|2|3 \arrow [rr,dash] \arrow [dd,dash] & & 4|1|3|2 \arrow[rr,dash] \arrow[dd,dash] & & 4|3|1|2 \arrow[dd,dash] & \\
4411
& 2|4|1|3 \arrow [rr,dash,dotted] \arrow[dd,dash,dotted] & & 2|4|3|1 \arrow[dd,dash,dotted] & & 3|4|2|1 \arrow [dl,dash] \arrow[dd,dash]\\
4412
1|4|2|3 \arrow [rr,dash] \arrow [dd,dash]& & 1|4|3|2 \arrow[dd,dash] & & 3|4|1|2 \arrow[dd,dash]\\
4413
& 2|1|4|3 \arrow[dd,dash,dotted] \arrow[dl,dash,dotted] & & 2|3|4|1 \arrow [dd, dash,dotted] \arrow [rr, dash,dotted] & & 3|2|4|1 \arrow[dd,dash] \\
4414
1|2|4|3 \arrow [dd,dash] & & 1|3|4|2 \arrow[rr,dash] \arrow[dd,dash] & & 3|1|4|2 \arrow[dd,dash] & \\
4415
& \mathbf{2|1|3|4} \arrow [rr,dash,dotted] \arrow [dl,dash,dotted] & & \mathbf{2|3|1|4} \arrow [rr,dash,dotted] & & \mathbf{3|2|1|4} \arrow [dl,dash] \\
4416
\mathbf{1|2|3|4} \arrow [rr,dash] & & \mathbf{1|3|2|4} \arrow [rr,dash] & & \mathbf{3|1|2|4} \\
4417
\end{tikzcd}
4418
}
4419
\end{center}
4420
\caption{Cubical realizations $\divcube{1}^\SU,\divcube{2}^\SU$ and $\divcube{3}^\SU$ of the permutahedra~$\Perm[2]$, $\Perm[3]$ and $\Perm[4]$, respectively, from \cref{constr:cubicPermutahedron1}.}
4421
\label{fig:cubicPermutahedron1}
4422
\end{figure}
4423
4424
\begin{remark}
4425
\label{rem:coordinates}
4426
A consequence of the construction is that each edge of $\divcube{n}^\SU$ is parallel to one of the canonical basis vectors $e_i$ of $\R^n$, and corresponds to shifting the element $i+1$ of $[n+1]\ssm \{1\}$.
4427
\end{remark}
4428
4429
\begin{proposition}
4430
\label{prop:subdiv cube Combinatorially Isomorphic to perm}
4431
The polytopal complex $\divcube{n}^\SU$ is combinatorially isomorphic to the permutahedron $\Perm[n+1]$.
4432
\end{proposition}
4433
4434
\begin{proof}
4435
By construction it is clear that the faces of $\divcube{n}^\SU$ and $\Perm[n+1]$ are in bijection, and that this bijection preserves the dimension.
4436
It remains to show that this bijection is a poset isomorphism.
4437
Let $\sigma_1 | \ldots | \sigma_i | \sigma_{i+1} | \ldots | \sigma_k \prec \sigma_1 | \ldots | \sigma_i \cup \sigma_{i+1} | \ldots | \sigma_k$ be a covering relation in the face poset of $\Perm[n+1]$.
4438
We need to see that the corresponding faces $F,G$ of $\divcube{n}^\SU$ satisfy $F \prec G$.
4439
From \cref{eq:sub} this clearly holds for lines.
4440
Since any face of $\divcube{n}^\SU$ is a product of lines, the result follows by induction on the dimension of the faces.
4441
\end{proof}
4442
4443
We now unpack how certain properties of $\Perm$ have been encoded in the cubical structure of $\divcube{n}^\SU$.
4444
This will later allow us to construct a cubical formula for the diagonal through maximal pairings of $k$-subdivision cubes of $\divcube{n}^\SU$, which we now introduce.
4445
4446
\begin{definition}
4447
\label{def:Subdivisions}
4448
For $k\geq 0$, a \defn{$k$-subdivision cube} of $\divcube{n}^\SU$ is a union of $k$-faces of $\divcube{n}^\SU$ whose underlying set is a $k$-dimensional rectangular prism.
4449
\end{definition}
4450
4451
An important example of a $k$-subdivision cubes are the $k$-faces of $\divcube{n}^\SU$ (other examples are provided in \cref{ex:subdivision cubes}).
4452
4453
\begin{lemma}
4454
\label{lem:k-subdiv cubes have max/min k faces}
4455
A $k$-subdivision cube has a unique maximal (\resp minimal) vertex with respect to the weak order on permutations.
4456
\end{lemma}
4457
4458
\begin{proof}
4459
By construction, the edges of $\divcube{n}^\SU$ are parallel to the basis vectors of $\R^n$ (\cref{rem:coordinates}), and correspond to inversions on permutations.
4460
Thus, the vector $\b{v}\eqdef(1,\ldots,1)$ induces the weak order on the vertices of $\divcube{n}^\SU$.
4461
Since each $k$-subdivision cube is a rectangular prism whose edges are not perpendicular to $\b{v}$, the scalar product with $\b{v}$ is maximized (\resp minimized) at a unique vertex.
4462
\end{proof}
4463
4464
\begin{definition}
4465
The \defn{maximal (\resp minimal) $k$-face} of a $k$-subdivision cube, with respect to the weak order, is the unique $k$-face in the subdivision cube which contains the maximal (\resp minimal) vertex.
4466
\end{definition}
4467
4468
\begin{construction}
4469
\label{const:unique-sub-cube}
4470
For a $k$-dimensional face $\sigma$ of the cubical permutahedron $\divcube{n}^\SU$, we construct the unique maximal $k$-subdivision cube, with respect to inclusion, whose maximal (\resp minimal) $k$-face with respect to the weak order is $\sigma$.
4471
\end{construction}
4472
4473
We only treat the case for the maximal $k$-face $\sigma$, the minimal face proceeds similarly.
4474
We build the maximal subdivision cubes inductively.
4475
Let $\sigma$ be an edge of $\divcube{n}^\SU$, and let $v$ be its maximal vertex.
4476
Let $\rho \in [n+1]\ssm\{1\}$ be the element shifted by this edge (\cref{rem:coordinates}).
4477
Shifting this element to the right as far as possible ($\rho$ will be shifted all the way to the right, or be blocked by a larger element), we get the desired $1$-subdivision line.
4478
4479
Suppose that we have constructed maximal subdivision cubes up to dimension $k$, and let $\sigma$ be a $(k+1)$-face of $\divcube{n}^\SU$, with maximal vertex $v$.
4480
Consider the $k+1$ elements of $[n]$, which correspond to dimensions spanned by $\sigma$ (\cref{rem:coordinates}), and let $\rho$ be the largest such element.
4481
Let $\sigma_L$ be the $1$-face, with maximal vertex~$v$, whose only non-trivial dimension corresponds to~$\rho$.
4482
From the initial step above, there is a unique maximal $1$-subdivision line $L$ with maximal $1$-face~$\sigma_L$.
4483
Let $\sigma_C$ be the $k$-face, with maximal vertex $v$, spanned by the complement of $\rho$ in $[n+1] \ssm \{1\}$.
4484
By induction, there is a maximal $k$-subdivision cube $C$ with maximal $k$-face $\sigma_C$.
4485
Then, it is clear that the product $L\times C$ defines a $(k+1)$-subdivision cube of $\divcube{n}^\SU$, with maximal vertex~$v$.
4486
Indeed, as $\rho$ is the maximal element corresponding to dimensions of $L\times C$, the faces of $L\times C$ are the $(k+1)$-dimensional rectangular prisms resulting from shifting $\rho$ through the $k$-faces of $C$, as in \cref{eq:sub}.
4487
4488
Finally, $L\times C$ is maximal under inclusion.
4489
If there was a larger $(k+1)$-subdivision cube enveloping $L\times C$, then one of its projections would be a $1$ or $k$-subdivision cube enveloping $L$ or $C$, contradicting the assumption that they are maximal.
4490
This finishes the construction.
4491
4492
\begin{definition}
4493
\label{def:hourglass}
4494
For a vertex $v$ of $\divcube{n}^\SU$, the \defn{hourglass} of $\divcube{n}^\SU$ at $v$ is the maximal pair of subdivision cubes $\hour^\SU$, with respect to inclusion, within the set of all pairs $(C,C')$ of subdivision cubes such that $C$ has maximal vertex $v$, $C'$ has minimal vertex $v$, and $\dim C +\dim C' = n$.
4495
\end{definition}
4496
4497
\begin{figure}[h!]
4498
\begin{center}
4499
{\small
4500
\begin{tikzcd}[sep=0.1cm, scale=0.3]
4501
& 4|2|1|3 \arrow[rr,dash] \arrow[dd,dash,dotted] \arrow[dl,dash] & & 4|2|3|1 \arrow [dd, dash,dotted] \arrow[rr,dash] & & 4|3|2|1 \arrow[dd,dash] \arrow[dl,dash, near end, "\mathbf{\blue{4|3|12}}"']\\
4502
4|1|2|3 \arrow [rr,dash] \arrow [dd,dash] & & 4|1|3|2 \arrow[rr,dash] \arrow[dd,dash] & & \mathbf{\blue{4|3|1|2}} \arrow[dd,dash] & \\
4503
& \red{\mathbf{14|23}} \arrow [rr,dash,dotted] \arrow[dd,dash,dotted] & & 2|4|3|1 \arrow[dd,dash,dotted] & & 3|4|2|1 \arrow [dl,dash] \arrow[dd,dash]\\
4504
1|4|2|3 \arrow [rr,dash] \arrow [dd,dash]& & 1|4|3|2 \arrow[dd,dash] \arrow[rr, phantom, "\blue{\mathbf{134|2}}" description,crossing over] & & 3|4|1|2 \arrow[dd,dash]\\
4505
& 2|1|4|3 \arrow[dd,dash,dotted] \arrow[dl,dash,dotted] & & 2|3|4|1 \arrow [dd, dash,dotted] \arrow [rr, dash,dotted] & & 3|2|4|1 \arrow[dd,dash] \\
4506
1|2|4|3 \arrow [dd,dash] \arrow[rr, phantom, "\red{\mathbf{1|234}}" description,crossing over] & & 1|3|4|2 \arrow[rr,dash] \arrow[dd,dash] & & 3|1|4|2 \arrow[dd,dash] & \\
4507
& 2|1|3|4 \arrow [rr,dash,dotted] \arrow [dl,dash,dotted] & & \red{\mathbf{13|24}} \arrow [rr,dash,dotted] & & 3|2|1|4 \arrow [dl,dash] \\
4508
1|2|3|4 \arrow [rr,dash] & & 1|3|2|4 \arrow [rr,dash] & & 3|1|2|4 \\
4509
\end{tikzcd}
4510
}
4511
\end{center}
4512
\caption{The hourglass~$\hour^\SU$ of $\divcube{3}^\SU$ at~$v = \blue{\mathbf{4|3|1|2}}$.}
4513
\label{fig:hourglass1}
4514
\end{figure}
4515
4516
The following examples are pictured in \cref{fig:hourglass1}.
4517
4518
\begin{example}
4519
\label{ex:subdivision cubes}
4520
The sets of faces~$\{1234\}$, $\{1|234\}$, $\{1|234, 14|23\}$, and $\{1|3|24,1|34|2\}$ are all subdivision cubes of $\divcube{3}$.
4521
In contrast, the sets of faces $\{134|2, 14|23\}$, $\{1|234, 134|2, 14|23\}$, and $\{1|2|34,1|23|4\}$
4522
are not subdivision cubes of $\divcube{3}^\SU$.
4523
For $v = \blue{\mathbf{4|3|1|2}}$, the only $1$-subdivision cube with minimal vertex $v$ is $4|3|12$, and the three $2$-subdivision cubes with maximal vertex $v$ are $\{134|2\}$, $\{13|24, 134|2\}$, $\{ 1|234, 13|24, 14|23, 134|2\}$.
4524
This defines the hourglass $\hour^\SU=\{ 1|234, 13|24, 14|23, 134|2\} \times \{4|3|12\}$ of $\divcube{3}^\SU$ at $v$.
4525
4526
Let us observe that the $\SCP$ corresponding to $v$ is $(\sigma,\tau) \eqdef (\blue{\mathbf{134|2}},\blue{\mathbf{4|3|12}} )$.
4527
The ordered partition $\sigma$ admits three distinct rights shifts, $\red{\mathbf{13|24}}$, $\red{\mathbf{14|23}}$, $\red{\mathbf{1|234}}$, and $\tau$ admits no left shifts.
4528
\cref{thm:cubical-SU} shows that $\hour^\SU$ is generated by all shifts of the $\SCP$ corresponding to $v$.
4529
\end{example}
4530
4531
\begin{theorem}
4532
\label{thm:cubical-SU}
4533
Let $v$ be a vertex of the cubical permutahedron $\divcube{n}^\SU$, and let $(\sigma,\tau)$ be its associated~$\SCP$.
4534
Then, we have
4535
\[
4536
\hour^\SU = \bigcup_{\b{M},\b{N}}R_{\b{M}}(\sigma) \times L_{\b{N}}(\tau) \ ,
4537
\]
4538
where the union is taken over all block-admissible sequences of $\SU$ shifts $\b{M},\b{N}$.
4539
\end{theorem}
4540
4541
\begin{proof}
4542
We prove the result inductively from lines, and consider the case of $\sigma$, $\tau$ proceeds similarly.
4543
Combining \cref{const:unique-sub-cube} with \cref{prop:maximal m-shift formulae}, we have that if $L$ is a maximal $1$-subdivision cube with maximal $1$-face $\sigma$, then its faces are generated by the right $1$-shifts $R_\rho^i$, for $i$ between $0$ and its maximal right height~$r_\rho$ (\cref{def:heights}).
4544
4545
Now consider the unique maximal $(k+1)$-subdivision cube of $\divcube{n}^\SU$ with maximal $(k+1)$-face~$\sigma$.
4546
By \cref{const:unique-sub-cube}, this subdivision cube is given by the product $L\times C$ of a line corresponding to the maximal element $\rho$ being shifted, and a $k$-cube $C$ corresponding to all other elements.
4547
By induction hypothesis, both the line $L$ and the cube $C$ are generated by all block-admissible right shifts from their unique maximal faces $\sigma_L$ and $\sigma_C$.
4548
Moreover, if $\rho$ is in the $i$th block of $\sigma_C$, then $\sigma$ is obtained from~$\sigma_C$ by merging the $i$\ordinal{} and $(i+1)$\ordinalst{} blocks.
4549
4550
On the one hand, the right height of any element being shifted in $C$ is the same as its right height in $L\times C$.
4551
This follows from the inductive description of $\divcube{n}^\SU$ (\cref{eq:sub}): every $k$-face in $C$ has $\rho$ in a singleton block, and as $\rho$ is larger than all other elements it blocks all other shifts.
4552
4553
On the other hand, we know from \cref{const:unique-sub-cube} that the $(k+1)$-faces of $L\times C$ are obtained by weaving $\rho$ through the $k$-faces of~$C$.
4554
Indeed, every $k$-face in $C$ has $\rho$ in a fixed singleton set, and every $k$-face on the opposite side of $L\times C$ has $\rho$ in another fixed singleton set; given \cref{eq:sub}, this final singleton block in any $k$-face of $C$ is either the last block, or is followed by a block containing an element larger than $\rho$.
4555
If we translate this observation back to the $(k+1)$-faces of $L\times C$ that are adjacent to the boundary of $L$, then this corresponds to an equivalent calculation of the right height of $\rho$ in $\sigma$.
4556
4557
Given the lattice description of the diagonal (\cref{prop:shift lattice}), we thus have that $L\times C$ is generated by all block-admissible sequences of right shifts of $\sigma$, which concludes the proof.
4558
\end{proof}
4559
4560
This recovers the formulas \cite[Form.~(1)~\&~(3)]{SaneblidzeUmble-comparingDiagonals}.
4561
4562
%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4563
4564
\subsubsection{$\LA$ cubical description}
4565
4566
The $\LA$ diagonal also admits a similar cubical description, which we may quickly induce by isomorphism.
4567
We first define inductively a subdivision $\divcube{n-1}^{\LA}$ of the $(n-1)$-dimensional cube which is combinatorially isomorphic to the permutahedron $\Perm$, analogous to the one from the preceding sections.
4568
4569
\begin{construction}
4570
\label{const:cubical-LA}
4571
Given a $(n-k)$-dimensional face $\sigma = \sigma_1| \cdots |\sigma_k$ of the $(n-1)$-dimensional permutahedron $\Perm[n]$, we set $n_j \eqdef \card{\sigma_{k-j+1}\cup\dots\cup \sigma_k}$, and define a subdivision $I_\sigma \eqdef I_1 \cup \dots \cup I_k$ of the interval $[0,1]$ by the same formulas as in \cref{constr:cubicPermutahedron1}.
4572
4573
Let $\divcube{0}^\LA$ be the $0$-dimensional cube (a point), trivially subdivided by the sole element $1$ of~$\Perm[1]$.
4574
Then, assuming we have constructed the subdivision $\divcube{n-1}^\LA$ of the $(n-1)$-cube, we construct $\divcube{n}^\LA$ as the subdivision of $\divcube{n-1}^\LA \times [0,1]$ given, for each face $\sigma$ of $\divcube{n-1}^\LA$, by the polytopal complex $\sigma \times I_\sigma$.
4575
We label the faces $\sigma \times I$ of the subdivided rectangular prism $\sigma \times I_\sigma$ by the following rule
4576
\begin{align*}
4577
\sigma \times I \eqdef
4578
\begin{cases}
4579
\sigma_1'| \cdots |\sigma_k'| 1 & \text{if } I = \{0\}, \\
4580
\sigma_1'| \cdots |\sigma_j'| 1 |\sigma_{j+1}'| \cdots |\sigma_k' & \text{if } I = I_j \cap I_{j+1} \text{ with } 1 \leq j \leq k-1 , \\
4581
1|\sigma_1'| \cdots |\sigma_k' & \text{if } I = \{1\}, \\
4582
\sigma_1'| \cdots |\sigma_j' \cup \{1\}| \cdots |\sigma_k' & \text{if } I = I_j \text{ with } 1\leq j \leq k ,
4583
\end{cases}
4584
\end{align*}
4585
where each block of $\sigma$ has been renumbered as $\sigma_i' \eqdef \{p+1 \ | \ p \in \sigma_i\}$ for all $1 \leq i \leq k$.
4586
We obtain a subdivision $\divcube{n}^\LA$ of the $n$-cube isomorphic to the permutahedron $\Perm[n+1]$.
4587
\end{construction}
4588
4589
\begin{figure}[h!]
4590
\begin{center}
4591
{\small
4592
\begin{tikzcd}[sep=0.3cm]
4593
\mathbf{2|1} \arrow[r,dash, "12"] & 1|2
4594
\end{tikzcd}
4595
}
4596
$\hookrightarrow$
4597
{\small
4598
\begin{tikzcd}[column sep=0.3cm]
4599
1|3|2 \arrow[r,dash, "1|23"] \arrow[d,dash, "13|2"]& 1|2|3 \arrow[d,dash, "13|2"]\\
4600
3|1|2 \arrow[d,dash, "3|12"] & 2|1|3 \arrow[d,dash, "2|13"]\\
4601
\mathbf{3|2|1} \arrow[r,dash,thick, "23|1"] & \mathbf{2|3|1}
4602
\end{tikzcd}
4603
}
4604
$\hookrightarrow$
4605
{\small
4606
\begin{tikzcd}[sep=0.15cm,scale=0.8]
4607
& 1|3|4|2 \arrow[rr,dash] \arrow[dd,dash,dotted] \arrow[dl,dash] & & 1|3|2|4 \arrow [dd, dash,dotted] \arrow[rr,dash] & & 1|2|3|4 \arrow[dd,dash] \arrow[dl,dash]\\
4608
1|4|3|2 \arrow [rr,dash] \arrow [dd,dash] & & 1|4|2|3 \arrow[rr,dash] \arrow[dd,dash] & & 1|2|4|3 \arrow[dd,dash] & \\
4609
& 3|1|4|2 \arrow [rr,dash,dotted] \arrow[dd,dash,dotted] & & 3|1|2|4 \arrow[dd,dash,dotted] & & 2|1|3|4 \arrow [dl,dash] \arrow[dd,dash]\\
4610
4|1|3|2 \arrow [rr,dash] \arrow [dd,dash]& & 4|1|2|3 \arrow[dd,dash] & & 2|1|4|3 \arrow[dd,dash]\\
4611
& 3|4|1|2 \arrow[dd,dash,dotted] \arrow[dl,dash,dotted] & & 3|2|1|4 \arrow [dd, dash,dotted] \arrow [rr, dash,dotted] & & 2|3|1|4 \arrow[dd,dash] \\
4612
4|3|1|2 \arrow [dd,dash] & & 4|2|1|3 \arrow[rr,dash] \arrow[dd,dash] & & 2|4|1|3 \arrow[dd,dash] & \\
4613
& \mathbf{3|4|2|1} \arrow [rr,dash,dotted] \arrow [dl,dash,dotted] & & \mathbf{3|2|4|1} \arrow [rr,dash,dotted] & & \mathbf{2|3|4|1} \arrow [dl,dash] \\
4614
\mathbf{4|3|2|1} \arrow [rr,dash] & & \mathbf{4|2|3|1} \arrow [rr,dash] & & \mathbf{2|4|3|1} \\
4615
\end{tikzcd}
4616
}
4617
\end{center}
4618
\caption{Cubical realizations of the permutahedra~$\Perm[2]$, $\Perm[3]$ and $\Perm[4]$ from \cref{const:cubical-LA}.}
4619
\label{fig:cubicPermutahedron2}
4620
\end{figure}
4621
4622
\cref{fig:cubicPermutahedron2} illustrates $\divcube{n}^{\LA}$ in dimensions $1$ to $3$.
4623
We have indicated in bold the embedding ${\divcube{n-1}^{\LA}\hookrightarrow \divcube{n}^{\LA}}$ induced by the natural inclusions $\R^n \hookrightarrow \R^{n+1}$.
4624
4625
The appropriate base definitions for the $\LA$ diagonal of $k$-subdivision cubes (\cref{def:Subdivisions}) and hourglasses ${\hour}$ (\cref{def:hourglass}) are the same as in the $\SU$ case.
4626
The proof that $\hour^\LA $ is indeed combinatorially isomorphic to $\Perm[n+1]$ proceeds similarly to \cref{prop:subdiv cube Combinatorially Isomorphic to perm}.
4627
4628
Recall from \cref{subsec:isos-LA-SU} the face poset isomorphism of the permutahedra which acts on~$A_1| \cdots |A_k$, by replacing each block $A_j$ by the block $r(A_j)\eqdef\set{n-i+1}{i \in A_j}$.
4629
We have the analogue of \cref{thm:cubical-SU} for the $\LA$ diagonal.
4630
4631
\begin{theorem}
4632
\label{prop:LA-cubical}
4633
Let $v$ be a vertex of the cubical permutahedron $\divcube{n}$, and let $(\sigma,\tau)$ be its associated~$\SCP$.
4634
Then, we have
4635
\[
4636
\hour^\LA = \bigcup_{\b{M},\b{N}}L_{\b{M}}(\sigma) \times R_{\b{N}}(\tau) \ ,
4637
\]
4638
where the union is taken over all block-admissible sequences of $\LA$ shifts $\b{M},\b{N}$.
4639
\end{theorem}
4640
4641
\begin{proof}
4642
It is straightforward to see that the involution $r$ induces a $k$-subdivision cube isomorphism $r:\divcube{n}^{\SU}\to \divcube{n}^{\LA}$ between the cubical subdivisions of \cref{constr:cubicPermutahedron1} and \cref{const:cubical-LA}, which sends the hourglass $\hour^\SU$ to the hourglass $\hour[r(v)]^\LA$.
4643
By \cref{thm:cubical-SU}, we know that $\hour^\SU$ is generated by $\SU$ shifts; we want to deduce that $\hour[r(v)]^\LA$ is generated by $\LA$ shifts.
4644
First, we observe that the diagram
4645
\begin{center}
4646
\begin{tikzcd}
4647
\SCP \arrow[r,"t(r\times r)",leftrightarrow] \arrow[d,leftrightarrow] & \SCP \arrow[d,leftrightarrow]\\
4648
\mathbb{S}_n \arrow[r,"r"',leftrightarrow] & \mathbb{S}_n ,
4649
\end{tikzcd}
4650
\end{center}
4651
where $t$ is the permutation of the two factors, and the vertical arrows are the bijection between $\SCP$s and permutations (\cref{def:strong-complementary-pairs}), is commutative.
4652
Thus, if we start from a $\SCP$ and consider its associated $\SU$ hourglass $\hour^\SU$, applying the subdivision cube isomorphism $r$ or applying the map $t(r \times r)$ both give the $\LA$ hourglass $\hour[r(v)]^\LA$.
4653
Combining this with the fact that the map $t(r\times r):\LAD\to \SUD$ is an isomorphism between the $\LA$ and $\SU$ diagonal (\cref{rem:Alternate Isomorphism}) which preserves left and right shifts (\cref{prop:trr is an isomorphism of shifts}), we obtain the desired result.
4654
\end{proof}
4655
4656
\begin{figure}[h!]
4657
\centerline{
4658
{\small
4659
\begin{tikzcd}[sep=0.1cm, scale=0.3, ampersand replacement=\&]
4660
\& 4|2|1|3 \arrow[rr,dash] \arrow[dd,dash,dotted] \arrow[dl,dash] \& \& 4|2|3|1 \arrow [dd, dash,dotted] \arrow[rr,dash] \& \& 4|3|2|1 \arrow[dd,dash] \arrow[dl,dash, near end, "\mathbf{\blue{4|3|12}}"']\\
4661
4|1|2|3 \arrow [rr,dash] \arrow [dd,dash] \& \& 4|1|3|2 \arrow[rr,dash] \arrow[dd,dash] \& \& \mathbf{\blue{4|3|1|2}} \arrow[dd,dash] \& \\
4662
\& \red{\mathbf{14|23}} \arrow [rr,dash,dotted] \arrow[dd,dash,dotted] \& \& 2|4|3|1 \arrow[dd,dash,dotted] \& \& 3|4|2|1 \arrow [dl,dash] \arrow[dd,dash]\\
4663
1|4|2|3 \arrow [rr,dash] \arrow [dd,dash]\& \& 1|4|3|2 \arrow[dd,dash] \arrow[rr, phantom, "\blue{\mathbf{134|2}}" description,crossing over] \& \& 3|4|1|2 \arrow[dd,dash]\\
4664
\& 2|1|4|3 \arrow[dd,dash,dotted] \arrow[dl,dash,dotted] \& \& 2|3|4|1 \arrow [dd, dash,dotted] \arrow [rr, dash,dotted] \& \& 3|2|4|1 \arrow[dd,dash] \\
4665
1|2|4|3 \arrow [dd,dash] \arrow[rr, phantom, "\red{\mathbf{1|234}}" description,crossing over] \& \& 1|3|4|2 \arrow[rr,dash] \arrow[dd,dash] \& \& 3|1|4|2 \arrow[dd,dash] \& \\
4666
\& 2|1|3|4 \arrow [rr,dash,dotted] \arrow [dl,dash,dotted] \& \& \red{\mathbf{13|24}} \arrow [rr,dash,dotted] \& \& 3|2|1|4 \arrow [dl,dash] \\
4667
1|2|3|4 \arrow [rr,dash] \& \& 1|3|2|4 \arrow [rr,dash] \& \& 3|1|2|4 \\
4668
\end{tikzcd}
4669
}
4670
$\overset{r}{\longrightarrow}$
4671
{\small
4672
\begin{tikzcd}[sep=0.1cm, scale=0.3, ampersand replacement=\&]
4673
\& 1|3|4|2 \arrow[rr,dash] \arrow[dd,dash,dotted] \arrow[dl,dash] \& \& 1|3|2|4 \arrow [dd, dash,dotted] \arrow[rr,dash] \& \& 1|2|3|4 \arrow[dd,dash] \arrow[dl,dash, near end, "\mathbf{\blue{1|2|34}}"']\\
4674
1|4|3|2 \arrow [rr,dash] \arrow [dd,dash] \& \& 1|4|2|3 \arrow[rr,dash] \arrow[dd,dash] \& \& \mathbf{\blue{1|2|4|3}} \arrow[dd,dash] \& \\
4675
\& \red{\mathbf{14|23}} \arrow [rr,dash,dotted] \arrow[dd,dash,dotted] \& \& 3|1|2|4 \arrow[dd,dash,dotted] \& \& 2|1|3|4 \arrow [dl,dash] \arrow[dd,dash]\\
4676
4|1|3|2 \arrow [rr,dash] \arrow [dd,dash]\& \& 4|1|2|3 \arrow[dd,dash] \arrow[rr, phantom, "\blue{\mathbf{124|3}}" description,crossing over] \& \& 2|1|4|3 \arrow[dd,dash]\\
4677
\& 3|4|1|2 \arrow[dd,dash,dotted] \arrow[dl,dash,dotted] \& \& 3|2|1|4 \arrow [dd, dash,dotted] \arrow [rr, dash,dotted] \& \& 2|3|1|4 \arrow[dd,dash] \\
4678
4|3|1|2 \arrow [dd,dash] \arrow[rr, phantom, "\red{\mathbf{4|123}}" description,crossing over] \& \& 4|2|1|3 \arrow[rr,dash] \arrow[dd,dash] \& \& 2|4|1|3 \arrow[dd,dash] \& \\
4679
\& 3|4|2|1 \arrow [rr,dash,dotted] \arrow [dl,dash,dotted] \& \& \red{\mathbf{24|13}} \arrow [rr,dash,dotted] \& \& 2|3|4|1 \arrow [dl,dash] \\
4680
4|3|2|1 \arrow [rr,dash] \& \& 4|2|3|1 \arrow [rr,dash] \& \& 2|4|3|1 \\
4681
\end{tikzcd}
4682
}
4683
}
4684
\caption{The isomorphism~$r$ applied to the $\SU$ cubical subdivision from \cref{fig:hourglass1}.}
4685
\label{fig:hourglass2}
4686
\end{figure}
4687
4688
\begin{example}
4689
Applying the isomorphism $r$ to \cref{ex:subdivision cubes} yields the illustration of \cref{fig:hourglass2}.
4690
As $r$ is an isomorphism of $k$-subdivision cubes, the maximal pair of $\SU$ subdivision cubes has been mapped to a maximal pair of $\LA$ subdivision cubes.
4691
Note that the maximal $\SU$ $2$-subdivision cube with maximal vertex $4|3|1|2$ was generated by $\SU$ right shifts.
4692
Its image under $r$ is the maximal $\LA$ $2$-subdivision cube with minimal vertex $1|2|4|3$, and it is generated by $\LA$ right shifts.
4693
\end{example}
4694
4695
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4696
4697
\subsection{Matrix description}
4698
\label{subsec:matrix}
4699
4700
For completeness, we recall from \cite{SaneblidzeUmble} the matrix description of facets of the $\SU$ diagonal.
4701
We previously saw that $\SCP$s and permutations are in bijection (\cref{def:strong-complementary-pairs}).
4702
There is also a third equivalent way to encode this data, the \defn{step matrices} of \cite[Def. 6]{SaneblidzeUmble}.
4703
Given a permutation, one defines its associated step matrix by starting in the bottom left corner, writing increasing sequences vertically and decreasing sequences horizontally one after the other, leaving all other entries $0$.
4704
See \cref{fig:step-matrix}.
4705
4706
\begin{figure}[h!]
4707
\begin{center}
4708
\raisebox{3em}{$6|5|2|4|7|1|3$}
4709
$\quad \quad$
4710
\begin{tikzpicture}[scale=.7]
4711
\node[anchor=east] (1) at (-1.5, -1) {$256$};
4712
\node[anchor=east] (2) at (-1.5, -2) {$4$};
4713
\node[anchor=east] (3) at (-1.5, -3) {$17$};
4714
\node[anchor=east] (4) at (-1.5, -4) {$3$};
4715
%
4716
\node[anchor=west] (5) at (1.5, -1) {$6$};
4717
\node[anchor=west] (6) at (1.5, -2) {$5$};
4718
\node[anchor=west] (7) at (1.5, -3) {$247$};
4719
\node[anchor=west] (8) at (1.5, -4) {$13$};
4720
%
4721
\draw[thick] (1.east) -- (5.west);
4722
\draw[thick] (1.east) -- (6.west);
4723
\draw[thick] (1.east) -- (7.west);
4724
\draw[thick] (2.east) -- (7.west);
4725
\draw[thick] (3.east) -- (7.west);
4726
\draw[thick] (3.east) -- (8.west);
4727
\draw[thick] (4.east) -- (8.west);
4728
\end{tikzpicture}
4729
$\quad \quad$
4730
\raisebox{3em}{
4731
$
4732
\begin{blockarray}{ccccc}
4733
& \sigma_1 & \sigma_2 & \sigma_3 & \sigma_4 \\
4734
\begin{block}{c[cccc]}
4735
\tau_4 & & & 1 & 3 \\
4736
\tau_3 & 2 & 4 & 7 & \\
4737
\tau_2 & 5 & & & \\
4738
\tau_1 & 6 & & & \\
4739
\end{block}
4740
\end{blockarray}
4741
$
4742
}
4743
\end{center}
4744
\caption{A permutation, its associated $\SCP$, and their step matrix.}
4745
\label{fig:step-matrix}
4746
\end{figure}
4747
Given a matrix $A$ whose only non-zero entries are the elements $[n]$, let $\sigma_i(A)$ denote the non-zero entries of the $i$\ordinal{} column, and $\tau_j(A)$ the non-zero entries of the $(r-j+1)$\ordinalst{} row, where $r$ is the number of rows of $A$. See the labelling in \cref{fig:step-matrix}.
4748
With this identification, the definitions of the shift operators can be translated directly:
4749
the right shift operator $R_{M}$ shifts the elements of a subset $M \subset \sigma_i(A)$ one column to the right, or one row up, replacing only elements of value $0$, and leaving $0$ elements in their wake, while the left shift operator $L_{M}$ shifts the elements of $M$ to the left, or down one row.
4750
4751
The fact that the shifts avoid collisions with other elements is a consequence of their admissibility.
4752
Recall from \cref{def:movable-subsets}, that a right $\SU$ $1$-shift $R_M$ is \defn{block-admissible} if $\min \sigma_i \notin M$ and $\min M > \max \sigma_{i+1}$, and that admissible sequence of right shifts proceed in increasing order (\cref{def:SU-admissible}).
4753
4754
\begin{proposition}
4755
Admissible sequences of matrix shift operators are well-defined.
4756
\end{proposition}
4757
4758
\begin{proof}
4759
We verify the claim that the admissible sequences of matrix shift operators never replace non-zero elements.
4760
It is straightforward to show that this is true when a shift operator is applied to a $\SCP$ $(\sigma,\tau)$.
4761
From here we proceed inductively.
4762
We assume that all prior shift operators have been well-defined, and we then check that applying another admissible shift operator is also well-defined.
4763
Suppose that an admissible right shift $R_{M}$ is not well-defined, as it tries to move a value $m$ into a non-zero matrix entry $n$.
4764
Then, $n$ must have been placed into that column by a prior left shift operator $L_{N_j}$, and consequently $n>\min N_j > \max \tau_{j-1}> m$.
4765
However, $n\in \sigma_{i+1}$ so $\max \sigma_{i+1}>n>m>\min M_i $ implies that $M$ is not block-admissible, a contradiction.
4766
\end{proof}
4767
4768
\begin{figure}[h!]
4769
\centerline{
4770
\begin{tikzcd}[column sep = 0.5cm, ampersand replacement=\&]
4771
\begin{bmatrix}
4772
& & 1 & 3 \\
4773
2 & 4 & 7 & \\
4774
5 & & & \\
4775
6 & & & \\
4776
\end{bmatrix}
4777
\arrow[r,"R^{\SU}_{56}"]
4778
\&
4779
\begin{bmatrix}
4780
& & 1 & 3 \\
4781
2 & 4 & 7 & \\
4782
& 5 & & \\
4783
& 6 & & \\
4784
\end{bmatrix}
4785
\arrow[r,"L^{\SU}_7"]
4786
\&
4787
\begin{bmatrix}
4788
& & 1 & 3 \\
4789
2 & 4 & & \\
4790
& 5 & 7 & \\
4791
& 6 & & \\
4792
\end{bmatrix}
4793
\arrow[r,"R^{\SU}_7"]
4794
\&
4795
\begin{bmatrix}
4796
& & 1 & 3 \\
4797
2 & 4 & & \\
4798
& 5 & & 7 \\
4799
& 6 & & \\
4800
\end{bmatrix}
4801
\\[-.3cm]
4802
\begin{bmatrix}
4803
& & & 2 \\
4804
& & & 3 \\
4805
& 1 & 4 & 6 \\
4806
5 & 7 & & \\
4807
\end{bmatrix}
4808
\arrow[r,"R^{\LA}_{23}"]
4809
\&
4810
\begin{bmatrix}
4811
& & 2 & \\
4812
& & 3 & \\
4813
& 1 & 4 & 6 \\
4814
5 & 7 & & \\
4815
\end{bmatrix}
4816
\arrow[r,"L^{\LA}_1"]
4817
\&
4818
\begin{bmatrix}
4819
& & 2 & \\
4820
& 1 & 3 & \\
4821
& & 4 & 6 \\
4822
5 & 7 & & \\
4823
\end{bmatrix}
4824
\arrow[r,"R^{\LA}_1"]
4825
\&
4826
\begin{bmatrix}
4827
& & 2 & \\
4828
1 & & 3 & \\
4829
& & 4 & 6 \\
4830
5 & 7 & & \\
4831
\end{bmatrix}
4832
\end{tikzcd}
4833
}
4834
\caption{Matrix shifts under the isomorphism $t(r \times r)$ between the $\LA$ and $\SU$ diagonals.}
4835
\label{fig:matrix-shifts}
4836
\end{figure}
4837
4838
\defn{Configuration matrices} \cite[Def. 7]{SaneblidzeUmble} are the matrices corresponding to $\SCP$s and those generated by admissible sequences of shifts.
4839
Consequently, they are in bijection with the facets of the $\SU$ diagonal.
4840
The translation of these results for the $\LA$ diagonal is clear.
4841
One can use the isomorphism $t(r\times r)$ as in the following example.
4842
4843
\begin{example}
4844
\label{ex:matrix shifts}
4845
The first row of \cref{fig:matrix-shifts} contains a sequence of admissible $\SU$ subset shifts applied to the matrix encoding of the SCP $256|4|17|3\times 6|5|147|13$.
4846
The second row is the image of these shifts under the isomorphism $t(r\times r)$.
4847
Note that the shifts of this example are also isomorphic to those of \cref{ex:shift translation by theta}, under the isomorphism $(rs\times rs)$.
4848
\end{example}
4849
4850
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4851
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4852
4853
\clearpage
4854
4855
\part{Higher algebraic structures}
4856
\label{part:higherAlgebraicStructures}
4857
4858
In this third part, we derive some higher algebraic consequences of the results obtained in \cref{part:diagonalsPermutahedra}.
4859
We first prove in \cref{subsec:top-operadic-structures} that there are exactly two topological operad structures on the family of operahedra (\resp multiplihedra) which are compatible with the generalized Tamari order, and thus two geometric universal tensor products of (non-symmetric non-unital) homotopy operads (\resp $\Ainf$-morphisms).
4860
Then, we show that these topological operad structures are isomorphic (\cref{subsec:iso-top-operads}).
4861
However, these isomorphisms do not commute with the diagonal maps (\cref{ex:iso-not-Hopf,ex:iso-not-Hopf-2}).
4862
Finally, we show that contrary to the case of permutahedra, the faces of the $\LA$ and $\SU$ diagonals of the operahedra (\resp multiplihedra) are in general not in bijection (\cref{subsec:tensor-products}).
4863
However, from a homotopical point of view, the two tensor products of homotopy operads (\resp $\Ainf$-morphisms) that they define are $\infty$-isomorphic (\cref{thm:infinity-iso,thm:infinity-iso-2}).
4864
4865
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4866
4867
\section{Higher tensor products}
4868
4869
%%%%%%%%%%%%%%%
4870
4871
\subsection{Topological operadic structures}
4872
\label{subsec:top-operadic-structures}
4873
4874
The permutahedra are part of a more general family of polytopes called Loday realizations of the \emph{operahedra}~\cite[Def.~2.9]{LaplanteAnfossi}, which encodes the notion of homotopy operad~\cite[Def.~4.11]{LaplanteAnfossi} (we consider here only \emph{non-symmetric non-unital} homotopy operads).
4875
Let $\PT_n$ be the set of planar trees with $n$ internal edges, which are labelled by $[n]$ using the infix order.
4876
For every planar tree $t$, there is a corresponding operahedron $P_t$ whose codimension~$k$ faces are in bijection with nestings of $t$ with $k$ non-trivial nests.
4877
4878
\begin{definition}[{\cite[Def.~2.1 \& 2.22]{LaplanteAnfossi}}]\label{def:nesting}
4879
A \defn{nest} of $t \in \PT_n$ is a subset of internal edges which induce a subtree, and a \defn{nesting} of $t$ is a family of nests which are either included in one another, or disjoint.
4880
See \cref{fig:nestings}.
4881
\end{definition}
4882
4883
\begin{figure}[h!]
4884
\begin{tikzpicture}[yscale=-1, every node/.style={draw, very thick, circle, inner sep=1pt}, edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]
4885
]
4886
\node(0) {0}
4887
child{node(1){1}}
4888
child{node(2){2}
4889
child{node(3){3}
4890
child{node(4){4}}
4891
}}
4892
child{node(5){5}};
4893
% \begin{pgfonlayer}{bg} % select the background layer
4894
\hedge[blue, very thick]{0,2,1}{0.3cm}
4895
\hedge[violet, very thick]{0,2,3,1}{0.4cm}
4896
\hedge[red, very thick]{0,5,4,1}{0.5cm}
4897
%\end{pgfonlayer}
4898
\end{tikzpicture}
4899
\qquad
4900
\begin{tikzpicture}[yscale=-1, every node/.style={draw, very thick, circle, inner sep=1pt}, edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]
4901
]
4902
\node(0) {0}
4903
child{node(1){1}}
4904
child{node(2){2}
4905
child{node(3){3}
4906
child{node(4){4}}
4907
}}
4908
child{node(5){5}};
4909
% \begin{pgfonlayer}{bg} % select the background layer
4910
\hedge[blue, very thick]{2,3}{0.3cm}
4911
\hedge[blue, very thick]{0,5}{0.3cm}
4912
\hedge[violet, very thick]{0,5,3}{0.4cm}
4913
\hedge[red, very thick]{0,5,4,1}{0.5cm}
4914
%\end{pgfonlayer}
4915
\end{tikzpicture}
4916
\caption{Two nestings of a tree with $5$ internal edges. These nestings, \cref{def:nesting}, are also $2$-colored, \cref{def:2-Colored Nesting}.}
4917
\label{fig:nestings}
4918
\end{figure}
4919
4920
Since the operahedra are generalized permutahedra~\cite[Coro.~2.16]{LaplanteAnfossi}, a choice of diagonal for the permutahedra induces a choice of diagonal for every operahedron~\cite[Coro.~1.31]{LaplanteAnfossi}.
4921
Every face of an operahedron is isomorphic to a product of lower-dimensional operahedra, via an isomorphism~$\Theta$ which generalizes the one from \cref{subsec:operadicProperty}, see Point (5) of~\cite[Prop.~2.3]{LaplanteAnfossi}.
4922
4923
\begin{definition}
4924
An \emph{operadic diagonal} for the operahedra is a choice of diagonal $\triangle_t$ for each Loday operahedron~$P_t$, such that $\triangle \eqdef \{\triangle_t\}$ commutes with the map $\Theta$, \ie it satisfies~\cite[Prop.~4.14]{LaplanteAnfossi}.
4925
\end{definition}
4926
4927
An operadic diagonal gives rise to topological operad structure on the set of Loday operahedra \cite[Thm 4.18]{LaplanteAnfossi}, and via the functor of cellular chains, to a universal tensor product of homotopy operads \cite[Prop. 4.27]{LaplanteAnfossi}.
4928
Here, by \defn{universal}, we mean a formula that applies uniformly to \emph{any} pair of homotopy operads.
4929
Since such an operad structure and tensor product are induced by a geometric diagonal, we shall call them \defn{geometric}.
4930
4931
\begin{theorem}
4932
\label{thm:operahedra}
4933
There are exactly
4934
\begin{enumerate}
4935
\item two geometric operadic diagonals of the Loday operahedra, the $\LA$ and $\SU$ diagonals,
4936
\item two geometric colored topological cellular operad structures on the Loday operahedra,
4937
\item two geometric universal tensor products of homotopy operads,
4938
\end{enumerate}
4939
which agree with the generalized Tamari order on fully nested trees.
4940
\end{theorem}
4941
4942
\begin{proof}
4943
Let us first examine Point (1).
4944
By \cref{thm:unique-operadic}, we know that if one of the two choices $\LAD$ or $\SUD$ is made on an operahedron $P_t$, one has to make the same choice on every lower-dimensional operahedron appearing in the decomposition $P_{t_1} \times \cdots \times P_{t_k} \cong F \subset P_t$ of a face $F$ of~$P_t$.
4945
Now suppose that one makes two distinct choices for two operahedra $P_t$ and $P_{t'}$.
4946
It is easy to find a bigger tree $t''$, of which both $t$ and $t'$ are subtrees.
4947
Therefore, $P_t$ and $P_{t'}$ appear as facets of $P_{t''}$ and by the preceding remark, any choice of diagonal for $P_{t''}$ will then contradict our initial two choices.
4948
Thus, these had to be the same from the start, which concludes the proof.
4949
4950
Point (2) then follows from the fact that a choice of diagonal for the Loday realizations of the operahedra \emph{forces} a unique topological cellular colored operad structure on them, see~\cite[Thm.~4.18]{LaplanteAnfossi}.
4951
Since universal tensor products of homotopy operads are induced by a colored operad structure on the operahedra~\cite[Coro.~4.24]{LaplanteAnfossi}, we obtain Point~(3).
4952
Finally, since only vectors with strictly decreasing coordinates induce the generalized Tamari order on the skeleton of the operahedra~\cite[Prop.~3.11]{LaplanteAnfossi}, we get the last part of the statement.
4953
\end{proof}
4954
4955
This answers a question raised in~\cite[Rem.~3.14]{LaplanteAnfossi}.
4956
4957
\begin{example}
4958
The Loday associahedra correspond to the Loday operahedra associated with linear trees~\cite[Sect. 2.2]{LaplanteAnfossi}, and define a suboperad.
4959
The restriction of the two operad structures of \cref{thm:operahedra} coincide in this case, and both the $\LA$ and $\SU$ diagonals induce the \emph{magical formula} of~\cite{MarklShnider, MasudaThomasTonksVallette, SaneblidzeUmble-comparingDiagonals} defining a universal tensor product of $\Ainf$-algebras.
4960
\end{example}
4961
4962
\begin{example}
4963
The restriction of \cref{thm:operahedra} to the permutahedra associated with $2$-leveled trees gives two distinct universal tensor products of permutadic $\Ainf$-algebras, as studied in~\cite{LodayRonco-permutads,Markl}.
4964
\end{example}
4965
4966
Two other important families of operadic polytopes are the \emph{Loday associahedra} and \emph{Forcey multiplihedra}, which encode respectively $\Ainf$-algebras and $\Ainf$-morphisms~\cite[Prop.~4.9]{LaplanteAnfossiMazuir}, as well as $\Ainf$-categories and $\Ainf$-functors~\cite[Sect.~4.3]{LaplanteAnfossiMazuir}.
4967
For every linear tree $t \in \PT_n$, there is a corresponding Loday associahedron $\K_n$, whose faces are in bijection with nestings of $t$, and a Forcey multiplihedron $\J_n$ whose faces are in bijection with $2$-colored nestings of $t$.
4968
4969
\begin{definition}[{\cite[Def. 3.2]{LaplanteAnfossiMazuir}}]\label{def:2-Colored Nesting}
4970
A \defn{$2$-colored nesting} is a nesting where each nest $N$ is either blue, red, or blue and red (purple), and which satisfies that if $N$ is blue or purple (\resp red or purple), then all nests contained in $N$ are blue (\resp all nests that contain $N$ are red).
4971
See \cref{fig:nestings}.
4972
\end{definition}
4973
4974
The Loday associahedra are faces of the Forcey multiplihedra: they correspond to $2$-colored nestings where all the nests are of the same color (either blue or red).
4975
4976
Forcey realizations of the multiplihedra are not generalized permutahedra, but they are projections of the Ardila--Doker realizations, which are \cite[Prop. 1.16]{LaplanteAnfossiMazuir}.
4977
A choice of diagonal for the permutahedra thus induces a choice of diagonal for every Ardila--Doker multiplihedron, and a subset of these choices (the ones which satisfy \cite[Prop. 2.7 \& 2.8]{LaplanteAnfossiMazuir}) further induce a choice of diagonal for the Forcey multiplihedra.
4978
Every face of a Forcey multiplihedron is isomorphic to a product of a Loday associahedron and possibly many lower-dimensional Forcey multiplihedra, via an isomorphism $\Theta$ similar to the one from \cref{subsec:operadicProperty}, see Point (4) of \cite[Prop. 1.10]{LaplanteAnfossiMazuir}.
4979
4980
\begin{definition}
4981
An \defn{operadic diagonal} for the multiplihedra is a choice of diagonal $\triangle_n$ for each multiplihedron $\J_n$, such that $\triangle \eqdef \{\triangle_n\}$ commutes with the map $\Theta$.
4982
\end{definition}
4983
4984
An operadic diagonal endows the Loday associahedra with a topological operad structure \cite[Thm.~1]{MasudaThomasTonksVallette}, and the Forcey multiplihedra with a topological operadic bimodule structure over the operad of Loday associahedra \cite[Thm.~1]{LaplanteAnfossiMazuir}.
4985
Via the functor of cellular chains, it defines universal tensor products of $\Ainf$-algebras and $\Ainf$-morphisms \cite[Sec.~4.2.1]{LaplanteAnfossiMazuir}.
4986
Here again, by \emph{universal} we mean a formula that applies uniformly to any pair of $\Ainf$-algebras or $\Ainf$-morphisms.
4987
We shall call such geometrically defined operadic structures and tensor products \emph{geometric}.
4988
4989
\begin{theorem}
4990
\label{thm:multiplihedra}
4991
There are exactly
4992
\begin{enumerate}
4993
\item two geometric operadic diagonals of the Forcey multiplihedra, the $\LA$ and $\SU$ diagonals,
4994
\item two geometric topological cellular operadic bimodule structures (over the Loday associahedra) on the Forcey multiplihedra,
4995
\item two compatible geometric universal tensor products of $\Ainf$-algebras and $\Ainf$-morphisms,
4996
\end{enumerate}
4997
which agree with the Tamari-type order on atomic $2$-colored nested linear trees.
4998
\end{theorem}
4999
5000
\begin{proof}
5001
Let us first examine Point (1).
5002
Consider the vectors $\b{v}_\LA \eqdef (1,2^{-1},2^{-2},\ldots,2^{-n+1})$ and $\b{v}_\SU \eqdef (2^n-1,2^{n}-2,2^n-2^2,\ldots,2^n-2^{n-1})$ in $\R^n$.
5003
As previously observed, they induce the $\LA$ and $\SU$ diagonals on the permutahedra (\cref{def:LA-and-SU}).
5004
One checks directly that both vectors satisfy \cite[Prop. 2.7 \& 2.8]{LaplanteAnfossiMazuir}, and thus define diagonals of the Forcey multiplihedron~$\J_n$ which agree with the Tamari-type order \cite[Prop. 2.10]{LaplanteAnfossiMazuir}.
5005
Moreover, these diagonals commute with the map $\Theta$ for the Forcey multiplihedra \cite[Prop. 2.14]{LaplanteAnfossiMazuir}; this is because after deleting the last coordinate of $\b{v}_\LA$ or $\b{v}_\SU$, and then applying $\Theta^{-1}$, we still have vectors which induce the $\LA$ or $\SU$ diagonal, respectively.
5006
5007
By \cref{thm:unique-operadic}, we know that if one of the two choices $\LAD$ or $\SUD$ is made on a multiplihedron $\J_n$, one has to make the same choice on every lower-dimensional multiplihedra and associahedra appearing in the product decomposition of any face of~$\J_n$.
5008
Now suppose that one makes two distinct choices for two multiplihedra $\J_n$ and $\J_{n'}$.
5009
It is easy to find a bigger multiplihedron $\J_{n''}$, for which $\J_n$ and $\J_{n'}$ appear in the product decomposition of a face of $\J_{n''}$ and by the preceding remark, any choice of diagonal for $\J_{n''}$ will then contradict our initial two choices.
5010
Thus, these had to be the same from the start, which conclude the proof of Point (1).
5011
5012
Point (2) then follows from the fact that a choice of diagonal for the Loday associahedra and the Forcey multiplihedra \emph{forces} a unique topological cellular colored operad and operadic bimodule structure on them, see~\cite[Thm.~1]{MasudaThomasTonksVallette} and \cite[Thm.~1]{LaplanteAnfossiMazuir}.
5013
Since a universal tensor products of $\Ainf$-algebras, and a compatible universal tensor products of $\Ainf$-morphisms are induced by an operad and operadic bimodule structures on the associahedra and multiplihedra respectively~\cite[Sec.~4.2.3]{LaplanteAnfossiMazuir}, we obtain Point~(3).
5014
Finally, since only vectors with strictly decreasing coordinates induce the Tamari-type order on the skeleton of the Loday multiplihedra~\cite[Prop. 2.10]{LaplanteAnfossiMazuir}, we get the last part of the statement.
5015
\end{proof}
5016
5017
This answers a question raised in~\cite[Rem.~3.9]{LaplanteAnfossiMazuir}.
5018
5019
\begin{remark}
5020
Note that in the case of the Loday associahedra, there is only one geometric operadic diagonal which induces the Tamari order atomic $2$-colored nested planar trees (equivalently, binary trees, see \cite[Fig.~6]{LaplanteAnfossiMazuir}).
5021
Therefore, there is only one geometric topological operad structure, and only one geometric universal tensor product.
5022
This is because any vector with strictly decreasing coordinates lives in the same chamber of the fundamental hyperplane arrangement of the Loday associahedra (see \cite[Ex.~1.21]{LaplanteAnfossi}).
5023
\end{remark}
5024
5025
\begin{remark}
5026
Considering all $2$-colored nested trees instead of only linear trees, one should obtain similar results for tensor products of $\infty$-morphisms of homotopy operads.
5027
\end{remark}
5028
5029
We shall see now that the two operad (\resp operadic bimodule) structures on the operahedra (\resp multiplihedra) are related to one another in the strongest possible sense: they are isomorphic as topological cellular colored operads (\resp topological operadic bimodule structure over the associahedra).
5030
5031
%%%%%%%%%%%%%%%
5032
5033
\subsection{Relating operadic structures}
5034
\label{subsec:iso-top-operads}
5035
5036
Recall that the topological cellular operad structure on the operahedra~\cite[Def.~4.17]{LaplanteAnfossi} is given by a family of partial composition maps
5037
\[
5038
\vcenter{\hbox{
5039
\begin{tikzcd}[column sep=1cm]
5040
\circ_i^{\LA}\ : \ P_{t'}\times P_{t''}
5041
\arrow[r, "\tr\times \id"]
5042
& P_{(t',\omega)}\times P_{t''}
5043
\arrow[r,hookrightarrow, "\Theta"]
5044
&
5045
P_t .
5046
\end{tikzcd}
5047
}} \]
5048
Here, the map $\tr$ is the \emph{unique} topological cellular map which commutes with the diagonal~$\LAD$, see~\cite[Prop.~7]{MasudaThomasTonksVallette}.
5049
This partial composition $\circ_i^\LA$ is an isomorphism in the category $\PolySub$~\cite[Def.~4.13]{LaplanteAnfossi} between the product $P_{t'}\times P_{t''}$ and the facet $t' \circ_i t''$ of~$P_t$.
5050
At the level of trees, the composition operation $\circ_i$ is given by \emph{substitution} \cite[Fig.~14]{LaplanteAnfossi}.
5051
Using the $\SU$ diagonal $\SUD$, one can define similarly a topological operad structure via the same formula, but with a different transition map $\tr$, which commutes with $\SUD$.
5052
5053
Recall that a face $F$ of $P_t$ is represented by a nested tree $(t,\mathcal{N})$, which can be written uniquely as a sequence of substitution of trivially nested trees
5054
$(t,\mathcal{N})=((\cdots((t_1\circ_{i_1} t_2) \circ_{i_2} t_3) \cdots )\circ_{i_k} t_{k+1})$.
5055
Here we use the increasing order on nestings~\cite[Def. 4.5]{LaplanteAnfossi}, and observe that any choice of sequence of $\circ_i$ operations yield the same nested tree, since these form an operad~\cite[Def.~4.7]{LaplanteAnfossi}.
5056
At the geometric level, we have an isomorphism
5057
\[((\cdots((\circ_{i_1}^\LA) \circ_{i_2}^\LA) \cdots) \circ_{i_k}^\LA): P_{t_1} \times P_{t_2} \times \cdots \times P_{t_{k+1}} \overset{\cong}{\longrightarrow} F \subset P_t \]
5058
between a uniquely determined product of lower dimensional operahedra, and the face $F=(t,\mathcal{N})$ of~$P_t$.
5059
Note that any choice of sequence of $\circ_i^\LA$ operations yield the same isomorphism, since they form an operad~\cite[Thm.~4.18]{LaplanteAnfossi}.
5060
The same holds when taking the~$\circ_i^\SU$ operations instead of the $\circ_i^\LA$.
5061
5062
\begin{construction}
5063
\label{const:top-iso}
5064
For any operahedron $P_t$, we define a map $\Psi_t : P_t \to P_t$
5065
\begin{itemize}
5066
\item on the interior of the top face by the identity $\id : \mathring P_t \to \mathring P_t$, and
5067
\item on the interior of the face $F=((\cdots((t_1 \circ_{i_1} t_2) \circ_{i_2} t_3) \cdots )\circ_{i_k} t_{k+1})$ of~$P_t$ by the composition of the two isomorphisms
5068
\[
5069
((\cdots ((\circ_{i_1}^\SU) \circ_{i_2}^\SU) \cdots) \circ_{i_k}^\SU) \ ((\cdots((\circ_{i_1}^\LA) \circ_{i_2}^\LA) \cdots) \circ_{i_k}^\LA)^{-1}: F \to F . \]
5070
\end{itemize}
5071
\end{construction}
5072
5073
\begin{theorem}
5074
\label{thm:top-iso}
5075
The map $\Psi \eqdef \{\Psi_t\}$ is an isomorphism of topological cellular symmetric colored operad between the $\LA$ and $\SU$ operad structures on the operahedra, in the category $\PolySub$.
5076
\end{theorem}
5077
5078
\begin{proof}
5079
By definition, we have that $\Psi$ is an isomorphism in the category $\PolySub$.
5080
It remains to show that it preserves the operad structures, \ie that the following diagram commutes
5081
\[
5082
\vcenter{\hbox{
5083
\begin{tikzcd}[column sep=2.2cm, row sep=1.3cm]
5084
P_{t'}\times P_{t''}
5085
\arrow[r, "\circ_i^\LA"]
5086
\arrow[d, "\Psi_{t'}\times\Psi_{t''}"']
5087
& P_{t} \arrow[d, "\Psi_t"] \\
5088
P_{t'}\times P_{t''}
5089
\arrow[r, "\circ_i^\SU"']
5090
& P_{t}
5091
\end{tikzcd}
5092
}}\]
5093
For two interior points $(x,y) \in \mathring P_{t'}\times \mathring P_{t''}$, the diagram clearly commutes by definition, since $\Psi_{t'}$ and~$\Psi_{t''}$ are the identity in that case.
5094
If $x$ is in a face $F=((\cdots((t_1 \circ_{i_1} t_2) \circ_{i_2} t_3) \cdots )\circ_{i_k} t_{k+1})$ of the boundary of~$P_{t'}$, then the lower composite is equal to $\circ_i^\SU (\circ_{i_1}^\SU \circ_{i_2}^\SU \cdots \circ_{i_k}^\SU \times \id)(\circ_{i_1}^\LA \circ_{i_2}^\LA \cdots \circ_{i_k}^\LA \times \id)^{-1}$, and so is the upper composite since $\Psi_t$ starts with the inverse $(\circ_i^\LA)^{-1}$ and the decomposition of~$F$ into $P_{t_1} \times \cdots \times P_{t_{k+1}} \times P_{t''}$ is unique.
5095
The case when $y$ is in the boundary of $P_{t''}$ is similar.
5096
Finally, the compatibility of $\Psi$ with units and the symmetric group actions are straightforward to check, see~\cite[Def.~4.17 \& Thm.~4.18]{LaplanteAnfossi}.
5097
\end{proof}
5098
5099
\begin{remark}
5100
\cref{const:top-iso} and \cref{thm:top-iso} do not depend on a specific choice of operadic diagonal.
5101
In this case, however, we do not lose any generality by using specifically the $\LA$ and $\SU$ operad structures.
5102
\end{remark}
5103
5104
\begin{example}
5105
\label{ex:iso-not-Hopf}
5106
Note that $\Psi$ is \emph{not} a morphism of ``Hopf" operads, \ie it does not commute with the respective diagonals $\LAD$ and $\SUD$.
5107
Consider the two square faces $F \eqdef 12|34$ and $G \eqdef 24|13$ of the $3$-dimensional permutahedron $\Perm[4]$, and choose a point $\b z \in (\mathring F + \mathring G)/2$.
5108
Then, $\LAD(z)$ and~$\SUD(z)$ are two different pair of points on the $1$-skeleton of $\Perm[4]$.
5109
Since $\circ_i^\LA$ and $\circ_i^\SU$ are the identity both on the interior of $\Perm[4]$ (by \cref{const:top-iso}) and on the $1$-skeleton of $\Perm[4]$ (see the proof of~\cite[Prop. 7]{MasudaThomasTonksVallette}), we directly obtain that
5110
\[
5111
{\LAD(z)=(\Psi \times \Psi)\LAD(z) \neq \SUD \Psi(z)=\SUD(z)}.
5112
\]
5113
\end{example}
5114
5115
Recall that the topological cellular operadic bimodule structure on the Forcey multiplihedra is given by a family of action-composition maps \cite[Def. 2.13]{LaplanteAnfossiMazuir}
5116
\[
5117
\vcenter{\hbox{
5118
\begin{tikzcd}[column sep = 16pt]
5119
\circ_{p+1}^\LA\ : \ \J_{p+1+r}\times \K_q
5120
\arrow[rr, "\tr\times \id"]
5121
& &
5122
\J_{(1,\ldots,q,\ldots,1)}\times \K_q
5123
\arrow[rr,hookrightarrow, "\Theta_{p,q,r}"]
5124
& &
5125
\J_{n}\ \ \text{and}
5126
\end{tikzcd}
5127
}}
5128
\]
5129
\[
5130
\vcenter{\hbox{
5131
\begin{tikzcd}[column sep = 16pt]
5132
\gamma_{i_1,\ldots,i_k}^\LA \ : \ \K_{k}\times \J_{i_1} \times \cdots \times \J_{i_k}
5133
\arrow[rr, "\tr\times \id"]
5134
& &
5135
\K_{(i_1,\ldots,i_k)} \times \J_{i_1} \times \cdots \times \J_{i_k}
5136
\arrow[rr,hookrightarrow, "\Theta^{i_1, \ldots , i_k}"]
5137
& &
5138
\J_{i_1+\cdots + i_k}\, .
5139
\end{tikzcd}
5140
}}
5141
\]
5142
Here, the map $\tr$ is the \emph{unique} topological cellular map which commutes with the diagonal $\LAD$, see \cite[Prop. 7]{MasudaThomasTonksVallette}.
5143
These action-composition maps $\circ_{p+1}^\LA$ and $\gamma_{i_1,\ldots,i_k}^\LA$ are isomorphisms in the category $\PolySub$ \cite[Sec.~2.1]{LaplanteAnfossiMazuir} between the products $\J_{p+1+r}\times \K_q$ and $\K_{k}\times \J_{i_1} \times \cdots \times \J_{i_k}$, and corresponding facets of $\J_n$ and $\J_{i_1 + \cdots + i_k}$, respectively.
5144
Using the $\SU$ diagonal $\SUD$, one defines similarly a topological operadic bimodule structure via the same formula, but with a different transition map $\tr$, which commutes with $\SUD$.
5145
5146
There is a bijection between $2$-colored planar trees and $2$-colored nested linear trees \cite[Lem.~3.4 \& Fig.~6]{LaplanteAnfossiMazuir}, which translate grafting of planar trees into substitution at a vertex of nested linear trees.
5147
The indices of the $\circ_{p+1}$ and $\gamma_{i_1,\ldots,i_k}$ operations above refer to grafting.
5148
Equivalently, a face of $\J_n$ is represented by a $2$-colored nested tree $(t,\mathcal{N})$, which can be written uniquely as a sequence of substitution of trivially nested $2$-colored trees $(t,\mathcal{N})=((\cdots((t_1\circ_{i_1} t_2) \circ_{i_2} t_3) \cdots )\circ_{i_k} t_{k+1})$.
5149
Here we use the left-levelwise order on nestings \cite[Def. 4.12]{LaplanteAnfossiMazuir}, and translate tree grafting operations $\circ_{p+1}$ and $\gamma_{i_1,\ldots,i_k}$ into nested tree substitution $\circ_{i_j}$.
5150
Note that any choice of substitutions yield the same $2$-colored nested tree, since these form an operadic bimodule.
5151
5152
At the geometric level, we have an isomorphism $((\cdots((\circ_{i_1}^\LA) \circ_{i_2}^\LA) \cdots) \circ_{i_k}^\LA)$ between a uniquely determined product of lower dimensional associahedra and multiplihedra, and the face $(t,\mathcal{N})$.
5153
Note that any choice of $\circ_i^\LA$ operations (\ie the $\circ_{p+1}^\LA$ and $\gamma_{i_1,\ldots,i_k}^\LA$ action-composition operations) yield the same isomorphism, since they form an operadic bimodule \cite[Thm.~1]{LaplanteAnfossiMazuir}.
5154
The same holds when taking the $\circ_i^\SU$ (\ie the $\circ_{p+1}^\SU$ and $\gamma_{i_1,\ldots,i_k}^\SU$ action-composition) operations instead.
5155
5156
\begin{construction}
5157
\label{const:top-iso-2}
5158
For any Forcey multiplihedron $\J_n$, we define a map $\Psi_n : \J_n \to \J_n$
5159
\begin{itemize}
5160
\item on the interior of the top face by the identity $\id : \mathring \J_n \to \mathring \J_n$, and
5161
\item on the interior of the face $F=((\cdots((t_1 \circ_{i_1} t_2) \circ_{i_2} t_3) \cdots )\circ_{i_k} t_{k+1})$ of~$\J_n$ by the composition of the two isomorphisms
5162
\[
5163
((\cdots ((\circ_{i_1}^\SU) \circ_{i_2}^\SU) \cdots) \circ_{i_k}^\SU) \ ((\cdots((\circ_{i_1}^\LA) \circ_{i_2}^\LA) \cdots) \circ_{i_k}^\LA)^{-1}: F \to F . \]
5164
\end{itemize}
5165
\end{construction}
5166
5167
\begin{theorem}
5168
\label{thm:top-iso-2}
5169
The map $\Psi \eqdef \{\Psi_n\}$ is an isomorphism of topological cellular operadic bimodule structure over the Loday associahedra between the $\LA$ and $\SU$ operadic bimodule structures on the Forcey multiplihedra, in the category $\PolySub$.
5170
\end{theorem}
5171
5172
\begin{proof}
5173
The proof is the same as the one of \cref{thm:top-iso}, with the multiplihedra $\circ_i^\LA$ and $\circ_i^\SU$ operations (that is, the action-composition maps $\circ_{p+1}^\LA$ and $\gamma_{i_1,\ldots,i_k}^\LA$, and $\circ_{p+1}^\SU$ and $\gamma_{i_1,\ldots,i_k}^\SU$) in place of the operahedra operations.
5174
\end{proof}
5175
5176
\begin{example}
5177
\label{ex:iso-not-Hopf-2}
5178
Note that $\Psi$ does not commute with the respective diagonals $\LAD$ and $\SUD$.
5179
Consider the two square faces $F \eqdef \purplea{\bluea{\bullet \bullet \bullet}\bullet}$ and $G \eqdef \reda{\bullet\purplea{\bullet\bullet}\bullet}$ of the $3$-dimensional Forcey multiplihedron $\J_4$, and choose a point $\b z \in (\mathring F + \mathring G)/2$.
5180
Then, $\LAD(z)$ and~$\SUD(z)$ are two different pair of points on the $1$-skeleton of $\J_4$ (see \cite[Ex.~3.7 \& Fig.~9]{LaplanteAnfossiMazuir}).
5181
Since the $\LA$ and $\SU$ action-composition maps are the identity both on the interior of $\J_4$ (by \cref{const:top-iso-2}) and on the $1$-skeleton of $\J_4$ (see the proof of~\cite[Prop. 7]{MasudaThomasTonksVallette}), we directly obtain that
5182
\[
5183
{\LAD(z)=(\Psi \times \Psi)\LAD(z) \neq \SUD \Psi(z)=\SUD(z)}.
5184
\]
5185
\end{example}
5186
5187
%%%%%%%%%%%%%%%
5188
5189
\subsection{Tensor products}
5190
\label{subsec:tensor-products}
5191
Recall that a homotopy operad $\mathcal{P}$ is a family of vector spaces $\{\mathcal{P}(n)\}_{n \geq 1}$ together with a family of operations $\{\mu_t\}$ indexed by planar trees $t$ \cite[Def. 4.11]{LaplanteAnfossi}.
5192
One can consider the category of homotopy operads with strict morphisms, that is morphisms of the underlying vector spaces which commute strictly with all the higher operations $\mu_t$, or with their $\infty$-morphisms, made of a tower of homotopies controlling the lack of commutativity of their first component with the higher operations \cite[Sec. 10.5.2]{LodayVallette}.
5193
5194
\begin{theorem}
5195
\label{thm:infinity-iso}
5196
For any pair of homotopy operads, the two universal tensor products defined by the $\LA$ and $\SU$ diagonals are not isomorphic in the category of homotopy operads and strict morphisms.
5197
However, they are isomorphic in the category of homotopy operads and their $\infty$-morphisms.
5198
\end{theorem}
5199
5200
\begin{proof}
5201
Since the two morphisms of topological operads $\LAD$ and $\SUD$ do not have the same cellular image, the tensor products that they define are not strictly isomorphic.
5202
However, they are both homotopic to the usual thin diagonal.
5203
Recall that homotopy operads are algebras over the colored operad $\mathcal{O}_\infty$, which is the minimal model of the operad $\mathcal{O}$ encoding (non-symmetric non-unital) operads \cite[Prop. 4.9]{LaplanteAnfossi}.
5204
Using the universal property of the minimal model $\mathcal{O}_\infty$, one can show that the algebraic diagonals $\LAD,\SUD : \mathcal{O}_\infty \to \mathcal{O}_\infty \otimes \mathcal{O}_\infty$ are homotopic, in the sense of \cite[Sec. 3.10]{MarklShniderStasheff}, see \cite[Prop. 3.136]{MarklShniderStasheff}.
5205
Then, by \cite[Cor.~2]{DotsenkoShadrinVallette} there is an $\infty$-isotopy, that is an $\infty$-isomorphism whose first component is the identity, between the two homotopy operad structures on the tensor product.
5206
\end{proof}
5207
5208
\begin{remark}
5209
Neither of the two diagonals $\LAD$ or $\SUD$ are cocommutative, or coassociative, as they are special cases of $\Ainf$-algebras \cite[Thm. 13]{MarklShnider}.
5210
\end{remark}
5211
5212
Note that restricting to linear trees, the two tensor products of $\Ainf$-algebras induced by the $\LA$ and $\SU$ diagonals coincide (and are thus strictly isomorphic).
5213
Restricting to $2$-leveled trees, we obtain two tensor product of permutadic $\Ainf$-algebras whose terms are in bijection.
5214
For the operahedra in general, such a bijection does not exist, as the following example demonstrates.
5215
5216
\begin{example}
5217
\label{ex:operahedra-LA-SU}
5218
The $\LA$ and $\SU$ diagonals of the operahedra associated with trees that have less than $4$ internal edges have the same number of facets.
5219
However, there are 24 planar trees with $5$ internal edges, such that the number of facets of the $\LA$ and $\SU$ diagonals are distinct, displayed in \cref{fig:trees}.
5220
To compute these numbers, we first computed the facets of the $\LA$ and $\SU$ diagonals of the permutahedra, and then used the projection from the permutahedra to the operahedra described in \cite[Prop. 3.20]{LaplanteAnfossi}.
5221
\end{example}
5222
5223
\begin{remark}
5224
The lack of symmetry in the trees in \cref{fig:trees} arises from the lack of symmetry inherent in the infix order, and in how the $\LA$ and $\SU$ diagonal treat maximal and minimal elements.
5225
A sufficient condition for the diagonals of a tree $t$ to have the same number of facets is to satisfy, $N$ is a nesting of $t$ if and only if $rN$ is a nesting of $t$.
5226
For a tree satisfying this condition, relabelling its edges via the function $r : [n]\to [n]$ defined by $r(i)\eqdef n-i+1$ exchanges the number of facets between the $\LA$ and $\SU$ diagonals.
5227
\end{remark}
5228
5229
We have an analogous result for universal tensor products of $\Ainf$-morphisms.
5230
Let $\Ainf^2$ denote the $2$-colored operad whose algebras are pairs of $\Ainf$-algebras together with an $\Ainf$-morphism between them \cite[Sec.~4.4.1]{LaplanteAnfossiMazuir}.
5231
The datum of a diagonal of the operad $\Ainf$ encoding $\Ainf$-algebras and a diagonal of the operadic bimodule $\mathrm{M}_\infty$ encoding $\Ainf$-morphisms is equivalent to the datum of a morphism of $2$-colored operads $\Ainf^2 \to \Ainf^2 \otimes \Ainf^2$.
5232
5233
\begin{theorem}
5234
\label{thm:infinity-iso-2}
5235
For any pair of $\Ainf$-morphisms, the two universal tensor products defined by the $\LA$ and $\SU$ diagonals are not isomorphic in the category of $\Ainf^2$-algebras and strict morphisms.
5236
However, they are isomorphic in the category of $\Ainf^2$-algebras and their $\infty$-morphisms.
5237
\end{theorem}
5238
5239
\begin{proof}
5240
Since the two morphisms of topological operadic bimodules on the multiplihedra $\LAD$ and $\SUD$ do not have the same cellular image, the tensor products that they define are not strictly isomorphic.
5241
However, they are both homotopic to the usual thin diagonal.
5242
Recall that the operad $\Ainf^2$ is the minimal model of the operad $\mathrm{As}^2$, whose algebras are pairs of associative algebras together with a morphism between them \cite[Prop.~4.9]{LaplanteAnfossi}.
5243
Using the universal property of the minimal model $\Ainf^2$, one can show that the algebraic diagonals $\LAD,\SUD : \Ainf^2 \to \Ainf^2 \otimes \Ainf^2$ are homotopic, in the sense of \cite[Sec.~3.10]{MarklShniderStasheff}.
5244
Then, by \cite[Cor.~2]{DotsenkoShadrinVallette} there is an $\infty$-isotopy, that is an $\infty$-isomorphism whose first component is the identity, between the two tensor products of $\Ainf$-morphisms.
5245
\end{proof}
5246
5247
\begin{remark}
5248
As studied in \cite[Sec.~4.4]{LaplanteAnfossiMazuir}, the above tensor products of $\Ainf$-morphisms are not coassociative, nor cocommutative.
5249
Moreover, there \emph{does not exist} a universal tensor product of $\Ainf$-morphisms which is compatible with composition \cite[Prop.~4.23]{LaplanteAnfossiMazuir}.
5250
\end{remark}
5251
5252
\begin{example}
5253
\label{ex:multiplihedra-LA-SU}
5254
The $\LA$ and $\SU$ diagonals of the multiplihedra associated with trees that have less than $4$ edges have the same number of facets.
5255
However, for linear trees with $5$ and $6$ internal edges, the number of facets of the $\LA$ and $\SU$ diagonals differ, as displayed in \cref{table:multiplihedra}.
5256
To compute these numbers, we first computed the facets of the $\LA$ and $\SU$ diagonals of the permutahedra, and then used the projection from the permutahedra to the multiplihedra described in the proof of \cite[Thm.~3.3.6]{Doker}.
5257
\end{example}
5258
5259
\begin{figure}[h]
5260
\begin{tabular}{cccccc}
5261
\imagebot{\begin{tikzpicture}[yscale=-1,every tree node/.style={draw, very thick, circle, inner sep=0.08cm}, level distance=0.7cm,sibling distance=0.3cm, , edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]]
5262
\Tree
5263
[.\node{};
5264
[.\node{};
5265
[.\node{};]
5266
]
5267
[.\node{};]
5268
[.\node{};]
5269
[.\node{};]
5270
]
5271
\node(1) at (0,0.5) {$(266,263)$};
5272
\end{tikzpicture}}
5273
&
5274
\imagebot{\begin{tikzpicture}[yscale=-1,every tree node/.style={draw, very thick, circle, inner sep=0.08cm}, level distance=0.7cm,sibling distance=0.3cm, , edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]]
5275
\Tree
5276
[.\node{};
5277
[.\node{};]
5278
[.\node{};
5279
[.\node{};]
5280
]
5281
[.\node{};]
5282
[.\node{};]
5283
]
5284
\node(1) at (0,0.5) {$(256,254)$};
5285
\end{tikzpicture}}
5286
&
5287
\imagebot{\begin{tikzpicture}[yscale=-1,every tree node/.style={draw, very thick, circle, inner sep=0.08cm}, level distance=0.7cm,sibling distance=0.3cm, , edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]]
5288
\Tree
5289
[.\node{};
5290
[.\node{};]
5291
[.\node{};]
5292
[.\node{};
5293
[.\node{};]
5294
]
5295
[.\node{};]
5296
]
5297
\node(1) at (0,0.5) {$(255,254)$};
5298
\end{tikzpicture}}
5299
&
5300
\imagebot{\begin{tikzpicture}[yscale=-1,every tree node/.style={draw, very thick, circle, inner sep=0.08cm}, level distance=0.7cm,sibling distance=0.3cm, , edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]]
5301
\Tree
5302
[.\node{};
5303
[.\node{};]
5304
[.\node{};]
5305
[.\node{};]
5306
[.\node{};
5307
[.\node{};]
5308
]
5309
]
5310
\node(1) at (0,0.5) {$(263,266)$};
5311
\end{tikzpicture}}
5312
&
5313
\imagebot{\begin{tikzpicture}[yscale=-1,every tree node/.style={draw, very thick, circle, inner sep=0.08cm}, level distance=0.7cm,sibling distance=0.3cm, , edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]]
5314
\Tree
5315
[.\node{};
5316
[.\node{};
5317
[.\node{};]
5318
[.\node{};]
5319
]
5320
[.\node{};]
5321
[.\node{};]
5322
]
5323
\node(1) at (0,0.5) {$(214,216)$};
5324
\end{tikzpicture}}
5325
&
5326
\imagebot{\begin{tikzpicture}[yscale=-1,every tree node/.style={draw, very thick, circle, inner sep=0.08cm}, level distance=0.7cm,sibling distance=0.3cm, , edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]]
5327
\Tree
5328
[.\node{};
5329
[.\node{};
5330
[.\node{};]
5331
]
5332
[.\node{};
5333
[.\node{};]
5334
]
5335
[.\node{};]
5336
]
5337
\node(1) at (0,0.5) {$(162,161)$};
5338
\end{tikzpicture}}
5339
\\
5340
\imagebot{\begin{tikzpicture}[yscale=-1,every tree node/.style={draw, very thick, circle, inner sep=0.08cm}, level distance=0.7cm,sibling distance=0.3cm, , edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]]
5341
\Tree
5342
[.\node{};
5343
[.\node{};]
5344
[.\node{};
5345
[.\node{};]
5346
[.\node{};]
5347
]
5348
[.\node{};]
5349
]
5350
\node(1) at (0,0.5) {$(212,213)$};
5351
\end{tikzpicture}}
5352
&
5353
\imagebot{\begin{tikzpicture}[yscale=-1,every tree node/.style={draw, very thick, circle, inner sep=0.08cm}, level distance=0.7cm,sibling distance=0.3cm, , edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]]
5354
\Tree
5355
[.\node{};
5356
[.\node{};]
5357
[.\node{};
5358
[.\node{};
5359
[.\node{};]
5360
]
5361
]
5362
[.\node{};]
5363
]
5364
\node(1) at (0,0.5) {$(129,127)$};
5365
\end{tikzpicture}}
5366
&
5367
\imagebot{\begin{tikzpicture}[yscale=-1,every tree node/.style={draw, very thick, circle, inner sep=0.08cm}, level distance=0.7cm,sibling distance=0.3cm, , edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]]
5368
\Tree
5369
[.\node{};
5370
[.\node{};]
5371
[.\node{};
5372
[.\node{};]
5373
]
5374
[.\node{};
5375
[.\node{};]
5376
]
5377
]
5378
\node(1) at (0,0.5) {$(160,161)$};
5379
\end{tikzpicture}}
5380
&
5381
\imagebot{\begin{tikzpicture}[yscale=-1,every tree node/.style={draw, very thick, circle, inner sep=0.08cm}, level distance=0.7cm,sibling distance=0.3cm, , edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]]
5382
\Tree
5383
[.\node{};
5384
[.\node{};
5385
[.\node{};
5386
[.\node{};]
5387
]
5388
[.\node{};]
5389
]
5390
[.\node{};]
5391
]
5392
\node(1) at (0,0.5) {$(142,141)$};
5393
\end{tikzpicture}}
5394
&
5395
\imagebot{\begin{tikzpicture}[yscale=-1,every tree node/.style={draw, very thick, circle, inner sep=0.08cm}, level distance=0.7cm,sibling distance=0.3cm, , edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]]
5396
\Tree
5397
[.\node{};
5398
[.\node{};
5399
[.\node{};]
5400
[.\node{};]
5401
]
5402
[.\node{};
5403
[.\node{};]
5404
]
5405
]
5406
\node(1) at (0,0.5) {$(141,144)$};
5407
\end{tikzpicture}}
5408
&
5409
\imagebot{\begin{tikzpicture}[yscale=-1,every tree node/.style={draw, very thick, circle, inner sep=0.08cm}, level distance=0.7cm,sibling distance=0.3cm, , edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]]
5410
\Tree
5411
[.\node{};
5412
[.\node{};
5413
[.\node{};
5414
[.\node{};]
5415
]
5416
]
5417
[.\node{};
5418
[.\node{};]
5419
]
5420
]
5421
\node(1) at (0,0.5) {$(91,92)$};
5422
\end{tikzpicture}}
5423
\\
5424
\imagebot{\begin{tikzpicture}[yscale=-1,every tree node/.style={draw, very thick, circle, inner sep=0.08cm}, level distance=0.7cm,sibling distance=0.3cm, , edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]]
5425
\Tree
5426
[.\node{};
5427
[.\node{};
5428
[.\node{};]
5429
]
5430
[.\node{};
5431
[.\node{};]
5432
[.\node{};]
5433
]
5434
]
5435
\node(1) at (0,0.5) {$(155,152)$};
5436
\end{tikzpicture}}
5437
&
5438
\imagebot{\begin{tikzpicture}[yscale=-1,every tree node/.style={draw, very thick, circle, inner sep=0.08cm}, level distance=0.7cm,sibling distance=0.3cm, , edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]]
5439
\Tree
5440
[.\node{};
5441
[.\node{};
5442
[.\node{};]
5443
]
5444
[.\node{};
5445
[.\node{};
5446
[.\node{};]
5447
]
5448
]
5449
]
5450
\node(1) at (0,0.5) {$(98,97)$};
5451
\end{tikzpicture}}
5452
&
5453
\imagebot{\begin{tikzpicture}[yscale=-1,every tree node/.style={draw, very thick, circle, inner sep=0.08cm}, level distance=0.7cm,sibling distance=0.3cm, , edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]]
5454
\Tree
5455
[.\node{};
5456
[.\node{};]
5457
[.\node{};
5458
[.\node{};]
5459
[.\node{};]
5460
[.\node{};]
5461
]
5462
]
5463
\node(1) at (0,0.5) {$(266,263)$};
5464
\end{tikzpicture}}
5465
&
5466
\imagebot{\begin{tikzpicture}[yscale=-1,every tree node/.style={draw, very thick, circle, inner sep=0.08cm}, level distance=0.7cm,sibling distance=0.3cm, , edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]]
5467
\Tree
5468
[.\node{};
5469
[.\node{};]
5470
[.\node{};
5471
[.\node{};
5472
[.\node{};]
5473
]
5474
[.\node{};]
5475
]
5476
]
5477
\node(1) at (0,0.5) {$(157,154)$};
5478
\end{tikzpicture}}
5479
&
5480
\imagebot{\begin{tikzpicture}[yscale=-1,every tree node/.style={draw, very thick, circle, inner sep=0.08cm}, level distance=0.7cm,sibling distance=0.3cm, , edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]]
5481
\Tree
5482
[.\node{};
5483
[.\node{};
5484
[.\node{};
5485
[.\node{};]
5486
]
5487
[.\node{};]
5488
[.\node{};]
5489
]
5490
]
5491
\node(1) at (0,0.5) {$(256,255)$};
5492
\end{tikzpicture}}
5493
&
5494
\imagebot{\begin{tikzpicture}[yscale=-1,every tree node/.style={draw, very thick, circle, inner sep=0.08cm}, level distance=0.7cm,sibling distance=0.3cm, , edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]]
5495
\Tree
5496
[.\node{};
5497
[.\node{};
5498
[.\node{};]
5499
[.\node{};
5500
[.\node{};]
5501
]
5502
[.\node{};]
5503
]
5504
]
5505
\node(1) at (0,0.5) {$(255,254)$};
5506
\end{tikzpicture}}
5507
\\
5508
\imagebot{\begin{tikzpicture}[yscale=-1,every tree node/.style={draw, very thick, circle, inner sep=0.08cm}, level distance=0.7cm,sibling distance=0.3cm, , edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]]
5509
\Tree
5510
[.\node{};
5511
[.\node{};
5512
[.\node{};]
5513
[.\node{};]
5514
[.\node{};
5515
[.\node{};]
5516
]
5517
]
5518
]
5519
\node(1) at (0,0.5) {$(263,266)$};
5520
\end{tikzpicture}}
5521
&
5522
\imagebot{\begin{tikzpicture}[yscale=-1,every tree node/.style={draw, very thick, circle, inner sep=0.08cm}, level distance=0.7cm,sibling distance=0.3cm, , edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]]
5523
\Tree
5524
[.\node{};
5525
[.\node{};
5526
[.\node{};
5527
[.\node{};]
5528
[.\node{};]
5529
]
5530
[.\node{};]
5531
]
5532
]
5533
\node(1) at (0,0.5) {$(212,213)$};
5534
\end{tikzpicture}}
5535
&
5536
\imagebot{\begin{tikzpicture}[yscale=-1,every tree node/.style={draw, very thick, circle, inner sep=0.08cm}, level distance=0.7cm,sibling distance=0.3cm, , edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]]
5537
\Tree
5538
[.\node{};
5539
[.\node{};
5540
[.\node{};
5541
[.\node{};
5542
[.\node{}; ]
5543
]
5544
]
5545
[.\node{};]
5546
]
5547
]
5548
\node(1) at (0,0.5) {$(129,127)$};
5549
\end{tikzpicture}}
5550
&
5551
\imagebot{\begin{tikzpicture}[yscale=-1,every tree node/.style={draw, very thick, circle, inner sep=0.08cm}, level distance=0.7cm,sibling distance=0.3cm, , edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]]
5552
\Tree
5553
[.\node{};
5554
[.\node{};
5555
[.\node{};
5556
[.\node{};]
5557
]
5558
[.\node{};
5559
[.\node{};]
5560
]
5561
]
5562
]
5563
\node(1) at (0,0.5) {$(160,161)$};
5564
\end{tikzpicture}}
5565
&
5566
\imagebot{\begin{tikzpicture}[yscale=-1,every tree node/.style={draw, very thick, circle, inner sep=0.08cm}, level distance=0.7cm,sibling distance=0.3cm, , edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]]
5567
\Tree
5568
[.\node{};
5569
[.\node{};
5570
[.\node{};
5571
[.\node{};]
5572
[.\node{};]
5573
[.\node{};]
5574
]
5575
]
5576
]
5577
\node(1) at (0,0.5) {$(266,263)$};
5578
\end{tikzpicture}}
5579
&
5580
\imagebot{\begin{tikzpicture}[yscale=-1,every tree node/.style={draw, very thick, circle, inner sep=0.08cm}, level distance=0.7cm,sibling distance=0.3cm, , edge from parent path={[very thick, draw] (\tikzparentnode) -- (\tikzchildnode)}]]
5581
\Tree
5582
[.\node{};
5583
[.\node{};
5584
[.\node{};
5585
[.\node{};
5586
[.\node{};]
5587
]
5588
[.\node{};]
5589
]
5590
]
5591
]
5592
\node(1) at (0,0.5) {$(154,157)$};
5593
\end{tikzpicture}}
5594
\end{tabular}
5595
\caption{The $24$ planar trees $t$ with $5$ internal edges for which the number of facets in the $\LA$ diagonal (left) and the $\SU$ diagonal (right) differ.}
5596
\label{fig:trees}
5597
\end{figure}
5598
5599
\begin{table}[h]
5600
\begin{center}
5601
\begin{tabular}{c|c|c|c|c|c}
5602
Internal edges & $\LA$ diagonal & $\LA$ only & Shared & $\SU$ only & $\SU$ diagonal \\
5603
\hline
5604
$n=1$ & 2 & 0 & 2 & 0 & 2 \\
5605
$n=2$ & 8 & 0 & 8 & 0 & 8 \\
5606
$n=3$ & 42 & 5 & 37 & 5 & 42 \\
5607
$n=4$ & 254 & 72 & 182 & 72 & 254 \\
5608
$n=5$ & 1678 & 759 & 919 & 757 & 1676 \\
5609
$n=6$ & 11790 & 7076 & 4714 & 7024 & 11738
5610
\end{tabular}
5611
\end{center}
5612
\caption{Number of facets in the $\LA$ and $\SU$ diagonals of the multiplihedra, indexed by linear trees with $n$ internal edges.}
5613
\label{table:multiplihedra}
5614
\end{table}
5615
5616
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5617
5618
\clearpage
5619
\bibliographystyle{alpha}
5620
\bibliography{diagonalsPermutahedra}
5621
\label{sec:biblio}
5622
5623
\end{document}
5624
5625