\documentclass[letterpaper,11pt,oneside,reqno]{amsart} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %bibliography \usepackage[sorting=nyt,style=alphabetic,backend=bibtex,doi=false,maxbibnames=9,maxcitenames=4]{biblatex} \makeatletter \def\blx@maxline{77} \makeatother \addbibresource{~/Dropbox/BiBTeX/bib.bib} \sloppy %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %main packages \usepackage{amsmath,amssymb,amsthm,amsfonts} \usepackage[colorlinks=true,linkcolor=blue,citecolor=red]{hyperref} \usepackage{graphicx,color,colortbl} \usepackage{upgreek} \usepackage[mathscr]{euscript} %equations \allowdisplaybreaks \numberwithin{equation}{section} %tikz \usepackage{tikz} \usetikzlibrary{ shapes, arrows, positioning, decorations.markings, decorations.pathmorphing } %conveniences \usepackage{array} \usepackage{adjustbox} \usepackage{cleveref} \usepackage{enumerate} %paper geometry \usepackage[DIV=12]{typearea} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %draft-specific %\newcommand{\note}[1]{ {\color{blue}\textsf{(#1)}}} \synctex=1 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %this paper specific %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \newtheorem{proposition}{Proposition}[section] \newtheorem{lemma}[proposition]{Lemma} \newtheorem{corollary}[proposition]{Corollary} \newtheorem{theorem}[proposition]{Theorem} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \theoremstyle{definition} \newtheorem{definition}[proposition]{Definition} \newtheorem{remark}[proposition]{Remark} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % include pictures % [scale, bottom #, top #, baseline] % \Voo \Voi \Vio \Vii \input{single_vertex_text.tex} % [scale, line width, baseline] % \ooYBoo \iiYBii \ioYBio \ioYBoi \oiYBoi \oiYBio \input{YB_vertices.tex} % [scale, baseline] % \SVoooo etc (6pcs) \input{six_vertex.tex} % [scale, bottom #, middle #, top #, baseline] % \ooWoo etc (16 pcs) for forward and \ooBoo (16 pcs) for backward \input{two_vertices_text.tex} % [scale, baseline] % \SVoooo etc (6pcs) \input{six_vertex_two.tex} % [scale] % \vertexlo etc (4pcs) \input{single_vertex_full.tex} % \colz, \coli, \colii, \colx, \colxx \input{table_colors.tex} %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \begin{document} \title[Yang-Baxter field for spin Hall-Littlewood symmetric functions]{Yang-Baxter field for spin Hall-Littlewood\\symmetric functions} \author[A. Bufetov]{Alexey Bufetov}\address{A. Bufetov, Massachusetts Institute of Technology, Department of Mathematics, 77 Massachusetts Avenue, Cambridge, MA 02139, USA}\email{alexey.bufetov@gmail.com} \author[L. Petrov]{Leonid Petrov}\address{L. Petrov, University of Virginia, Department of Mathematics, 141 Cabell Drive, Kerchof Hall, P.O. Box 400137, Charlottesville, VA 22904, USA, and Institute for Information Transmission Problems, Bolshoy Karetny per. 19, Moscow, 127994, Russia}\email{lenia.petrov@gmail.com} \date{} \begin{abstract} Employing bijectivisation of summation identities, we introduce local stochastic moves based on the Yang-Baxter equation for $U_q(\widehat{\mathfrak{sl}_2})$. Combining these moves leads to a new object which we call the spin Hall-Littlewood Yang-Baxter field --- a probability distribution on two-dimensional arrays of particle configurations on the discrete line. We identify joint distributions along down-right paths in the Yang-Baxter field with spin Hall-Littlewood processes, a generalization of Schur processes. We consider various degenerations of the Yang-Baxter field leading to new dynamic versions of the stochastic six vertex model and of the Asymmetric Simple Exclusion Process. \end{abstract} \maketitle \tableofcontents \section{Introduction} \subsection{Overview} The past two decades have seen a wave of progress in understanding large scale, long time asymptotics of driven nonequilibrium stochastic particle systems in the one space and one time dimension belonging to the Kardar-Parisi-Zhang (KPZ) universality class (about the KPZ class see, e.g., \cite{CorwinKPZ}, \cite{Corwin2016Notices}, \cite{halpin2015kpzCocktail}). Much of this progress has been achieved by discovering exact distributional formulas in these particle systems, and leveraging these formulas towards asymptotic analysis. Stochastic particle systems possessing such exact formulas are known under the name \emph{integrable}. Since the early days (e.g., \cite{johansson2000shape}), success in discovering integrability (at least for special initial data) has often been triggered by applications of techniques coming from the algebra of symmetric functions \cite[Ch. I]{Macdonald1995}. Among the most notable frameworks for these applications are Schur processes \cite{okounkov2001infinite}, \cite{okounkov2003correlation}, \cite{Borodin2010Schur}, \cite{Betea_etal2014} and Macdonald processes \cite{BorodinCorwin2011Macdonald}, \cite{BCGS2013}. The success of this approach naturally leads to a more extensive study of structural properties of various families of symmetric functions and their relations to probabilistic systems. In this work we investigate stochastic systems related to \emph{spin Hall-Littlewood symmetric rational functions} introduced in \cite{Borodin2014vertex}. These functions are naturally at the interplay of the theory of symmetric functions and the Yang-Baxter equation (see, e.g., \cite{tsilevich2006quantum}, \cite{betea2016refined}, \cite{betea2015refined}, \cite{wheeler2015refined} for other related examples). The main results of the present paper are: \begin{itemize} \item We consider the general idea of bijectivisation (\Cref{sec:bijectivisation}) and apply it to the Yang-Baxter equation obtaining local stochastic moves acting on vertex model configurations (\Cref{sec:main_construction}). We hope that the usefulness of this general idea will not be limited by the results of this paper. \item We introduce the spin Hall-Littlewood Yang-Baxter field (\Cref{sec:YB_field}), a two-dimensional array of random particle configurations on the discrete line. Its main properties are explicit formulas for distributions along any down-right path (\Cref{thm:YB_field_spin_HL_process}), and Markov projections turning the Yang-Baxter field into a two-dimensional scalar field or its multilayer versions (\Cref{prop:YB_Markov_projections,prop:dyn6V_is_YB_field}). \item We consider a number of degenerations of the Yang-Baxter field, including new dynamic versions of the stochastic six vertex model (\Cref{sec:dynamicS6V}) and the Asymmetric Simple Exclusion Process (\Cref{sub:degen_ASEP}). Our results about these dynamic models generalize those of the recent works \cite{BorodinBufetovWheeler2016} and \cite{BufetovMatveev2017}. \end{itemize} Let us describe our results in more detail. \subsection{Random fields of Young diagrams} \label{sub:YB_field_intro_section} One of the key properties behind probabilistic applications of Macdonald (in particular, Schur) symmetric functions is that they satisfy \emph{Cauchy summation identities} \cite[Ch. I.4 and Ch. VI.2]{Macdonald1995} (see also \Cref{sub:Cauchy_identity} below for Cauchy identities for the spin Hall-Littlewood symmetric functions). Regarding these identities as expressing probability normalizing constants (=~partition functions) allows to define and analyze \emph{Macdonald processes}. These are certain probability distributions on collections of Young diagrams\footnote{In probabilistic applications, Young diagrams are often interpreted as particle configurations on the discrete line.} whose probability weights are proportional to products of the (skew) Macdonald symmetric polynomials. A lot of recent research is devoted to the study of these processes and their degenerations, with applications to KPZ type and other asymptotics, e.g., see \cite{Oconnell2009_Toda}, \cite{COSZ2011}, \cite{OSZ2012}, \cite{BorodinCorwin2011Macdonald}, \cite{BorodinCorwinFerrariVeto2013}, \cite{BorodinGorin2013beta}, \cite{BorodinPetrov2013Lect}. It is much less articulated in the existing literature that one can consider Macdonald (Schur, etc.) \emph{fields} --- certain ways to couple many processes together leading to two-dimensional arrays of random Young diagrams. Such fields are highly non-unique, and coming up with a ``good'' way to couple processes together involves additional considerations like the presence of Markov projections (see below). Various elements of Young diagram random fields have appeared in the literature mainly as ways to match observables of $(1+1)$-dimensional stochastic interacting particle systems with observables of Macdonald or Schur processes. The latter observables then can be analyzed to the point of asymptotics thanks to the algebraic structure coming from symmetric functions. Two ways to construct such random fields were mainly employed which we briefly discuss in \Cref{sub:intro_RSK,sub:intro_BF} below. \subsection{RSK type fields} \label{sub:intro_RSK} RSK type fields were applied in probabilistic context in connection with Schur measures as early as in \cite{baik1999distribution}, \cite{johansson2000shape}, \cite{PhahoferSpohn2002} to study asymptotics of longest increasing subsequences, last passage percolation, TASEP (Totally Asymmetric Simple Exclusion Process), and PNG (polynuclear growth). These fields arise (in the Schur case) as results of applying the Robinson-Schensted-Knuth (RSK) insertion algorithm to a random input, hence the name. More precisely, a Schur RSK type field can be realized using Fomin growth diagrams (an equivalent way to interpret the RSK insertion \cite{Fomin1986}, \cite{fomin1995schensted}) with random integer inputs. The idea to apply RSK insertion to random input seems to have first appeared in \cite{Vershik1986}, and was substantially developed in \cite{Baryshnikov_GUE2001}, \cite{OConnell2003}, \cite{OConnell2003Trans}. Recently RSK type fields associated with deformations of Schur processes (see \Cref{fig:polys}) were constructed for Whittaker processes \cite{Oconnell2009_Toda}, \cite{COSZ2011}, \cite{OSZ2012}, $q$-Whittaker processes \cite{OConnellPei2012}, \cite{BorodinPetrov2013NN}, \cite{Pei2013Symmetry}, \cite{MatveevPetrov2014}, \cite{pei2016qRSK}, and Hall-Littlewood processes \cite{BufetovPetrov2014}, \cite{BorodinBufetovWheeler2016}, \cite{BufetovMatveev2017}. Constructions at the Whittaker level relied on the geometric (also sometimes called ``tropical'') lifting of the RSK correspondence \cite{Kirillov2000_Tropical}, \cite{NoumiYamada2004}, while the $q$-Whittaker and Hall-Littlewood developments required nontrivial randomizations of the original RSK insertion algorithm. Via Markov projections, this work uncovered connections of Whittaker, $q$-Whittaker, and Hall-Littlewood processes with known and new $(1+1)$-dimensional stochastic particle systems. In the Whittaker case, these are various integrable models of directed random polymers \cite{OConnellYor2002}, \cite{Seppalainen2012}. For the $q$-Whittaker processes, these are the $q$-TASEP and related systems \cite{BorodinCorwin2013discrete}, \cite{CorwinPetrov2013}, \cite{MatveevPetrov2014}. In the Hall-Littlewood case these are the ASEP \cite{macdonald1968bioASEP}, \cite{Spitzer1970} and the stochastic six vertex model \cite{GwaSpohn1992}, \cite{BCG6V}. \begin{figure}[htpb] %\begin{noindent} \centering \scalebox{1}{ \begin{tikzpicture} [scale=1, very thick] \node[fill=white,draw,rectangle] (schur) at (0,0) {Schur}; \node[fill=white,draw,rectangle] (whit) at (4,0) {Whittaker}; \node[fill=white,draw,rectangle] (HL) at (-4,2) {Hall-Littlewood ($t$)}; \node[fill=white,draw,rectangle] (sHL) at (-5.6,4.2) {spin Hall-Littlewood ($t,s$)}; \node[fill=white,draw,rectangle] (jack) at (-.6,2) {Jack}; \node[fill=white,draw,rectangle] (M) at (0,4.2) {Macdonald ($q,t$)}; %\node[fill=white,draw,rectangle] (sqW) at (5,4.2) {spin $q$-Whittaker ($q,s$)}; \node[fill=white,draw,rectangle] (qW) at (4,2) {$q$-Whittaker ($q$)}; \draw[->] (M)--(jack) node [midway, xshift=23, yshift=-10] {\parbox{.085\textwidth}{$t=q^{\beta/2}$\\\phantom{.}\hfill$\to1$}}; \draw[->] (jack)--(schur) node [midway, xshift=-18, yshift=2] {$\beta=2$}; \draw[->] (sHL)--(HL) node [midway, xshift=-20] {$s=0$}; \draw[->] (HL)--(schur) node [midway, xshift=-25] {$t=0$}; \draw[->] (M)--(HL) node [midway, yshift=10, xshift=-5] {$q=0$}; \draw[->] (M)--(qW) node [midway, xshift=25] {$t=0$}; %\draw[->] (sqW)--(qW) node [midway, xshift=20] {$s=0$}; \draw[->] (qW)--(schur) node [midway, xshift=25] {$q=0$}; \draw[->] (qW)--(whit) node [midway, xshift=20] {$q\nearrow1$}; \end{tikzpicture}} \caption{A part of the hierarchy of symmetric functions satisfying summation identities of Cauchy type. Arrows mean degenerations.} \label{fig:polys} %\end{noindent} \end{figure} \subsection{BF type fields} \label{sub:intro_BF} Another method of constructing random fields of Young diagrams is based on interpreting the skew Cauchy identity as an intertwining relation between certain Markov transition matrices, and stitching these matrices together into a multivariable Markov chain using an idea of Diaconis and Fill \cite{DiaconisFill1990}. In symmetric functions context this method was first applied (in the Schur case) in a work by Borodin and Ferrari \cite{BorFerr2008DF}, hence the name. In principle, this approach is applicable to a wider variety of models than the RSK one, and does not require intricate combinatorial constructions. This generality comes at a cost of having fewer Markovian projections than the RSK constructions, especially away from the Schur case. An exception in the literature is that the half-continuous BF type field in the setting of $q$-Whittaker processes has led to the discovery of the continuous time $q$-TASEP, a notable deformation of the TASEP with a richer algebraic structure \cite{BorodinCorwin2011Macdonald}. A unified approach to both the RSK type and the BF type fields in the half-continuous setting (details on half-continuous degenerations of random fields may be found in \Cref{sub:degen_half_continuous_DS6V,sub:hc_degen_Schur}) was suggested in \cite{BorodinPetrov2013NN}. In fully discrete setting, elements of BF type fields for Schur polynomials appeared in \cite{warrenwindridge2009some}, \cite{BorFerr2008DF}. \subsection{Yang-Baxter field} \label{sub:intro_intro_YB_fields} We present a third way of constructing random fields associated with symmetric functions and the corresponding processes. Our approach is based on the Yang-Baxter equation which is behind many families of symmetric functions including Schur, Hall-Littlewood, and spin Hall-Littlewood ones. We focus on the latter family for which Cauchy summation identities were recently established in \cite{Borodin2014vertex} with the help of the Yang-Baxter equation for the quantum $\mathfrak{sl}_2$ \cite{baxter2007exactly}. In the setting of spin Hall-Littlewood processes, random fields have not been considered in the literature yet. The Yang-Baxter field we construct in the present paper yields a new object even in the most basic Schur case (\Cref{sub:hc_degen_Schur}). The main advantages of our approach are its simplicity and clear structure of Markov projections yielding new $(1+1)$-dimensional stochastic systems (see \Cref{sub:degen_intro_section} below). In comparison, an RSK type approach would likely require very nontrivial combinatorial considerations (cf. \cite{BufetovMatveev2017} for the Hall-Littlewood case), further complicated by the fact that the spin Hall-Littlewood functions are not homogeneous polynomials while the usual Hall-Littlewood ones are (see \Cref{rmk:spinHL_not_RSK} for more details). A BF type approach, while clearly being applicable in the spin Hall-Littlewood case, might not readily produce Markov projections. Our construction of the Yang-Baxter field uses a very basic idea of bijectivisation of the Yang-Baxter equation. We briefly describe this idea next. \subsection{Bijectivisation of the Yang-Baxter equation} \label{sub:bijectivisation_intro_section} In probability theory it is well known that considering couplings of probability measures is a powerful idea. For our construction of Yang-Baxter field we apply a similar idea to summation identities which form the Yang-Baxter equation for quantum $\mathfrak{sl}_2$. We refer to it as a \emph{bijectivisation} of these combinatorial summation identities. As a byproduct of couplings thus constructed we obtain conditional distributions, and we regard them as local stochastic (Markov) moves acting on vertex model configurations. The bijectivisation of the Yang-Baxter equation we consider is also not unique, but the space of possible parameters is quite small. We use this freedom to choose a bijectivisation with the least ``noise'', in the spirit of RSK type approach, cf. \cite[Section 7.4]{BorodinPetrov2013NN}. See \Cref{sub:discussion} for details. We believe that one of important novelties of this paper is the application of this idea of coupling to combinatorial summation identities. Here we use it in only one situation, in the setting of the Yang-Baxter equation powering the spin Hall-Littlewood symmetric functions. However, it seems likely that this idea might lead to new interesting constructions and results for other forms of Yang-Baxter equation as well. \subsection{Dynamic stochastic six vertex model and dynamic ASEP} \label{sub:degen_intro_section} A certain Markov projection of our Yang-Baxter random field yields a scalar-valued random field indexed by the nonnegative integer quadrant. This scalar field can be interpreted as a random field of values of the height function in a certain generalization of the stochastic six vertex model in which the vertex probabilities additionally depend on the value of the height function. For this reason one can call this model a \emph{dynamic stochastic six vertex model} (DS6V). Its detailed description is given in \Cref{sub:dynamicS6V_subsection}. The joint distribution of the values of the height function in DS6V along \emph{down-right paths}\footnote{Also referred to as \emph{space-like paths} in the language of stochastic particle systems, cf. \cite{derrida1991dynamics}, \cite{Ferrari_Airy_Survey}, \cite{BorFerr08push}.} can be identified with that of certain observables of a spin Hall-Littlewood process (\Cref{cor:dyn6V_spin_HL_process}). In the degeneration turning the spin Hall-Littlewood symmetric functions into the Hall-Littlewood ones, the DS6V model becomes the usual stochastic six vertex model of \cite{GwaSpohn1992}, \cite{BCG6V}, and \Cref{cor:dyn6V_spin_HL_process} turns into the statement established in \cite{BufetovMatveev2017}. \medskip Along with single-layer projections leading to DS6V, one can consider multilayer projections of the full Yang-Baxter field, as was done in \cite[Sections 4.4 and 4.5]{BufetovMatveev2017} for the Hall-Littlewood RSK field. In particular, one can check that the two-layer projection of our Yang-Baxter field, in the Hall-Littlewood degeneration, coincides with the two-layer stochastic six vertex model of \cite[Section 4.4]{BufetovMatveev2017}. However, the corresponding degeneration of the full Yang-Baxter field is different from the full Hall-Littlewood RSK field. Details may be found in \Cref{sec:degenerations}. \medskip In a continuous time limit around the diagonal, the DS6V model turns into the following dynamic version of the ASEP depending on parameters $t\ge 0$, $-10$. Consider a continuous time particle system $\{y_\ell(\tau)\}_{\ell\in \mathbb{Z}_{\ge1},\;\tau\in \mathbb{R}_{\ge0}}$ on $\mathbb{Z}$ (no more than one particle at a site), started from the step initial configuration $y_\ell(0)=-\ell$. In continuous time, each particle $y_{\ell}$, $\ell\ge1$, tries to jump to the right by one at rate\footnote{That is, the waiting time till the jump is an independent exponential random variable with mean equal to $(\textnormal{rate})^{-1}$.} $\dfrac{u-st^{\ell}}{u-st^{\ell-1}}$, and to the left by one at rate $t\,\dfrac{u-st^{\ell-1}}{u-st^{\ell}}$. If the destination is occupied, the corresponding jump is blocked and $y_{\ell}$ does not move. See \Cref{fig:dyn_ASEP}. The height function in this dynamic ASEP can be identified in distribution with a certain limit of observables of spin Hall-Littlewood processes. When $s=0$, the dynamic dependence of jump rates on the height function disappears, and the system turns into the usual ASEP. See \cite{BufetovMatveev2017} for connections of ASEP to Hall-Littlewood processes. \begin{figure}[htpb] %\begin{noindent} \centering \begin{tikzpicture} [scale=1,very thick] \def\pt{.17} \def\ee{.1} \def\h{.45} \draw[->] (-.5,0) -- (8.5,0); \foreach \ii in {(0,0),(\h,0),(3*\h,0),(4*\h,0),(5*\h,0),(6*\h,0), (8*\h,0),(10*\h,0),(9*\h,0),(12*\h,0),(13*\h,0),(14*\h,0),(15*\h,0),(16*\h,0),(17*\h,0),(18*\h,0)} { \draw \ii circle(\pt); } \foreach \ii in {(2*\h,0),(7*\h,0),(11*\h,0),(15*\h,0),(8*\h,0),(16*\h,0)} { \draw[fill] \ii circle(\pt); } \node at (16*\h,-.5) {$y_1$}; \node at (15*\h,-.5) {$y_2$}; \node at (11*\h,-.5) {$y_3$}; \node at (8*\h,-.5) {$y_4$}; \node at (7*\h,-.5) {$y_5$}; \node at (2*\h,-.5) {$y_6$}; \draw[->, very thick] (2*\h,.3) to [in=180,out=90] (2.5*\h,.65) to [in=90, out=0] (3*\h,.3) node [xshift=5,yshift=20] {$\frac{u-st^6}{u-st^5}$}; \draw[->, very thick] (2*\h,.3) to [in=0,out=90] (1.5*\h,.65) to [in=90, out=180] (1*\h,.3) node [xshift=-5,yshift=20] {$t\frac{u-st^5}{u-st^6}$}; \draw[->, very thick] (8*\h,.3) to [in=0,out=90] (7.5*\h,.65) to [in=90, out=180] (7*\h,.3); \draw[ultra thick] (7.5*\h,.65)--++(.1,.2)--++(-.2,-.4)--++(.1,.2)--++(-.1,.2)--++(.2,-.4); \draw[->, very thick] (8*\h,.3) to [in=180,out=90] (8.5*\h,.65) to [in=90, out=0] (9*\h,.3) node [xshift=5,yshift=20] {$\frac{u-st^4}{u-st^3}$}; \end{tikzpicture} \caption{A new dynamic version of the ASEP.} \label{fig:dyn_ASEP} %\end{noindent} \end{figure} The connection between spin Hall-Littlewood process and DS6V and dynamic ASEP hint at the possible integrability of the latter models, which might lead to asymptotic results for them. We do not address this question in the present paper. Note also that other dynamic generalizations of the stochastic six vertex model and the ASEP were recently considered in \cite{borodin2017elliptic}, \cite{aggarwal2017dynamical}, \cite{BorodinCorwin2017dynamic} in connection with vertex models related to the Yang-Baxter equation for the elliptic quantum group $E_{\tau,\eta}(\mathfrak{sl}_2)$. These dynamic models are different from the ones introduced in the present work. \subsection{Outline} In \Cref{sec:bijectivisation} we outline the general idea of bijectivisation of summation identities. In \Cref{sec:main_construction} we describe the higher spin six vertex weights, the Yang-Baxter equation they satisfy, and its bijectivisation with minimal ``noise''. In \Cref{sec:spin_HL_functions} we recall the spin Hall-Littlewood symmetric functions and Cauchy summation identities they satisfy. This section closely follows \cite{Borodin2014vertex}. In \Cref{sec:local_transition_probabilities} we use our bijectivisation of the Yang-Baxter equation sequentially to produce a bijective proof of the skew Cauchy identity for the spin Hall-Littlewood symmetric functions. In \Cref{sec:YB_field} we define our main object, the Yang-Baxter field, and discuss its connection with spin Hall-Littlewood measure and processes. In \Cref{sec:dynamicS6V} we consider a projection of the Yang-Baxter field onto the column number zero leading to a new dynamic version of the stochastic six vertex model. We also discuss a dynamic Yang-Baxter equation for these dynamic six vertex weights. In \Cref{sec:degenerations} we consider various degenerations of the dynamic stochastic six vertex model. One of these degenerations produces a new dynamic version of the ASEP. In \Cref{app:YB_equation,app:YB_probabilities} we explicitly list all identities comprising the Yang-Baxter equation, and all the forward and backward local transition probabilities coming out of our bijectivisation of the Yang-Baxter equation. In \Cref{sub:another_Cauchy} we discuss another versions of the skew Cauchy identity satisfied by the spin Hall-Littlewood symmetric functions. In \Cref{app:inhomogeneous_construction} we briefly outline extensions of our main constructions to the case of inhomogeneous parameters spin Hall-Littlewood symmetric functions. \subsection{Acknowledgments} We appreciate helpful discussions with Alexei Borodin, Ivan Corwin, Grigori Olshanski, and Nicolai Reshetikhin. The work was started when the authors attended the 2017 IAS PCMI Summer Session on Random Matrices, and we are grateful to the organizers for their hospitality and support. LP is partially supported by the NSF grant DMS-1664617. \section{Bijectivisation of summation identities} \label{sec:bijectivisation} \subsection{General formalism} Here we explain the formal concept of bijectivisation of summation identities which will be applied to the Yang-Baxter equation in \Cref{sec:main_construction} below. Let $A$ and $B$ be two fixed finite nonempty sets, and each element $a\in A$ and $b\in B$ is assigned certain weight $w(a)$ or $w(b)$, respectively. Assume that the following summation identity holds: \begin{equation} \label{summation_identity} \sum_{a\in A}w(a)=\sum_{b\in B}w(b). \end{equation} \begin{definition} \label{def:bijectivisation} We say that the following data provides a \emph{bijectivisation} of identity \eqref{summation_identity}: \begin{itemize} \item There are \emph{forward transition weights} $p^{\mathrm{fwd}}(a,b)$ which satisfy \begin{equation*} \sum_{b\in B}p^{\mathrm{fwd}}(a,b)=1\quad \textnormal{for each $a\in A$}; \end{equation*} \item There are \emph{backward transition weights} $p^{\mathrm{bwd}}(b,a)$ which satisfy \begin{equation*} \sum_{a\in A}p^{\mathrm{bwd}}(b,a)=1\quad \textnormal{for each $b\in B$}; \end{equation*} \item The transition weights satisfy the \emph{reversibility condition} \begin{equation} \label{reversibility_condition} w(a)p^{\mathrm{fwd}}(a,b)= w(b)p^{\mathrm{bwd}}(b,a)\quad \textnormal{for each $a\in A$ and $b\in B$.} \end{equation} \end{itemize} \end{definition} The term ``bijectivisation'' is justified by the following two observations. First, if $A$ and $B$ have the same numbers of elements, $w(a)=w(b)=1$ for all $a\in A$, $b\in B$, and each $p^{\mathrm{fwd}}(a,b)$ and $p^{\mathrm{bwd}}(b,a)$ is either $0$ or $1$, then such a bijectivisation is simply a bijection between $A$ and $B$. Second, let us get back to the general situation of \Cref{def:bijectivisation} and assume that a bijectivisation $\left\{ p^{\mathrm{fwd}}(a,b), p^{\mathrm{bwd}}(b,a)\right\}$ is given. Start from the left-hand side of \eqref{summation_identity} and write \begin{equation*} \sum_{a\in A}w(a)= \sum_{a\in A}w(a) \Biggl( \sum_{b\in B}p^{\mathrm{fwd}}(a,b) \Biggr) = \sum_{b\in B}w(b) \Biggl( \sum_{a\in A}p^{\mathrm{bwd}}(b,a) \Biggr) =\sum_{b\in B}w(b). \end{equation*} Then, due to the reversibility condition \eqref{reversibility_condition}, in the middle two double sums the terms are in \emph{one-to-one correspondence}. Thus, one can say that the transition weights $\left\{ p^{\mathrm{fwd}}(a,b), p^{\mathrm{bwd}}(b,a)\right\}$ produce a \emph{refinement} (or a \emph{bijective proof}) of the initial identity~\eqref{summation_identity}. \begin{remark} \label{rmk:bij_not_unique} Clearly, if both $A$ and $B$ have more than one element, then a bijectivisation is highly non-unique. However, in a concrete situation (such as for the Yang-Baxter equation in \Cref{sec:main_construction}) a particular bijectivisation might be more natural than the others. This choice would depend on additional structure of individual terms in~\eqref{summation_identity}. \end{remark} \subsection{Stochastic bijectivisation} Now assume that the weights $w(a)$ and $w(b)$ in \eqref{summation_identity} are \emph{stochastic}, i.e., they are positive\footnote{If some weights are equal to zero then let us remove the corresponding elements from $A$ and $B$.} and sum to one: $\sum_{a\in A}w(a)=\sum_{b\in B}w(b)=1$. The latter condition can always be achieved for positive weights $w(a),w(b)$ by dividing \eqref{summation_identity} by their sum. If the transition weights in a bijectivisation $\left\{ p^{\mathrm{fwd}}(a,b), p^{\mathrm{bwd}}(b,a)\right\}$ are all nonnegative, we call such bijectivisation \emph{stochastic}. Another standard term used in Probability Theory for such an object is \emph{coupling}. A stochastic bijectivisation may be interpreted as a joint probability distribution on $A\times B$ having prescribed marginal distributions $\left\{ w(a) \right\}_{a\in A}$ and $\left\{ w(b) \right\}_{b\in B}$. The forward and backward transition weights become families of conditional distributions coming from this joint distribution on $A\times B$. The reversibility condition \eqref{reversibility_condition} simply states the compatibility between the two conditional distributions $p^{\mathrm{fwd}}$ and $p^{\mathrm{bwd}}$. One can also interpret $\{ p^{\mathrm{fwd}}(a,b)\}_{a\in A, b\in B}$ as a Markov transition matrix from $A$ to $B$, and similarly for $p^{\mathrm{bwd}}$. This explains the terms ``transition weights'' and ``reversibility condition''. If a stochastic bijectivisation has all transition weights $p^{\mathrm{fwd}}, p^{\mathrm{bwd}}$ equal to 0 or 1, we call such bijectivisation \emph{deterministic}. \subsection{Examples} \label{sub:examples} Let us discuss two examples of bijectivisation relevant to the Yang-Baxter equation considered in \Cref{sec:main_construction} below. \subsubsection{One of the sets is a singleton} \label{ssub:singleton} For the first example, assume that $B=\left\{ b \right\}$ is a singleton while $A=\{a_1,\ldots,a_n\}$ is an arbitrary finite set. The bijectivisation is unique in this case and is given by \begin{equation*} p^{\mathrm{fwd}}(a_i,b)=1,\qquad p^{\mathrm{bwd}}(b,a_i)=\frac{w(a_i)}{w(b)},\qquad i=1,\ldots,n . \end{equation*} \subsubsection{Both sets have two elements} \label{ssub:two_and_two} For the second example, consider the situation when both sets $A=\{a_1,a_2\}$, $B=\{b_1,b_2\}$ have two elements, and all the four weights $w(a_i),w(b_j)$ are nonzero. In this case there are 8 forward and backward transition weights which must solve 4 equations of the form $p^{\mathrm{fwd}}(a_1,b_1)+p^{\mathrm{fwd}}(a_1,b_2)=1$, plus 4 more reversibility equations involving the weights $w(a_i),w(b_j)$. However, since the weights satisfy \eqref{summation_identity}, the reversibility equations are not independent, and hence the rank of the system of linear equations on the transition weights is equal to 7. (Another way to see this is to use quantities from \eqref{reversibility_condition} as variables: there are 4 variables and 3 linearly independent conditions on them.) Therefore, there is a one-parameter family of bijectivisations. One readily checks that these solutions can be expressed in the following form: \begin{equation} %\begin{noindent} \label{general_2_2_solution} \begin{array}{ll} p^{\mathrm{fwd}}(a_1,b_1)=\gamma, & \quad p^{\mathrm{fwd}}(a_1,b_2)=1-\gamma, \\[12pt] p^{\mathrm{fwd}}(a_2,b_1)= 1-\dfrac{w(b_2)}{w(a_2)}+(1-\gamma) \dfrac{w(a_1)}{w(a_2)} , & \quad p^{\mathrm{fwd}}(a_2,b_2)= \dfrac{w(b_2)}{w(a_2)}-(1-\gamma)\dfrac{w(a_1)}{w(a_2)} , \\[12pt] p^{\mathrm{bwd}}(b_1,a_1)= \gamma\dfrac{w(a_1)}{w(b_1)}, &\quad p^{\mathrm{bwd}}(b_1,a_2)= 1-\gamma\dfrac{w(a_1)}{w(b_1)}, \\[12pt] p^{\mathrm{bwd}}(b_2,a_1)= (1-\gamma)\dfrac{w(a_1)}{w(b_2)}, &\quad p^{\mathrm{bwd}}(b_2,a_2) =1-(1-\gamma) \dfrac{w(a_1)}{w(b_2)}. \end{array} %\end{noindent} \end{equation} Let us also consider a particular case of the above example when $w(a_1)=w(b_1)$ (thus automatically $w(a_2)=w(b_2)$). In this case the $\gamma$-dependent general solution \eqref{general_2_2_solution} simplifies. Namely, it depends on the weights $w(\cdot)$ only through the combination $(1-\gamma)w(a_1)/w(a_2)$. Thus, the most natural bijectivisation of the summation identity \begin{equation} \label{2_2_equation_easy_case} w(a_1)+w(a_2)=w(b_1)+w(b_2),\qquad w(a_1)=w(b_1),\quad w(a_2)=w(b_2) \end{equation} corresponds to choosing $\gamma=1$, does not depend on the weights $w(\cdot)$, and is \emph{deterministic}. Namely, the term $w(a_1)$ is simply mapped to the term $w(b_1)$ equal to it, and similarly for $w(a_2)$ and $w(b_2)$. \section{Yang-Baxter equation and its bijectivisation} \label{sec:main_construction} The goal of this section is to apply bijectivisation of \Cref{sec:bijectivisation} to Yang-Baxter equation for the (horizontal spin-$\frac12$) higher spin six vertex model. This model corresponds to the quantum group $U_q(\widehat{\mathfrak{sl}_2})$. The main outcome of this section is the definition of forward and backward transition weights in \Cref{sub:YB_bijectivisation_new_label}. \subsection{Vertex weights} \label{sub:vertex_weights} Here we recall vertex weights of the higher spin six vertex model introduced in \cite{KulishReshetikhin_YB_1981}. In our formulas we adopt the parametrization used in \cite{Borodin2014vertex}. The vertex weights depend on the main ``\emph{quantization}'' parameter $t\in(0,1)$, the \emph{vertical spin} parameter $s$, and the \emph{spectral} parameter $u$, with only the latter explicitly indicated in the notation. These weights are associated to a vertex $(i_1,j_1;i_2,j_2)$ on the lattice $\mathbb{Z}^2$ which has $i_1$ and $i_2$ incoming and outgoing vertical arrows, and $j_1$ and $j_2$ incoming and outgoing horizontal arrows, respectively. We assume that our vertex model has horizontal spin-$\frac{1}{2}$ and generic higher vertical spin, which is equivalent to saying that the vertex weights are nonzero only if $j_1,j_2\in\left\{0,1\right\}$ and $i_1,i_2\in\mathbb{Z}_{\ge0}$. (See also \Cref{sub:degen_finite_spin} for a discussion of models with finite vertical spin $I$ obtained by specializing the vertical spin parameter $s$ to $t^{-I/2}$, $I\in\mathbb{Z}_{\ge1}$.) The arrows at any vertex should satisfy the \emph{preservation property} $i_1+j_1=i_2+j_2$. Depending on $j_1,j_2$, we will denote vertices by \begin{equation} \label{single_vertex_notation} (g,0;g,0)=\Voo{.6}gg{-4.5pt} \,, \quad (g,0;g-1,1)=\Voi{.6}g{g-1}{-4.5pt} \,, \quad (g,1;g,1)=\Vii{.6}gg{-4.5pt} \,, \quad (g,1;g+1,0)=\Vio{.6}g{g+1}{-4.5pt} \, \end{equation} (see also \Cref{fig:vertex_weights} for a more detailed graphical representation). Here $g\in\mathbb{Z}_{\ge0}$ is arbitrary, with the agreement that $g\ge1$ in the second vertex. The weights of these vertices are defined as \begin{equation} %\begin{noindent} \label{vertex_weights} \begin{array}{cll} & \Big[ \Voo{.6}gg{-4.5pt} \Big]_{u}:= \dfrac{1-st^gu}{1-su}, & \qquad \Big[ \Voi{.6}g{g-1}{-4.5pt} \Big]_{u}:= \dfrac{(1-s^2t^{g-1})u}{1-su}, \\[8pt] &\Big[ \Vii{.6}gg{-4.5pt} \Big]_{u} := \dfrac{u-st^g}{1-su}, & \qquad \Big[ \Vio{.6}{g}{g+1}{-4.5pt} \Big]_{u} := \dfrac{1-t^{g+1}}{1-su}. \end{array} %\end{noindent} \end{equation} Weights \eqref{vertex_weights} are very special in that they satisfy a Yang-Baxter equation which we recall in the next subsection. \begin{figure}[htbp] \begin{tabular}{c|c|c|c} \vertexoo{1} & \vertexol{1} & \vertexll{1} & \vertexlo{1} \\\hline\rule{0pt}{20pt} $\dfrac{1-s t^{g}u}{1-s u}$ & $\dfrac{(1-s ^{2}t^{g-1})u}{1-s u}$ & $\dfrac{u-s t^{g}}{1-s u}$ & $\dfrac{1-t^{g+1}}{1-s u}$ \phantom{\Bigg|} \\ \end{tabular} \caption{Possible vertices in the (horizontal spin-$\frac{1}{2}$) higher spin six vertex model, with their weights \eqref{vertex_weights}.} \label{fig:vertex_weights} \end{figure} \begin{remark} The higher spin weights \eqref{vertex_weights} of \cite{KulishReshetikhin_YB_1981} generalize the original six vertex weights \cite{pauling1935structure}, \cite{Lieb1967SixVertex}, \cite{baxter2007exactly} to the case when the vertical representation is arbitrary highest weight (corresponding to the spin parameter $s$), and the horizontal representation is still one-dimensional. Using a procedure called fusion \cite{KR1987Fusion}, one can define vertex weights corresponding to both representations being arbitrary. Explicit formulas for fused vertex weights may be found in, e.g., \cite{Mangazeev2014}, see also \cite{CorwinPetrov2015} for a probabilistic interpretation. In the present paper we only use the simpler weights \eqref{vertex_weights} and do not employ the fused ones. \end{remark} \begin{remark} We denote the quantization parameter of the higher spin six vertex model by $t$ instead of $q$ used in \cite{Borodin2014vertex}, \cite{CorwinPetrov2015}, \cite{BorodinPetrov2016inhom}. This is done to highlight properties (in particular, Cauchy summation identities) of the spin Hall-Littlewood symmetric functions which degenerate at $s=0$ to the corresponding properties of the usual Hall-Littlewood symmetric polynomials. Vertex models in the context of Hall-Littlewood polynomials and their properties were recently studied in, e.g., \cite{BorodinBufetovWheeler2016}, \cite{BufetovMatveev2017}, and we follow these papers when using the parameter $t$. Note that setting $s=t=0$ reduces the picture to the one associated with the classical Schur polynomials, see \Cref{sec:degenerations}. \end{remark} \subsection{Yang-Baxter equation} \label{sub:YB} The Yang-Baxter equation \cite{YangSystem1967}, \cite{baxter2007exactly}, \cite{KulishReshSkl1981yang} can be regarded as the origin of integrability of the stochastic higher spin six vertex model, cf. \cite{BorodinPetrov2016inhom}. It can be written in a rather compact form involving $4\times 4$ matrices containing certain combinations of vertex weights. For example, see \cite[Proposition 2.5]{Borodin2014vertex} for the statement for our particular parametrization. However, as we aim to construct a bijectivisation of the Yang-Baxter equation in the sense of \Cref{sec:bijectivisation}, we need to write the Yang-Baxter equation out in full detail, considering each of its matrix elements separately. Let us first define weights of auxiliary \emph{cross vertices}. The cross vertices' incoming and outgoing arrow directions are rotated by $45^\circ$, and along each direction there can be at most one arrow. Therefore, due to the arrow preservation there are 6 possible cross vertices. Their weights depend on two arbitrary spectral parameters $u,v$ and are defined as follows: \begin{equation} \begin{array}{clll} & \left[ \ooYBoo{.35}{3}{14.5pt} \right]_{u,v}:=1, & \qquad \left[ \oiYBio{.35}{3}{14.5pt} \right]_{u,v}:=\uprho=\dfrac{u-v}{u-tv}, & \qquad \left[ \oiYBoi{.35}{3}{14.5pt} \right]_{u,v}:=1-\uprho=\dfrac{(1-t)v}{u-tv}, \\[8pt]& \left[ \iiYBii{.35}{3}{14.5pt} \right]_{u,v}:=1, & \qquad \left[ \ioYBoi{.35}{3}{14.5pt} \right]_{u,v}:=t\uprho=\dfrac{t(u-v)}{u-tv}, & \qquad \left[ \ioYBio{.35}{3}{14.5pt} \right]_{u,v}:=1-t\uprho=\dfrac{(1-t)u}{u-tv}. \end{array} \label{cross_vertex_weights} \end{equation} Here we employed the shorthand notation $\uprho:=(u-v)/(u-tv)$. Let us now introduce notation for weights of \emph{pairs of vertices} where one vertex as in \Cref{fig:vertex_weights} is put on top of another. Because each of the two vertices in a pair can have at most one incoming and at most one outgoing horizontal arrow, there are $2^4=16$ types of such pairs. Indeed, choosing the numbers of horizontal arrows and saying that there are, say, $g$ incoming vertical arrows at the bottom determines the other numbers of vertical arrows by the arrows preservation. The weight a pair of vertices\footnote{The total weight of each particular arrow configuration containing several vertices is, by definition, equal to the product of weights of arrow configurations over all individual vertices.} depends on two spectral parameters $u,v$, where $u$ corresponds to the bottom vertex. We will denote pairs of vertices and their weights similarly to \eqref{single_vertex_notation}--\eqref{vertex_weights}, as in the following example: \begin{align*} \bigg[ \oiWoi{.6}g{g+1}g{7pt} \bigg]_{u,v} = \Big[ \Vio{.6}g{g+1}{-4.5pt} \Big]_{u} \Big[ \Voi{.6}{g+1}g{-4.5pt} \Big]_{v}= \frac{(1-t^{g+1})(1-s^2t^g)v}{(1-su)(1-sv)} . \end{align*} We are now in a position to discuss the Yang-Baxter equation. In words, this equation states that the partition function (i.e., the sum of weights of all arrow configurations) in a configuration of a cross vertex followed by a pair of vertices with spectral parameters $u,v$ is the same as the partition function of a pair of vertices with parameters $v,u$ followed by a cross vertex, provided that the boundary conditions on all 6 external edges are the same. (In fact, thus defined partition functions are always sums of at most two terms.) This leads to 16 types of identities \eqref{YB1.1}--\eqref{YB4.4} (each depending on~$g$) which are listed in \Cref{app:YB_equation}. \begin{remark} \label{rmk:YB_equation_numbers} The numbering of identities \eqref{YB1.1}--\eqref{YB4.4} reflects the boundary conditions on the left and right (the first and the second number, respectively). More precisely, equation numbers $\left\{ 1,2,3,4 \right\}$ correspond to the boundary conditions %\begin{noindent} $\left\{ \scalebox{.6}{\begin{tikzpicture} [scale=1.5, ultra thick, baseline=6pt] \draw[dotted] (-.5,0.1)--++(.4,0); \draw[dotted] (-.5,.4)--++(.4,0); \end{tikzpicture}}, \scalebox{.6}{\begin{tikzpicture} [scale=1.5, ultra thick, baseline=6pt] \draw[->] (-.5,0.1)--++(.4,0); \draw[dotted] (-.5,.4)--++(.4,0); \end{tikzpicture}}, \scalebox{.6}{\begin{tikzpicture} [scale=1.5, ultra thick, baseline=6pt] \draw[dotted] (-.5,0.1)--++(.4,0); \draw[->] (-.5,.4)--++(.4,0); \end{tikzpicture}}, \scalebox{.6}{\begin{tikzpicture} [scale=1.5, ultra thick, baseline=6pt] \draw[->] (-.5,0.1)--++(.4,0); \draw[->] (-.5,.4)--++(.4,0); \end{tikzpicture}} \right\}$. %\end{noindent} \end{remark} For example, identity \eqref{YB3.3} among these reads \begin{equation} \label{text_YB3.3} \biggl[ \oiYBoi{.3}{3}{13.5pt}\ioWoi{.6}ggg{6.75pt} \biggr]_{u,v} + \biggl[ \oiYBio{.3}{3}{13.5pt}\oiWoi{.6}g{g+1}g{6.75pt} \biggr]_{u,v} = \biggl[ \ioWoi{.6}g{g}g{6.75pt}\oiYBoi{.3}{3}{13.5pt} \biggr]_{v,u} + \biggl[ \ioWio{.6}g{g-1}g{6.75pt}\ioYBoi{.3}{3}{13.5pt} \biggr]_{v,u}. \end{equation} Here in the left-hand side $u$ is the spectral parameter of the bottom vertex, while in the right-hand side the spectral parameter $u$ is at the top vertex. The weights of the cross vertices in both sides are given by \eqref{cross_vertex_weights} and are not affected by the flipping of the spectral parameters. Writing out \eqref{text_YB3.3} as an identity between rational functions, we obtain: \begin{multline*} \frac{(1-t)v}{u-tv} \frac{(1-st^gu)(v-st^g)}{(1-su)(1-sv)} + \frac{u-v}{u-tv} \frac{(1-t^{g+1})(1-s^2t^g)v}{(1-su)(1-sv)} \\= \frac{(1-st^gv)(u-st^g)}{(1-sv)(1-su)} \frac{(1-t)v}{u-tv} + \frac{(1-s^2t^{g-1})v(1-t^g)}{(1-sv)(1-su)} \frac{t(u-v)}{u-tv}, \end{multline*} which can be readily checked by hand. All other explicit Yang-Baxter identities are listed in \Cref{app:YB_equation}. \subsection{Bijectivisation of the Yang-Baxter equation} \label{sub:YB_bijectivisation_new_label} Our aim is now to bijectivise (in the sense of \Cref{sec:bijectivisation}) each of the 16 types of identities \eqref{YB1.1}--\eqref{YB4.4} given in \Cref{app:YB_equation}. The forward weights corresponding to the Yang-Baxter equation with spectral parameters $u,v$\footnote{That is, in the left-hand side of the Yang-Baxter equation the parameter $u$ is at the bottom vertex, $v$ is at the top vertex, and the weights of the cross vertices in both sides are given by \eqref{cross_vertex_weights}.} will be denoted by $P^{\mathrm{fwd}}_{u,v}$, and the backward ones --- by $P^{\mathrm{bwd}}_{u,v}$. Now, note that both sides of each of the Yang-Baxter identities \eqref{YB1.1}--\eqref{YB4.4} have at most two terms, and so the discussion from \Cref{sub:examples} applies. First, we see that \Cref{ssub:singleton} provides unique bijectivisation of 12 out of 16 types of the Yang-Baxter identities, except \eqref{YB2.2}, \eqref{YB2.3}, \eqref{YB3.2}, and \eqref{YB3.3}. Second, among these four remaining identities, \eqref{YB2.3} and \eqref{YB3.2} are of the form \eqref{2_2_equation_easy_case}, that is, we can identify equal terms on both sides. Thus, let us choose the corresponding natural deterministic bijectivisations of these identities as explained in the end of \Cref{ssub:two_and_two}. Finally, it remains to choose bijectivisations of identities \eqref{YB2.2} and \eqref{YB3.3} for which one cannot deterministically identify terms in both sides. Let us consider \eqref{YB2.2}, identity \eqref{YB3.3} can be treated very similarly. Moreover, for any bijectivisation of the former identity there is a unique bijectivisation of the latter satisfying the symmetries discussed in \Cref{sub:YB_transition_symmetries} below. Thus, having a bijectivisation of \eqref{YB2.2} we will then simply write down the bijectivisation of \eqref{YB3.3} obtained using these symmetries. Identity \eqref{YB2.2} has the form $w(a_1)+w(a_2)=w(b_1)+w(b_2)$, where \begin{equation*} %\begin{noindent} a_1=\; \ioYBio{.3}{3}{13.5pt}\oiWio{.6}ggg{6.75pt}\;,\qquad a_2=\; \ioYBoi{.3}{3}{13.5pt}\ioWio{.6}g{g-1}g{6.75pt}\;,\qquad b_1=\; \oiWio{.6}g{g}{g}{6.75pt}\ioYBio{.3}{3}{13.5pt}\;,\qquad b_2=\; \oiWoi{.6}g{g+1}{g}{6.75pt}\oiYBio{.3}{3}{13.5pt}\;, %\end{noindent} \end{equation*} and the weights are given by (here $g\ge1$ because one of the arrow configurations contains $g-1$ vertical arrows): \begin{align*} w(a_1) & = \frac{(1-t)u}{u-tv} \frac{(u-st^g)(1-st^g v)}{(1-su)(1-sv)} ,\qquad w(a_2)= \frac{u-v}{u-tv} \frac{t(1-t^g)(1-s^2t^{g-1})u}{(1-su)(1-sv)}, \\ w(b_1) & = \frac{(1-t)u}{u-tv} \frac{(v-st^g)(1-st^gu)}{(1-sv)(1-su)} ,\qquad w(b_2)= \frac{u-v}{u-tv} \frac{(1-t^{g+1})(1-s^2t^g)u}{(1-sv)(1-su)}. \end{align*} All bijectivisations of \eqref{YB2.2} form a one-parameter family \eqref{general_2_2_solution} employing the above weights. To select a particular solution out of this one-parameter family, let us argue as follows. Note that $w(a_2)$ vanishes when $u=v$, $t=0$, or $s^2=t^{1-g}$. When $w(a_2)=0$, identity \eqref{YB2.2} simplifies and due to the discussion in \Cref{ssub:singleton} has a unique bijectivisation. In particular, in this case it should be $P^{\mathrm{bwd}}_{u,v}(b_1,a_2)=0$ (i.e., no mass can be transferred into the term $w(a_2)=0$), which means that \begin{equation*} \gamma(u,v,s,t,g)=\frac{w(b_1)}{w(a_1)}= \frac{(v-st^g)(1-st^gu)}{(u-st^g)(1-st^gv)} \qquad \textnormal{when $u=v$, $t=0$, or $s^2=t^{1-g}$.} \end{equation*} We will not address the question of whether the above conditions determine $\gamma(u,v,s,t,g)$ uniquely (in a suitable class of functions), but instead will take $\gamma(u,v,s,t,g)$ equal to the expression in the right-hand side \emph{for all possible values} of $u,v,s,t,g$ (more discussion about the choice of our particular bijectivisation may be found in \Cref{sub:discussion} below). This choice of $\gamma$ leads via \eqref{general_2_2_solution} to the following relatively simple forward and backward transition weights: \begin{equation*} %\begin{noindent} \begin{array}{ll} P^{\mathrm{fwd}}_{u,v}(a_1,b_1)=\dfrac{(v-st^g)(1-st^gu)}{(u-st^g)(1-st^gv)}, & \quad P^{\mathrm{fwd}}_{u,v}(a_1,b_2)=\dfrac{(u-v)(1-s^2t^{2g})}{(u-st^g)(1-st^gv)}, \\[12pt] P^{\mathrm{fwd}}_{u,v}(a_2,b_1)=0 , & \quad P^{\mathrm{fwd}}_{u,v}(a_2,b_2)=1 , \\[12pt] P^{\mathrm{bwd}}_{u,v}(b_1,a_1)=1 , &\quad P^{\mathrm{bwd}}_{u,v}(b_1,a_2)=0 , \\[12pt] P^{\mathrm{bwd}}_{u,v}(b_2,a_1)= \dfrac{(1-t)(1-s^2t^{2g})}{(1-t^{g+1})(1-s^2t^{g})} , &\quad P^{\mathrm{bwd}}_{u,v}(b_2,a_2) = \dfrac{(t-t^{g+1})(1-s^2t^{g-1})} {(1-t^{g+1})(1-s^2t^g)} . \end{array} %\end{noindent} \end{equation*} This is the bijectivisation of identity \eqref{YB2.2} that we will use in the present work. A similar argument leads to the following forward and backward transition weights corresponding to the Yang-Baxter identity \eqref{YB3.3}: \begin{align*} %\begin{noindent} P^{\mathrm{fwd}}_{u,v} \biggl( \oiYBio{.3}{3}{13.5pt}\oiWoi{.6}{g-1}g{g-1}{6.75pt}\ ,\ \ioWoi{.6}{g-1}{g-1}{g-1}{6.75pt}\oiYBoi{.3}{3}{13.5pt} \biggr) &= 1- P^{\mathrm{fwd}}_{u,v} \biggl( \oiYBio{.3}{3}{13.5pt}\oiWoi{.6}{g-1}g{g-1}{6.75pt}\ ,\ \ioWio{.6}{g-1}{g-2}{g-1}{6.75pt}\ioYBoi{.3}{3}{13.5pt} \biggr) = \dfrac{(1-t)(1-s^2t^{2g-2})}{(1-t^{g})(1-s^2t^{g-1})} ; \\ P^{\mathrm{fwd}}_{u,v} \biggl( \oiYBoi{.3}{3}{13.5pt}\ioWoi{.6}{g}g{g}{6.75pt}\ ,\ \ioWoi{.6}{g}{g}{g}{6.75pt}\oiYBoi{.3}{3}{13.5pt} \biggr) &= P^{\mathrm{bwd}}_{u,v} \biggl( \ioWio{.6}{g+1}{g}{g+1}{6.75pt}\ioYBoi{.3}{3}{13.5pt} \ ,\ \oiYBio{.3}{3}{13.5pt}\oiWoi{.6}{g+1}{g+2}{g+1}{6.75pt} \biggr) =1 ; \\ P^{\mathrm{bwd}}_{u,v} \biggl( \ioWoi{.6}{g}{g}{g}{6.75pt}\oiYBoi{.3}{3}{13.5pt} \ ,\ \oiYBoi{.3}{3}{13.5pt}\ioWoi{.6}{g}g{g}{6.75pt} \biggr) &= 1- P^{\mathrm{bwd}}_{u,v} \biggl( \ioWoi{.6}{g}{g}{g}{6.75pt}\oiYBoi{.3}{3}{13.5pt} \ ,\ \oiYBio{.3}{3}{13.5pt}\oiWoi{.6}{g}{g+1}{g}{6.75pt} \biggr) = \frac{(v-st^g)(1-st^gu)}{(u-st^g)(1-st^gv)} . %\end{noindent} \end{align*} All the forward and backward transition weights obtained above are organized into tables in \Cref{fig:fwd_YB,fig:bwd_YB}, respectively. In \Cref{app:YB_probabilities} these weights are listed in full detail. \begin{figure}[htpb] %\begin{noindent} \centering \scalebox{.9}{ $ \begin{array}{c||c||c|c||c|c||c} P^{\mathrm{fwd}}_{u,v}& \ooYBoo{.3}{3}{13.5pt} & \ioYBio{.3}{3}{13.5pt} & \oiYBio{.3}{3}{13.5pt} & \oiYBoi{.3}{3}{13.5pt} & \ioYBoi{.3}{3}{13.5pt} & \iiYBii{.3}{3}{13.5pt} \phantom{\bigg.} \\\hline \ooYBoo{.3}{3}{13.5pt} \phantom{\Bigg.} & 1 & \dfrac{(1-t)v}{u-t v}\dfrac{1-st^{g}u}{1-st^{g}v} & \coli \dfrac{u-v}{u-tv} \dfrac{1-st^{g+1}v}{1-st^{g}v} & \dfrac{(1-t)u}{u-tv} \dfrac{1-st^g v}{1-st^gu} & \colx \dfrac{t(u-v)}{u-tv} \dfrac{1-s t^{g-1}u}{1-s t^gu} & 1 \\\hline \ioYBio{.3}{3}{13.5pt} \phantom{\Bigg.} & 1 & \dfrac{v-st^g}{u-st^g} \dfrac{1-st^gu}{1-st^gv} & \coli \dfrac{u-v}{u-st^g} \dfrac{1-s^2t^{2g}}{1-st^gv} & 1 & \colx0 & 1 \\\hline \oiYBio{.3}{3}{13.5pt} \phantom{\Bigg.} & \colx1 & \colx0 & 1 & \colx \dfrac{1-t}{1-t^{g}} \dfrac{1-s^2t^{2g-2}}{1-s^2t^{g-1}} & \colxx \dfrac{t-t^{g}}{1-t^{g}} \dfrac{1-s^2t^{g-2}}{1-s^2t^{g-1}} & \colx1 \\\hline \oiYBoi{.3}{3}{13.5pt} \phantom{\Bigg.} & 1 & 1 & \coli0 & 1 & \colx0 & 1 \\\hline \ioYBoi{.3}{3}{13.5pt} \phantom{\Bigg.} & \coli1 & \coli0 & \colii1 & \coli0 & 1 & \coli1 \\\hline \iiYBii{.3}{3}{13.5pt} \phantom{\Bigg.} & 1 & \dfrac{(1-t)u}{u-t v}\dfrac{v-st^g}{u-st^g} & \coli\dfrac{u-v}{u-tv} \dfrac{u-st^{g+1}}{u-st^g} & \dfrac{(1-t)v}{u-tv} \dfrac{u-st^{g}}{v-st^{g}} & \colx\dfrac{t(u-v)}{u-tv} \dfrac{v-st^{g-1}}{v-st^{g}} & 1 \end{array} $ } \caption{Forward transition weights corresponding to the Yang-Baxter equation. Here $g$ is the number of vertical arrows in the middle before the move of the cross vertex. The coloring reflects the change of the number of vertical arrows in the middle after the move: pink and red correspond to transitions $g\to g+1$ and $g\to g+2$, while lighter and darker gray mean $g\to g-1$ and $g\to g-2$, respectively.} \label{fig:fwd_YB} %\end{noindent} \end{figure} \begin{figure}[htpb] %\begin{noindent} \centering \scalebox{.9}{ $ \begin{array}{c||c||c|c||c|c||c} P^{\mathrm{bwd}}_{u,v}& \ooYBoo{.3}{3}{13.5pt} & \ioYBio{.3}{3}{13.5pt} & \oiYBio{.3}{3}{13.5pt} & \oiYBoi{.3}{3}{13.5pt} & \ioYBoi{.3}{3}{13.5pt} & \iiYBii{.3}{3}{13.5pt} \phantom{\bigg.} \\\hline \ooYBoo{.3}{3}{13.5pt} \phantom{\Bigg.} & 1 & \dfrac{(1-t)u}{u-tv}\dfrac{1-st^gv}{1-st^gu} & \coli \dfrac{u-v}{u-tv}\dfrac{1-st^{g+1}v}{1-st^gv} & \dfrac{(1-t)v}{u-tv}\dfrac{1-st^gu}{1-st^gv} & \colx \dfrac{t(u-v)}{u-tv}\dfrac{1-st^{g-1}u}{1-st^gu} & 1 \\\hline \ioYBio{.3}{3}{13.5pt} \phantom{\Bigg.} & 1 & 1 & \coli 0 & 1 & \colx 0 & 1 \\\hline \oiYBio{.3}{3}{13.5pt} \phantom{\Bigg.} & \colx 1 & \colx \dfrac{1-t}{1-t^g}\dfrac{1-s^{2}t^{2g-2}}{1-s^2t^{g-1}} & 1 & \colx 0 & \colxx \dfrac{t-t^g}{1-t^g}\dfrac{1-s^2t^{g-2}}{1-s^2t^{g-1}} & \colx 1 \\\hline \oiYBoi{.3}{3}{13.5pt} \phantom{\Bigg.} & 1 & 1 & \coli \dfrac{u-v}{u-st^g}\dfrac{1-s^2t^{2g}}{1-st^gv} & \dfrac{v-st^g}{u-st^g}\dfrac{1-st^gu}{1-st^gv} & \colx 0 & 1 \\\hline \ioYBoi{.3}{3}{13.5pt} \phantom{\Bigg.} & \coli 1 & \coli 0 & \colii 1 & \coli 0 & 1 & \coli 1 \\\hline \iiYBii{.3}{3}{13.5pt} \phantom{\Bigg.} & 1 & \dfrac{(1-t)v}{u-tv}\dfrac{u-st^g}{v-st^g} & \coli \dfrac{u-v}{u-tv}\dfrac{u-st^{g+1}}{u-st^g} & \dfrac{(1-t)u}{u-tv}\dfrac{v-st^g}{u-st^g} & \colx \dfrac{t(u-v)}{u-tv}\dfrac{v-st^{g-1}}{v-st^g} & 1 \end{array} $ } \caption{Backward transition weights corresponding to the Yang-Baxter equation. This table uses the same conventions as in \Cref{fig:fwd_YB}.} \label{fig:bwd_YB} %\end{noindent} \end{figure} \subsection{Symmetries} \label{sub:YB_transition_symmetries} The forward and backward transition weights just defined in \Cref{sub:YB_bijectivisation_new_label} satisfy the following symmetries: \begin{proposition} \label{prop:symm1} Fix any boundary conditions $k_1,k_2,k_1',k_2'\in\left\{ 0,1 \right\}$ and $i_1,i_2\in \mathbb{Z}_{\ge0}$. Then for any $g_1,g_2\in \mathbb{Z}_{\ge0}$ we have the following identity between forward and backward transition weights: \begin{equation*} %\begin{noindent} P^{\mathrm{fwd}}_{u,v} \biggl( \scalebox{.6}{ \begin{tikzpicture} [scale=1.5, very thick, baseline=6.75pt] \draw[densely dashed] (-.75,.25)--(-.5,0)--++(.4,0); \draw[densely dashed] (-.75,.25)--(-.5,.5)--++(.4,0); \draw (-1,0)--(-.75,.25); \draw (-1,.5)--(-.75,.25); \node[anchor=east] at (-1.05,0) {\LARGE$k_1$}; \node[anchor=east] at (-1.05,.5) {\LARGE$k_2$}; \node[anchor=west] at (.52,0) {\LARGE$k_1'$}; \node[anchor=west] at (.52,.5) {\LARGE$k_2'$}; \draw (.1,0)--++(.4,0); \draw (.1,.5)--++(.4,0); \node at (0,-.25) {\LARGE$i_1$}; \node at (0,.25) {\LARGE$g_1$}; \node at (0,.75) {\LARGE$i_2$}; \end{tikzpicture}} \ ,\ \scalebox{.6}{ \begin{tikzpicture} [scale=1.5, very thick, baseline=6.75pt] \draw (-.5,0)--++(.4,0); \draw (-.5,.5)--++(.4,0); \node[anchor=east] at (-.52,0) {\LARGE$k_1$}; \node[anchor=east] at (-.52,.5) {\LARGE$k_2$}; \node[anchor=west] at (1.02,0) {\LARGE$k_1'$}; \node[anchor=west] at (1.02,.5) {\LARGE$k_2'$}; \draw[densely dashed] (.1,0)--++(.4,0)--++(.25,.25); \draw[densely dashed] (.1,.5)--++(.4,0)--++(.25,-.25); \draw (.75,.25)--++(.25,.25); \draw (.75,.25)--++(.25,-.25); \node at (0,-.25) {\LARGE$i_1$}; \node at (0,.25) {\LARGE$g_2$}; \node at (0,.75) {\LARGE$i_2$}; \end{tikzpicture}} \biggr) = P^{\mathrm{bwd}}_{u,v} \biggl( \scalebox{.6}{ \begin{tikzpicture} [scale=1.5, very thick, baseline=6.75pt] \draw (-.5,0)--++(.4,0); \draw (-.5,.5)--++(.4,0); \node[anchor=east] at (-.52,0) {\LARGE$k_2'$}; \node[anchor=east] at (-.52,.5) {\LARGE$k_1'$}; \node[anchor=west] at (1.02,0) {\LARGE$k_2$}; \node[anchor=west] at (1.02,.5) {\LARGE$k_1$}; \draw[densely dashed] (.1,0)--++(.4,0)--++(.25,.25); \draw[densely dashed] (.1,.5)--++(.4,0)--++(.25,-.25); \draw (.75,.25)--++(.25,.25); \draw (.75,.25)--++(.25,-.25); \node at (0,-.25) {\LARGE$i_2$}; \node at (0,.25) {\LARGE$g_1$}; \node at (0,.75) {\LARGE$i_1$}; \end{tikzpicture}} \ ,\ \scalebox{.6}{ \begin{tikzpicture} [scale=1.5, very thick, baseline=6.75pt] \draw[densely dashed] (-.75,.25)--(-.5,0)--++(.4,0); \draw[densely dashed] (-.75,.25)--(-.5,.5)--++(.4,0); \draw (-1,0)--(-.75,.25); \draw (-1,.5)--(-.75,.25); \node[anchor=east] at (-1.05,0) {\LARGE$k_2'$}; \node[anchor=east] at (-1.05,.5) {\LARGE$k_1'$}; \node[anchor=west] at (.52,0) {\LARGE$k_2$}; \node[anchor=west] at (.52,.5) {\LARGE$k_1$}; \draw (.1,0)--++(.4,0); \draw (.1,.5)--++(.4,0); \node at (0,-.25) {\LARGE$i_2$}; \node at (0,.25) {\LARGE$g_2$}; \node at (0,.75) {\LARGE$i_1$}; \end{tikzpicture}} \biggr), %\end{noindent} \end{equation*} with the agreement that weights on both sides are well-defined (i.e., $g_1$ and/or $g_2$ is $\ge1$ if needed). In both weights the numbers of arrows at the boundary are given, and the number of vertical arrows in the middle ($g_1$ or $g_2$) determines the numbers of arrows along the dashed edges connecting the cross vertices with the two-vertex configurations. \end{proposition} \begin{proof} Straightforward verification. \end{proof} \begin{proposition} \label{prop:symm2} For any $k_1,k_2,k_1',k_2'\in \left\{ 0,1 \right\}$ and $i_1,i_2,g_1,g_2\in \mathbb{Z}_{\ge0}$ we have the following symmetry of the forward transition weights with respect to the change $(u,v)\to(v^{-1},u^{-1})$ in the spectral parameters: \begin{multline*} %\begin{noindent} P^{\mathrm{fwd}}_{u,v} \biggl( \scalebox{.6}{ \begin{tikzpicture} [scale=1.5, very thick, baseline=6.75pt] \draw[densely dashed] (-.75,.25)--(-.5,0)--++(.4,0); \draw[densely dashed] (-.75,.25)--(-.5,.5)--++(.4,0); \draw (-1,0)--(-.75,.25); \draw (-1,.5)--(-.75,.25); \node[anchor=east] at (-1.05,0) {\LARGE$k_1$}; \node[anchor=east] at (-1.05,.5) {\LARGE$k_2$}; \node[anchor=west] at (.52,0) {\LARGE$k_1'$}; \node[anchor=west] at (.52,.5) {\LARGE$k_2'$}; \draw (.1,0)--++(.4,0); \draw (.1,.5)--++(.4,0); \node at (0,-.25) {\LARGE$i_1$}; \node at (0,.25) {\LARGE$g_1$}; \node at (0,.75) {\LARGE$i_2$}; \end{tikzpicture}} \ ,\ \scalebox{.6}{ \begin{tikzpicture} [scale=1.5, very thick, baseline=6.75pt] \draw (-.5,0)--++(.4,0); \draw (-.5,.5)--++(.4,0); \node[anchor=east] at (-.52,0) {\LARGE$k_1$}; \node[anchor=east] at (-.52,.5) {\LARGE$k_2$}; \node[anchor=west] at (1.02,0) {\LARGE$k_1'$}; \node[anchor=west] at (1.02,.5) {\LARGE$k_2'$}; \draw[densely dashed] (.1,0)--++(.4,0)--++(.25,.25); \draw[densely dashed] (.1,.5)--++(.4,0)--++(.25,-.25); \draw (.75,.25)--++(.25,.25); \draw (.75,.25)--++(.25,-.25); \node at (0,-.25) {\LARGE$i_1$}; \node at (0,.25) {\LARGE$g_2$}; \node at (0,.75) {\LARGE$i_2$}; \end{tikzpicture}} \biggr) \\= P^{\mathrm{fwd}}_{v^{-1},u^{-1}} \biggl( \scalebox{.6}{ \begin{tikzpicture} [scale=1.5, very thick, baseline=6.75pt] \draw[densely dashed] (-.75,.25)--(-.5,0)--++(.4,0); \draw[densely dashed] (-.75,.25)--(-.5,.5)--++(.4,0); \draw (-1,0)--(-.75,.25); \draw (-1,.5)--(-.75,.25); \node[anchor=east] at (-1.05,0) {\LARGE$1-k_2$}; \node[anchor=east] at (-1.05,.5) {\LARGE$1-k_1$}; \node[anchor=west] at (.52,0) {\LARGE$1-k_2'$}; \node[anchor=west] at (.52,.5) {\LARGE$1-k_1'$}; \draw (.1,0)--++(.4,0); \draw (.1,.5)--++(.4,0); \node at (0,-.25) {\LARGE$i_2$}; \node at (0,.25) {\LARGE$g_1$}; \node at (0,.75) {\LARGE$i_1$}; \end{tikzpicture}} \ ,\ \scalebox{.6}{ \begin{tikzpicture} [scale=1.5, very thick, baseline=6.75pt] \draw (-.5,0)--++(.4,0); \draw (-.5,.5)--++(.4,0); \node[anchor=east] at (-.52,0) {\LARGE$1-k_2$}; \node[anchor=east] at (-.52,.5) {\LARGE$1-k_1$}; \node[anchor=west] at (1.02,0) {\LARGE$1-k_2'$}; \node[anchor=west] at (1.02,.5) {\LARGE$1-k_1'$}; \draw[densely dashed] (.1,0)--++(.4,0)--++(.25,.25); \draw[densely dashed] (.1,.5)--++(.4,0)--++(.25,-.25); \draw (.75,.25)--++(.25,.25); \draw (.75,.25)--++(.25,-.25); \node at (0,-.25) {\LARGE$i_2$}; \node at (0,.25) {\LARGE$g_2$}; \node at (0,.75) {\LARGE$i_1$}; \end{tikzpicture}} \biggr). %\end{noindent} \end{multline*} An analogous identity holds for the backward transition weights. \end{proposition} \begin{proof} This can also be checked in a straightforward way, but the verification can be made shorter with the help of the previous \Cref{prop:symm1}. \end{proof} \subsection{Nonnegativity and probabilistic interpretation} \label{sub:YB_transition_nonnegativity} Let us now address the question of nonnegativity of the forward and backward transition weights obtained in \Cref{sub:YB_bijectivisation_new_label}. \begin{proposition} \label{prop:nonnegative_transition_weights} Assume that our parameters satisfy \begin{equation} \label{weights_nonnegativity_region} 0\le t<1,\qquad -1, densely dashed, line width=2] (2.9,1.5) to [out=45,in=135] (5.7,1.5); \draw[<-, densely dashed, line width=2] (2.9,-.5) to [out=-45,in=-135] (5.7,-.5); \node at (4.3, 1.6) {\LARGE{}forward}; \node at (4.3, -.5) {\LARGE{}backward}; \begin{scope}[shift={(5.6,0)}] \draw[line width=1.8] (.9,1)--++(1.2,0); \draw[line width=1.8] (.9,0)--++(1.2,0); \draw[line width=5] (1.5,-.5)--++(0,2); \draw[line width=1.8] (1.5,0)--++(.6,0)--++(1,1); \draw[line width=1.8] (1.5,1)--++(.6,0)--++(1,-1); \node[circle, draw, fill=white] at (1.5,0) {\large$v$}; \node[circle, draw, fill=white] at (1.5,1) {\large$u$}; \node[anchor=north] at (1.5,-.55) {\LARGE$i_1$}; \node[anchor=south] at (1.5,1.55) {\LARGE$i_2$}; \node[anchor=west] at (1.55,.5) {\LARGE$g_2$}; \node[anchor=east] at (.88,0) {\LARGE$k_1$}; \node[anchor=east] at (.88,1) {\LARGE$k_2$}; \node at (2.2,-.25) {\LARGE$j_1'$}; \node at (2.2,1.28) {\LARGE$j_2'$}; \node[anchor=west] at (3.12,0) {\LARGE$k_1'$}; \node[anchor=west] at (3.12,1) {\LARGE$k_2'$}; \end{scope} \end{tikzpicture} } \caption{Randomized Yang-Baxter moves turning fixed $g_1$ to random $g_2$ or vice versa. Note that the numbers of arrows $j_1,j_2$ or $j_1',j_2'$ in the middle are uniquely determined by $k_1,k_2,k_1',k_2',i_1,i_2$ and $g_1$ or $g_2$, respectively, and thus do not need to be specified explicitly.} \label{fig:probabilistic_transition} \end{figure} \subsection{On the choice of bijectivisation} \label{sub:discussion} In Section \Cref{sub:YB_bijectivisation_new_label} we presented a particular choice of bijectivisation of the Yang-Baxter equation, and the rest of the paper will be devoted to the study of the objects associated with the choice. However, there are other reasonable choices, for which a very similar discussion would be possible. To simplify the exposition, we will not focus on them and just briefly mention possible variations in this section. The Yang-Baxter equation consists of 16 identities between rational functions listed in \Cref{app:YB_equation}. Twelve of them contain only one term in at least one side of an equation and thus have a unique bijectivisation. Identities \eqref{YB2.2}, \eqref{YB2.3}, \eqref{YB3.2}, and \eqref{YB3.3} contain two terms on each side, so according to \Cref{ssub:two_and_two} each of these identities admits a one-parameter family of bijectivisations. It is easy to check that the choice of bijectivisations of these identities presented in \Cref{sub:YB_bijectivisation_new_label} uniquely determined by the following properties: \begin{enumerate} \item (\emph{Nonnegativity}) Transition probabilities are non-negative. \item (\emph{Minimal ``noise'' property}) As many transition probabilities as possible are equal to $0$. \end{enumerate} Indeed, in \eqref{YB2.3} and \eqref{YB3.2} two of the forward probabilities can be made zero, and in \eqref{YB2.2} and \eqref{YB3.3} one forward probability can be made zero. Which of these probabilities are zero is uniquely determined by the non-negativeness. Let us discuss the above conditions. The first one is a must have since we want to obtain a stochastic object. Thus, it forces our four parameters to lie within certain segments of the real line. However, the second condition has a combinatorial flavor which is not crucial for obtaining reasonable probabilistic models. For example, one can introduce another bijectivisation by replacing it with a condition \begin{enumerate} \item [(2')] (\emph{Independence from input}) Forward transition probabilities do not depend on the state of the cross vertex before the move. \end{enumerate} Condition (2') uses the idea of \cite{DiaconisFill1990} (applied in a symmetric function setting in \cite{BorFerr2008DF}). Also, as far as we know, the dynamics coming from condition (2') was used by Andrea Sportiello \cite{Sportiello-private} for simulations in our setting. However, this idea was not applied to bijectivise the Cauchy identity (which requires both forward and backward probabilities) or to construct a random field of signatures (\Cref{sec:YB_field}). We focus on condition (2) rather than (2') (or any other choice of four parameters satisfying condition (1)) because due to less interaction it leads to slightly simpler models. However, since 12 out of 16 identities coming from the Yang-Baxter equation work in the same way for any bijectivisation, all these models are fairly similar. In particular, the dynamic version of the six vertex model (\Cref{sec:dynamicS6V}) and all its degenerations (\Cref{sec:degenerations}) will appear for all bijectivizations. Finally, let us notice that yet another motivation for a certain specific choice of bijectivisation might come from the algebraic side related to the matrix interpretation of the Yang-Baxter equation. We were not able to find a natural condition along these lines. \section{Spin Hall-Littlewood symmetric functions} \label{sec:spin_HL_functions} In this section we recall the symmetric rational functions defined in \cite{Borodin2014vertex} and their basic properties including the Cauchy summation identities. In this section we do not assume that the transition weights are nonnegative. \subsection{Signatures} \label{sub:signatures} We need to introduce some notation. For each $N\in \mathbb{Z}_{\ge1}$ let \begin{equation*} \mathsf{Sign}_{N}:=\left\{ \lambda\in \mathbb{Z}^{N}\colon \lambda_1\ge \ldots\ge\lambda_N \right\} \end{equation*} denote the set of \emph{signatures} with $N$ components.\footnote{Signatures are also sometimes called \emph{highest weights} as the set $\mathsf{Sign}_{N}$ indexes irreducible representations of the unitary group $U(N)$, e.g., see \cite{Weyl1946}.} For $\lambda\in\mathsf{Sign}_{N}$ denote $\ell(\lambda):=N$ and call this the \emph{length} of $\lambda$. By agreement, $\mathsf{Sign}_0$ consists of the single empty signature $\varnothing$. We will also use the notation $|\lambda|:=\lambda_1+\ldots+\lambda_N$. A signature $\lambda\in \mathsf{Sign}_{N}$ is called \emph{nonnegative} if $\lambda_N\ge0$. The set of nonnegative signatures is denoted by $\mathsf{Sign}_{N}^+\subset \mathsf{Sign}_{N}$. Let us set $\mathsf{Sign}:=\bigcup_{N=0}^{\infty}\mathsf{Sign}_{N}$ and $\mathsf{Sign}^+:=\bigcup_{N=0}^{\infty}\mathsf{Sign}_{N}^+$. Nonnegative signatures are often referred to as (integer) \emph{partitions}, which are represented pictorially as \emph{Young diagrams}, e.g., see \cite[Ch. I.1]{Macdonald1995}. While this way of representing signatures is extremely useful in many contexts, we will employ another graphical representation of signatures which works equally well for signatures having negative parts. Namely, associate to each $\mu\in \mathsf{Sign}_{N}$ a configuration of $N$ vertical arrows on $\mathbb{Z}$, with multiple arrows per site allowed, by putting an arrow at each of the locations $\mu_1,\ldots,\mu_N\in \mathbb{Z} $. In other words, write $\mu$ in multiplicative notation as $\mu=\ldots(-1)^{m_{-1}}0^{m_0}1^{m_1}2^{m_2}\ldots $, where $m_i:=\#\left\{ j\colon \mu_j=i \right\}$, $i\in \mathbb{Z}$. Then put $m_i$ vertical arrows at each site $i\in \mathbb{Z}$. Note that all but finitely many sites $i\in \mathbb{Z}$ will be empty. See \Cref{fig:arrow_signature}, left, for an illustration. \begin{figure}[htpb] \centering \begin{tikzpicture}[scale=.7, thick] \draw[->] (-3.5,0)--++(9.5,0); \foreach \ii in {-3,...,5} { \draw (\ii,.1)--++(0,-.2) node[below, yshift=-10] {$\ii$}; } \draw [line width=2,->] (4,-.4)--++(0,.8); \draw [line width=2,->] (3,-.4)--++(0,.8); \draw [line width=2,->] (1.2,-.4)--++(0,.8); \draw [line width=2,->] (1,-.4)--++(0,.8); \draw [line width=2,->] (.8,-.4)--++(0,.8); \draw [line width=2,->] (-2,-.4)--++(0,.8); \end{tikzpicture} \caption{Representing a signature $\mu=(4,3,1,1,1,-2)\in \mathsf{Sign}_6$ as a configuration of $6$ vertical arrows on $\mathbb{Z}$.} \label{fig:arrow_signature} \end{figure} \subsection{Definition of spin Hall-Littlewood functions} \label{sub:spin_HL_definition} Let us now recall the definitions of the symmetric rational functions $F_{\lambda/\mu}$ and $G_{\lambda/\mu}^c$ introduced in \cite{Borodin2014vertex}. Similar objects were also considered earlier as Bethe ansatz eigenfunctions, e.g., see \cite[Ch. VII]{QISM_book}, and also \cite{Povolotsky2013}, \cite{BCPS2014} for more stochastic particle systems connections. We begin by defining versions of the spin Hall-Littlewood functions depending on one variable, the spectral parameter $u\in\mathbb{C}$. \subsubsection{Functions $F_{\lambda/\mu}(u)$} \label{ssub:F_definition} Let a signature $\mu\in \mathsf{Sign}_{N-1}$ \emph{interlace} with a signature $\lambda\in \mathsf{Sign}_{N}$ (notation: $\mu\prec\lambda$) which by definition means that \begin{equation} \label{interlacing_definition} \lambda_{N}\le \mu_{N-1}\le \lambda_{N-1}\le \ldots\le \lambda_2\le \mu_1\le \lambda_1 . \end{equation} There exists a unique configuration of arrows on the grid $\mathbb{Z}\times \left\{ -1,0,1 \right\}$ connecting $\mu$ to $\lambda$ (see \Cref{fig:connecting_interlacing}, left): \begin{itemize} \item vertical arrows $(\mu_i,-1)\to(\mu_i,0)$ entering from the bottom; \item vertical arrows $(\lambda_j,0)\to(\lambda_j,1)$ exiting at the top; \item horizontal arrows along $\mathbb{Z}\times \left\{0\right\}$ such that the local configuration of arrows around each vertex of $\mathbb{Z}\times\left\{0\right\}$ looks like one of the vertices in \Cref{fig:vertex_weights}, and configurations of arrows at neighboring vertices are compatible. There configuration of horizontal arrows is packed at $-\infty$, and is empty at $+\infty$. \end{itemize} \begin{figure}[htpb] \centering \begin{tikzpicture}[scale=.7, thick] \draw (7.5,0)--++(9.5,0); \draw[densely dotted, line width=.4] (7.5,1)--++(9.2,0); \draw[densely dotted, line width=.4] (7.5,-1)--++(9.2,0); \foreach \ii in {-4,...,4} { \draw (\ii+12,.1)--++(0,-.2) node[below, yshift=-25] {$\ii$}; \draw[densely dotted, line width=.4] (\ii+12,-1.3)--++(0,2.6); } \draw [line width=2,->] (15,-1)--++(0,1); \draw [line width=2,->] (15,0)--++(0,1); \draw [line width=2,->] (13.2,-1)--++(0,.9); \draw [line width=2,->] (13.2,-.1)--++(.1,.1)--++(.7,0)--++(0,1); \draw [line width=2,->] (13,-.1)--++(.1,.1)--++(0,1); \draw [line width=2,->] (12.8,-.1)--++(.1,.1)--++(0,1); \draw [line width=2,->] (13,-1)--++(0,.9); \draw [line width=2,->] (12.8,-1)--++(0,.9); \draw [line width=2,->] (10,-1)--++(0,.9); \draw [line width=2,->] (10,-.1)--++(.1,.1)--++(0,1); \draw [line width=2,->] (7,0)--++(1,0); \draw [line width=2,->] (8,0)--++(1,0); \draw [line width=2,->] (9,0)--++(.9,0); \draw [line width=2,->] (9.9,0)--++(0,1); \node at (11.5,1.1) {$\lambda$}; \node at (11.5,-1.1) {$\mu$}; \end{tikzpicture} \qquad \begin{tikzpicture}[scale=.7, thick] \draw (7.5,0)--++(9.5,0); \draw[densely dotted, line width=.4] (7.5,1)--++(9.2,0); \draw[densely dotted, line width=.4] (7.5,-1)--++(9.2,0); \foreach \ii in {-4,...,4} { \draw (\ii+12,.1)--++(0,-.2) node[below, yshift=-25] {$\ii$}; \draw[densely dotted, line width=.4] (\ii+12,-1.3)--++(0,2.6); } \draw [line width=2,->] (15,-1)--++(0,1); \draw [line width=2,->] (15,0)--++(1,0); \draw [line width=2,->] (16,0)--++(1,0); \draw [line width=2,->] (13.2,-1)--++(0,.9); \draw [line width=2,->] (13.2,-.1)--++(.1,.1)--++(.7,0)--++(0,1); \draw [line width=2,->] (13,-.1)--++(.1,.1)--++(0,1); \draw [line width=2,->] (12.8,-.1)--++(.1,.1)--++(0,1); \draw [line width=2,->] (13,-1)--++(0,.9); \draw [line width=2,->] (12.8,-1)--++(0,.9); \draw [line width=2,->] (10,-1)--++(0,.9); \draw [line width=2,->] (10,-.1)--++(.1,.1)--++(0,1); \draw [line width=2,->] (7,0)--++(1,0); \draw [line width=2,->] (8,0)--++(1,0); \draw [line width=2,->] (9,0)--++(.9,0); \draw [line width=2,->] (9.9,0)--++(0,1); \node at (11.5,1.1) {$\nu$}; \node at (11.5,-1.1) {$\mu$}; \end{tikzpicture} \caption{ Left: a configuration of horizontal arrows connecting $\mu=(3,1,1,1,-2)$ to $\lambda=(3,2,1,1,-2,-2)$, with $\mu\prec\lambda$. Right: a configuration of horizontal arrows connecting the same $\mu$ to $\nu=(2,1,1,-2,-2)$, with $\nu\mathop{\dot\prec}\mu$. } \label{fig:connecting_interlacing} \end{figure} For each $m\in \mathbb{Z}$, denote the numbers of incoming and outgoing vertical and horizontal arrows at vertex $m\times\left\{ 0 \right\}$ by $i_{1,2}(m)\in\mathbb{Z}_{\ge0}$ and $j_{1,2}(m)\in\left\{ 0,1 \right\}$, respectively (this notation follows the beginning of \Cref{sub:vertex_weights}). Using this configuration of horizontal arrows connecting $\mu$ to $\lambda$, define \begin{equation} %\begin{noindent} \label{F_skew_one_variable_definition} F_{\lambda/\mu}(u):= \prod_{m=-\infty}^{-1} \frac{\Bigl[ \scalebox{.6}{ \begin{tikzpicture}[scale=1.5, ultra thick, baseline=-4.5pt] \node at (-1,0) {\LARGE$j_1(m)$}; \node at (1,0) {\LARGE$j_2(m)$}; \node at (0,.25) {\LARGE$i_1(m)$}; \node at (0,-.25) {\LARGE$i_2(m)$}; \end{tikzpicture} } \Bigr]_{u}} {\Bigl[ \SVoioi{.6}{-4.5pt} \Bigr]_{u}}\; \prod_{m=0}^{\infty} \Bigl[ \scalebox{.6}{ \begin{tikzpicture}[scale=1.5, ultra thick, baseline=-4.5pt] \node at (-1,0) {\LARGE$j_1(m)$}; \node at (1,0) {\LARGE$j_2(m)$}; \node at (0,.25) {\LARGE$i_1(m)$}; \node at (0,-.25) {\LARGE$i_2(m)$}; \end{tikzpicture} } \Bigr]_{u} , %\end{noindent} \end{equation} where we use notation \eqref{vertex_weights} for the vertex weights depending on the spectral parameter $u$. Observe that both products above are finite since $i_{1,2}(-m)=i_{1,2}(m)=0$, $j_{1,2}(-m)=1$, $j_{1,2}(m)=0$ for all sufficiently large $m$. If $\mu\not\prec\lambda$, set $F_{\lambda/\mu}(u)\equiv 0$. When $\mu,\lambda\in \mathsf{Sign}^+$, $F_{\lambda/\mu}$ defined by \eqref{F_skew_one_variable_definition} coincides with the one given in \cite{Borodin2014vertex}. Moreover, \eqref{F_skew_one_variable_definition} extends the definition so that $F_{\lambda/\mu}$ for arbitrary signatures $\mu\prec\lambda$ satisfies the following translation property: \begin{equation} \label{F_shifting_property} F_{\lambda+(r^{N})/\mu+(r^{N-1})}(u)=\left( \frac{u-s}{1-su} \right)^{r} F_{\lambda/\mu}(u), \qquad \mu\in \mathsf{Sign}_{N-1}, \quad \lambda\in \mathsf{Sign}_N, \end{equation} where in the left-hand side we add arbitrary $r\in \mathbb{Z}$ to all parts of both $\mu$ and $\lambda$. \subsubsection{Functions $G_{\mu/\nu}^c(u)$} \label{ssub:G_definition} Let $\mu,\nu\in \mathsf{Sign}_N$. If these signatures satisfy \begin{equation} \label{interlace2} \nu_N\le \mu_N\le \nu_{N-1}\ldots \mu_2\le \nu_2\le \mu_1, \end{equation} then we also say that $\nu$ and $\mu$ \emph{interlace}, but use a slightly different notation $\nu\mathop{\dot\prec}\mu$ for this. Let us connect $\mu$ to $\nu$ by a configuration of horizontal arrows in the same sense as in \Cref{ssub:F_definition} above. Note that now the ``larger'' signature $\mu$ is placed at the \emph{bottom}. This implies that the configuration of horizontal arrows connecting $\mu$ to $\nu$ contains infinitely many horizontal arrows, both at $-\infty$ and at $+\infty$ (see \Cref{fig:connecting_interlacing}, right). Using this configuration of arrows connecting $\mu$ to $\nu$, define \begin{equation} %\begin{noindent} \label{G_skew_one_variable_definition} G_{\mu/\nu}^{c}(u):= \prod_{m=-\infty}^{+\infty} \frac{\Bigl[ \scalebox{.6}{ \begin{tikzpicture}[scale=1.5, ultra thick, baseline=-4.5pt] \node at (-1,0) {\LARGE$j_1(m)$}; \node at (1,0) {\LARGE$j_2(m)$}; \node at (0,.25) {\LARGE$i_1(m)$}; \node at (0,-.25) {\LARGE$i_2(m)$}; \end{tikzpicture} } \Bigr]_{u^{-1}}} {\Bigl[ \SVoioi{.6}{-4.5pt} \Bigr]_{u^{-1}}}, %\end{noindent} \end{equation} where we used the same notation $i_{1,2}(m), j_{1,2}(m)$ for the numbers of arrows at individual vertices of $\mathbb{Z}\times \left\{ 0 \right\}$ as in \Cref{ssub:F_definition}. Again, observe that the product in \eqref{G_skew_one_variable_definition} is actually finite. If $\nu\mathop{\dot{\not\prec}}\mu$, set $G_{\mu/\nu}^{c}(u)\equiv 0$. \begin{remark} \label{rmk:G_coincides_with_Bor17} Let us connect \eqref{G_skew_one_variable_definition} to the definition of $G_{\mu/\nu}^{c}$ given in \cite{Borodin2014vertex}. Denote \begin{equation*} \Bigl[ \scalebox{.6}{ \begin{tikzpicture}[scale=1.5, ultra thick, baseline=-4.5pt] \node at (-.4,0) {\LARGE$j_1$}; \node at (.4,0) {\LARGE$j_2$}; \node at (0,.25) {\LARGE$i_1$}; \node at (0,-.25) {\LARGE$i_2$}; \end{tikzpicture} } \Bigr]^{\bullet}_{u} := \frac{\Bigl[ \scalebox{.6}{ \begin{tikzpicture}[scale=1.5, ultra thick, baseline=-4.5pt] \node at (-.4,0) {\LARGE$j_1$}; \node at (.4,0) {\LARGE$j_2$}; \node at (0,.25) {\LARGE$i_1$}; \node at (0,-.25) {\LARGE$i_2$}; \end{tikzpicture} } \Bigr]_{u^{-1}}} {\Bigl[ \SVoioi{.6}{-4.5pt} \Bigr]_{u^{-1}}}, \end{equation*} then from \eqref{vertex_weights} we have \begin{equation*} %\begin{noindent} \begin{array}{cll} & \Big[ \Voo{.6}gg{-4.5pt} \Big]_{u}^{\bullet}= \dfrac{u-st^g}{1-su}, & \qquad \Big[ \Voi{.6}g{g-1}{-4.5pt} \Big]_{u}^{\bullet}= \dfrac{1-s^2t^{g-1}}{1-su}, \\[8pt] & \Big[ \Vii{.6}gg{-4.5pt} \Big]_{u}^{\bullet} = \dfrac{1-st^gu}{1-su}, & \qquad \Big[ \Vio{.6}{g}{g+1}{-4.5pt} \Big]_{u}^{\bullet} = \dfrac{(1-t^{g+1})u}{1-su}. \end{array} %\end{noindent} \end{equation*} Observe that in the above graphical definition of $G_{\mu/\nu}^{c}$ the ``larger'' signature $\mu$ is placed at the bottom. Replacing the right-pointing horizontal arrows by empty edges, and vice versa replacing empty edges by \emph{left-pointing} horizontal arrows leads to the conjugated vertex weights $w^{c}_u$ defined in \cite{Borodin2014vertex}: \begin{equation*} %\begin{noindent} \begin{array}{cll} & \Big[ \BVii{.6}gg{-4.5pt} \Big]_{u}^{c}= \dfrac{u-st^g}{1-su}, & \qquad \Big[ \BVio{.6}{g+1}{g}{-4.5pt} \Big]_{u}^{c}= \dfrac{1-s^2t^{g}}{1-su}, \\[8pt] & \Big[ \BVoo{.6}gg{-4.5pt} \Big]_{u}^{c} = \dfrac{1-st^gu}{1-su}, & \qquad \Big[ \BVoi{.6}{g-1}{g}{-4.5pt} \Big]_{u}^{c} = \dfrac{(1-t^{g})u}{1-su}. \end{array} %\end{noindent} \end{equation*} Then $G_{\mu/\nu}^c(u)$ is equal to the product of the conjugated weights $[\cdots]^{c}_u$ similar to \eqref{G_skew_one_variable_definition} but without the denominators (also with $\nu$ at the top and $\mu$ at the bottom). Note that \cite{Borodin2014vertex} also defines functions $G_{\mu/\nu}(u)$ without the conjugation, but we do not use them in the present paper. \end{remark} \subsubsection{Multivariable functions $F$ and $G^c$} \label{ssub:F_G_multivar_definition} Using the single-variable functions \eqref{F_skew_one_variable_definition} and \eqref{G_skew_one_variable_definition}, one can define the corresponding multivariable functions $F$ and $G^c$. Let $K\in \mathbb{Z}_{\ge1}$, $\lambda,\mu\in \mathsf{Sign}$, such that $\ell(\lambda)=\ell(\mu)+K$, $\ell(\mu)=N\in \mathbb{Z}_{\ge0}$. Set \begin{equation} \label{F_skew_multivariable} F_{\lambda/\mu}(u_1,\ldots,u_K ):= \sum_{ \{ \kappa^{(j)} \}} F_{\lambda/\kappa^{(K-1)}}(u_1) F_{\kappa^{(K-1)}/\kappa^{(K-2)}}(u_2)\ldots F_{\kappa^{(1)}/\mu}(u_K), \end{equation} where the sum runs over all $(K-1)$-tuples of signatures $\kappa^{(j)}\in \mathsf{Sign}_{N+j}$, $j=1,\ldots,K-1 $, such that $\mu\prec \kappa^{(1)}\prec\ldots\prec \kappa^{(K-1)}\prec\lambda$. Equivalently, $F_{\lambda/\mu}(u_1,\ldots,u_K )$ can be thought of as the partition function of a path configuration similar to the one in \Cref{fig:connecting_interlacing}, left, but consisting of $K$ horizontal layers. The signatures $\mu$ and $\lambda$ encode, respectively, the bottom and the top boundary conditions, and there are additional $K$ paths entering on the left. The multivariable version of $G^{c}$ is defined in a similar way. Fix $K\in \mathbb{Z}_{\ge1}$, $N\in \mathbb{Z}_{\ge0}$, and let $\mu,\nu\in \mathsf{Sign}_{N}$. Set \begin{equation} \label{G_skew_multivariable} G_{\lambda/\mu}^c(u_1,\ldots,u_K ):= \sum_{ \{\kappa^{(j)}\} } G_{\mu/\kappa^{(K-1)}}^{c}(u_1) G_{\kappa^{(K-1)}/\kappa^{(K-2)}}^{c}(u_2) \ldots G_{\kappa^{(1)}/\nu}(u_K), \end{equation} where the sum is taken over all $(K-1)$-tuples of signatures $\kappa^{(j)}\in \mathsf{Sign}_N$, $j=1,\ldots,K-1 $, satisfying $\nu\mathop{\dot\prec}\kappa^{(1)}\mathop{\dot\prec}\ldots \mathop{\dot\prec}\kappa^{(K-1)}\mathop{\dot\prec}\mu$. Equivalently, $G_{\lambda/\mu}^c(u_1,\ldots,u_K )$ is the partition function of path configurations similar to the one in \Cref{fig:connecting_interlacing}, right, but consisting of $K$ horizontal layers. The signatures $\mu$ and $\nu$ encode, respectively, the bottom and the top boundary conditions. The Yang-Baxter equation for the vertex weights used to define the functions $F_{\lambda/\mu}(u_1,\ldots, u_K)$ and $G_{\mu/\nu}^{c}(u_1,\ldots, u_K)$ readily implies that these functions are symmetric with respect to permutations of the $u_j$'s. See \cite[Theorem 3.5]{Borodin2014vertex} for details. In special cases when the lower diagram is simple, the skew functions $F$ and $G^c$ admit explicit formulas expressing them as sums over permutations. Let us recall such a formula for $F_{\lambda/\varnothing}$. A formula for $G^c_{\mu/(0,\ldots,0 )}$ (where the number of zeros is the same as the number of components in $\mu$) is of similar nature but is more complicated, so we omit it here and refer to \cite[Theorem 5.1]{Borodin2014vertex}, \cite[Theorem 4.14]{BorodinPetrov2016inhom} for details on the statements and their proofs. For the function $F_{\lambda/\varnothing}$ with $\lambda\in \mathsf{Sign}_N^+$ we have \begin{equation} \label{F_symmetrization_formula} F_{\lambda/\varnothing}(u_1,\ldots,u_N ) = \frac{(1-t)^N}{\prod_{i=1}^{N}(1-su_i)} \sum_{\sigma\in S(N)} \prod_{1\le i] (-2,-1)--++(0,1); \draw [line width=2,->] (.9,-1)--++(0,1); \draw [line width=2,->] (1.1,-1)--++(0,1); \draw [line width=2,->] (3,-1)--++(0,1); \node at (0.5,2.1) {$\mu$}; \draw [line width=2,->] (-3.1,1)--++(.1,.1)--++(0,.9); \draw [line width=2,->] (-1,1)--++(0,1); \draw [line width=2,->] (0,1)--++(0,1); \draw [line width=2,->] (1,1)--++(0,1); \draw [line width=2,->] (4,1)--++(0,1); \node at (.5,0.5) {$\kappa$}; \draw [line width=2,->] (-3,0)--++(0,.9); \draw [line width=2,->] (0,0)--++(0,1); \draw [line width=2,->] (2,0)--++(0,1); \draw [line width=2,->] (-5,1)--++(1,0); \draw [line width=2,->] (-4,1)--++(.9,0); \draw [line width=2,->] (-5,0)--++(1,0); \draw [line width=2,->] (-4,0)--++(1,0); \draw [line width=2,->] (-3,.9)--++(.1,.1)--++(.9,0); \draw [line width=2,->] (-2,1)--++(1,0); \draw [line width=2,->] (-2,0)--++(1,0); \draw [line width=2,->] (-1,0)--++(1,0); \draw [line width=2,->] (.9,0)--++(.1,.1)--++(0,.9); \draw [line width=2,->] (1.1,0)--++(.9,0); \draw [line width=2,->] (2,1)--++(1,0); \draw [line width=2,->] (3,1)--++(1,0); \draw [line width=2,->] (3,0)--++(1,0); \draw [line width=2,->] (4,0)--++(1,0); \draw [line width=2,->] (5,0)--++(1,0); \draw [line width=2,->] (6,0)--++(1,0); \end{tikzpicture}\qquad \quad \begin{tikzpicture}[scale=.6, thick] \draw (-4.5,0)--++(10.5,0) node[below right] {$u$}; \draw (-4.5,1)--++(10.5,0) node[above right] {$v$}; \draw[densely dotted, line width=.4] (-4.5,2)--++(10.2,0); \draw[densely dotted, line width=.4] (-4.5,-1)--++(10.2,0); \foreach \ii in {-4,...,5} { \draw (\ii,1.1)--++(0,-.2); \draw (\ii,.1)--++(0,-.2) node[below, yshift=-25] {$\ii$}; \draw[densely dotted, line width=.4] (\ii,-1.3)--++(0,3.6); } \node at (0.5,-1.1) {$\lambda$}; \draw [line width=2,->] (-2,-1)--++(0,.9); \draw [line width=2,->] (.9,-1)--++(0,1); \draw [line width=2,->] (1.1,-1)--++(0,1); \draw [line width=2,->] (3,-1)--++(0,1); \node at (0.5,2.1) {$\mu$}; \draw [line width=2,->] (-3,1)--++(0,1); \draw [line width=2,->] (-1.1,1)--++(.1,.1)--++(0,.9); \draw [line width=2,->] (0,1)--++(0,1); \draw [line width=2,->] (1,1)--++(0,1); \draw [line width=2,->] (4,1)--++(0,1); \node at (.5,0.5) {$\nu$}; \draw [line width=2,->] (-5,1)--++(1,0); \draw [line width=2,->] (-4,1)--++(1,0); \draw [line width=2,->] (-5,0)--++(1,0); \draw [line width=2,->] (-4,0)--++(1,0); \draw [line width=2,->] (-3,0)--++(.9,0); \draw [line width=2,->] (-2.1,0)--++(.1,.1)--++(0,.9); \draw [line width=2,->] (-2,-.1)--++(.1,.1)--++(.9,0); \draw [line width=2,->] (-2,1)--++(.9,0); \draw [line width=2,->] (-1,0)--++(0,.9); \draw [line width=2,->] (-1,.9)--++(.1,.1)--++(.9,0); \draw [line width=2,->] (.9,0)--++(.1,.1)--++(0,.9); \draw [line width=2,->] (1.1,0)--++(.9,0); \draw [line width=2,->] (2,0)--++(0,1); \draw [line width=2,->] (2,1)--++(1,0); \draw [line width=2,->] (3,1)--++(1,0); \draw [line width=2,->] (3,0)--++(1,0); \draw [line width=2,->] (4,0)--++(1,0); \draw [line width=2,->] (5,0)--++(0,1); \draw [line width=2,->] (5,1)--++(1,0); \draw [line width=2,->] (6,1)--++(1,0); \end{tikzpicture} \caption{ Illustration of the sums in the skew Cauchy identity \eqref{skew_Cauchy_identity} with $\lambda=(3,1,1,-2)$, $\mu=(4,1,0,-1,-3)$, and $N=4$. Left: The (finite) sum runs over $\kappa\in\mathsf{Sign}_N$ with $\lambda\mathop{\dot\succ}\kappa\prec \mu$. Right: The (infinite) sum runs over $\nu\in \mathsf{Sign}_{N+1}$ with $\lambda\prec \nu\mathop{\dot\succ}\mu$. Spectral parameters corresponding to the two horizontal layers are also indicated. } \label{fig:skew_Cauchy} %\end{noindent} \end{figure} Via iteration (cf. \eqref{F_skew_multivariable}, \eqref{G_skew_multivariable}), the skew Cauchy identity of \Cref{thm:skew_Cauchy_one} implies the following multivariable identity: \begin{corollary}[Multivariable skew Cauchy identity] \label{cor:skew_multi_Cauchy} Let $u_1,\ldots,u_K,v_1,\ldots,v_L\in \mathbb{C}$ be such that each pair $(u_i,v_j)$ satisfies \eqref{condition_on_convergence}. For any $N\in \mathbb{Z}_{\ge0}$, $\lambda\in \mathsf{Sign}_{N}$, and $\mu\in \mathsf{Sign}_{N+K}$, we have \begin{multline} \label{skew_multi_Cauchy} \sum_{\kappa\in\mathsf{Sign}_N} G_{\lambda/\kappa}^{c}(v_1^{-1},\ldots,v_L^{-1} ) F_{\mu/\kappa}(u_1,\ldots,u_K ) \\= \prod_{i=1}^{K}\prod_{j=1}^{L}\frac{v_j-u_i}{v_j-tu_i} \sum_{\nu\in \mathsf{Sign}_{N+K}} F_{\nu/\lambda}(u_1,\ldots,u_K )\, G_{\nu/\mu}^{c}(v_1^{-1},\ldots,v_L^{-1} ) \end{multline} \end{corollary} Next, setting $\lambda=\varnothing$ and $\mu=(0^K)$ in \Cref{cor:skew_multi_Cauchy}, we get: \begin{corollary}[Ordinary Cauchy identity] Let $u_1,\ldots,u_K,v_1,\ldots,v_L\in \mathbb{C}$ be such that each pair $(u_i,v_j)$ satisfies \eqref{condition_on_convergence}. Then we have \label{cor:nonskew_multi_Cauchy} \begin{equation} \label{nonskew_multi_Cauchy} \prod_{i=1}^{K}\frac{1-t^i}{1-su_i} = \prod_{i=1}^{K}\prod_{j=1}^{L}\frac{v_j-u_i}{v_j-tu_i} \sum_{\nu\in\mathsf{Sign}^+_{K}} F_{\nu/\varnothing}(u_1,\ldots,u_K )\, G_{\nu/(0^K)}^{c}(v_1^{-1},\ldots,v_L^{-1} ) . \end{equation} Note that here the sum runs over nonnegative signatures because all parts of $\mu=(0^K)$ are nonnegative. \end{corollary} % \section{Transition probabilities $\mathsf{U}^{\mathrm{fwd}}$ and $\mathsf{U}^{\mathrm{bwd}}$ on signatures} \label{sec:local_transition_probabilities} In this section, employing the vertex level forward and backward transition probabilities from \Cref{sec:main_construction}, we define the transition probabilities on signatures $\mathsf{U}_{v,u}^{\mathrm{fwd}} ( \kappa\to \nu\mid \lambda,\mu )$ and $\mathsf{U}_{v,u}^{\mathrm{bwd}}( \nu\to\kappa\mid \lambda,\mu )$. The latter probabilities are in particular used to give a new bijective proof of the skew Cauchy identity (\Cref{thm:skew_Cauchy_one}). \subsection{Definition of transition probabilities on signatures} \label{sub:definition_of_three_signature_probabilities} Throughout this section we assume that our parameters satisfy \begin{equation} \label{weights_nonnegativity_region_v_u} 0\le t<1,\qquad -1] (-.5,0.1)--++(.4,0); \draw[->] (-.5,.4)--++(.4,0); \end{tikzpicture} }$ on the far left and $\scalebox{.6}{ \begin{tikzpicture} [scale=1.5, ultra thick, baseline=6pt] \draw[->] (-.5,0.1)--++(.4,0); \draw[dotted] (-.5,.4)--++(.4,0); \end{tikzpicture} }$ on the far right, and, moreover, cannot contain vertical arrows to the left of $\mu_{N+1}$ and to the right of $\lambda_1$. Add the cross vertex $\iiYBii{.25}{3.5}{10.5pt}$ to the left of an arbitrary location $M\le \mu_{N+1}$. Then for each $r=M,M+1,\ldots $ perform the forward randomized Yang-Baxter move which drags the cross to the right through the column number $r$. Let these forward Yang-Baxter moves have probabilities $P_{v,u}^{\mathrm{fwd}}$ given in \Cref{fig:fwd_YB}. This sequence of forward Yang-Baxter moves will not affect the signatures $\lambda,\mu$, and will randomly change $\kappa$, cf. \Cref{fig:U_transition_probabilities}. \begin{figure}[htpb] \centering \begin{tikzpicture}[scale=.8, thick] %\begin{noindent} \draw (-4.5,0)--++(4,0)--++(1,1)--++(6,0); \draw (-4.5,1)--++(4,0)--++(1,-1)-++(6,0); \node at (-5.5,0) {$u$}; \node at (-5.5,1) {$v$}; \node at (7.5,-0) {$v$}; \node at (7.5,1) {$u$}; \draw[densely dotted, line width=.4] (-4.5,2)--++(4,0); \draw[densely dotted, line width=.4] (-4.5,-1)--++(4,0); \foreach \ii in {-4,...,-1} { \draw (\ii,1.1)--++(0,-.2); \draw (\ii,.1)--++(0,-.2) node[below, yshift=-42] {$\ii$}; \draw[densely dotted, line width=.4] (\ii,-1.3)--++(0,3.6); } \foreach \ii in {0,...,5} { \draw (\ii+1,1.1)--++(0,-.2); \draw (\ii+1,.1)--++(0,-.2) node[below, yshift=-42] {$\ii$}; \draw[densely dotted, line width=.4] (\ii+1,-1.3)--++(0,3.6); } \node at (-2,-1.5) {$\lambda_4$}; \node at (1,-1.5) {$\lambda_3$}; \node at (3,-1.5) {$\lambda_2$}; \node at (5,-1.5) {$\lambda_1$}; \draw [line width=2,->] (-2,-1)--++(0,.95); \node at (-3,2.5) {$\mu_5$}; \node at (-1,2.5) {$\mu_4$}; \node at (1,2.5) {$\mu_3$}; \node at (3,2.5) {$\mu_2$}; \node at (6,2.5) {$\mu_1$}; \node at (-2.3,0.5) {$\nu_5$}; \node at (-1.3,0.5) {$\nu_4$}; \node[anchor=west] at (1.1,0.5) {$\kappa_3=\kappa_2$}; \node at (4.4,0.5) {$\kappa_1$}; \draw [line width=2,->] (6,0)--++(1,0); \draw [line width=2,->] (-5,1)--++(1,0); \draw [line width=2,->] (-4,1)--++(1,0); \draw [line width=2,->] (-3,1)--++(0,1); \draw [line width=2,->] (-1,1)--++(0,1); \draw [line width=2,->] (-5,0)--++(1,0); \draw [line width=2,->] (-4,0)--++(1,0); \draw [line width=2,->] (-3,0)--++(.95,0); \draw [line width=2,->] (-2,0)--++(0,1); \draw [line width=2,->] (-2,1)--++(.95,0); \draw [line width=2,->] (-2,0)--++(1,0); \draw [line width=2,->] (-1,0)--++(0,.95); \draw [line width=2,->] (-1,1)--++(.5,0); \draw [line width=2,->] (-.5,1)--++(.5,-.5); \draw [line width=2,->] (0,.5)--++(.5,-.5); \draw [line width=2,->] (.5,0)--++(.4,0); \draw [line width=2,->] (.9,0)--++(0,.9); \draw [line width=2,->] (.9,.9)--++(.1,.1)--++(0,1); \draw [line width=2,->] (1.1,0)--++(0,.9); \draw [line width=2,->] (1.1,.9)--++(.1,.1)--++(.9,0); \draw [line width=2,->] (1,-1)--++(0,1); \draw [line width=2,->] (2,1)--++(1,0); \draw [line width=2,->] (3,1)--++(0,1); \draw [line width=2,->] (3,-1)--++(0,1); \draw [line width=2,->] (3,0)--++(1,0); \draw [line width=2,->] (4,0)--++(0,1); \draw [line width=2,->] (4,1)--++(1,0); \draw [line width=2,->] (5,1)--++(1,0); \draw [line width=2,->] (6,1)--++(0,1); \draw [line width=2,->] (5,-1)--++(0,1); \draw [line width=2,->] (5,0)--++(1,0); \end{tikzpicture} \caption{Performing randomized Yang-Baxter moves to sample $\nu$ given $\kappa$ under $\mathsf{U}_{v,u}^{\mathrm{fwd}}$ (dragging the cross to the right) or $\kappa$ given $\nu$ under $\mathsf{U}_{v,u}^{\mathrm{bwd}}$ (the cross is dragged the left). } \label{fig:U_transition_probabilities} %\end{noindent} \end{figure} \begin{lemma} \label{lemma:cross_far_to_the_right} As $r\to+\infty$, the state of the cross vertex stabilizes at $\oiYBio{.25}{3.5}{10.5pt}$\,. \end{lemma} \begin{proof} Once the cross vertex passes to the right of $\lambda_1$ it can only be in one of two states, $\oiYBio{.25}{3.5}{10.5pt}$ or $\ioYBio{.25}{3.5}{10.5pt}$\,, since the boundary conditions far to the right are $\scalebox{.6}{ \begin{tikzpicture} [scale=1.5, ultra thick, baseline=6pt] \draw[->] (-.5,0.1)--++(.4,0); \draw[dotted] (-.5,.4)--++(.4,0); \end{tikzpicture} }$. From the table in \Cref{fig:fwd_YB} we see that $P_{v,u}^{\mathrm{fwd}}\left( \oiYBio{.25}{3.5}{10.5pt},\oiYBio{.25}{3.5}{10.5pt} \right)=1$. Moreover, since there are no vertical arrows to the right of $\lambda_1$, we have $P_{v,u}^{\mathrm{fwd}}\left( \ioYBio{.25}{3.5}{10.5pt},\oiYBio{.25}{3.5}{10.5pt} \right) = 1-\frac{(u-s)(1-sv)}{(v-s)(1-su)}$, which is strictly positive by \eqref{condition_on_convergence}. Therefore, the state $\ioYBio{.25}{3.5}{10.5pt}$ of the cross vertex eventually turns into $\oiYBio{.25}{3.5}{10.5pt}$ with probability 1 (which in fact corresponds to choosing $\nu_1$ somewhere to the right of $\lambda_1$), and the latter state is preserved forever. \end{proof} We see that the process of (randomized) dragging of the cross vertex to the right essentially terminates. Cutting cross vertex $\oiYBio{.25}{3.5}{10.5pt}$ which has stabilized far on the right, we obtain the final two-layer arrow configuration which looks as in \Cref{fig:skew_Cauchy}, right. That is, the boundary conditions are now $\scalebox{.6}{ \begin{tikzpicture} [scale=1.5, ultra thick, baseline=6pt] \draw[->] (-.5,0.1)--++(.4,0); \draw[->] (-.5,.4)--++(.4,0); \end{tikzpicture} }$ on the far left and $\scalebox{.6}{ \begin{tikzpicture} [scale=1.5, ultra thick, baseline=6pt] \draw[dotted] (-.5,0.1)--++(.4,0); \draw[->] (-.5,.4)--++(.4,0); \end{tikzpicture} }$ on the far right, while the fixed signature $\kappa\in\mathsf{Sign}_N$ in the middle has been replaced by a \emph{random} signature $\nu\in \mathsf{Sign}_{N+1}$. Moreover, this new signature satisfies $\lambda\prec \nu\mathop{\dot\succ}\mu$ because in the final two-layer configuration there can be at most one horizontal arrow per edge. \begin{definition} \label{def:Ufwd} The law of the random signature $\nu\in \mathsf{Sign}_{N+1}$ described above will be denoted by $\mathsf{U}_{v,u}^{\mathrm{fwd}}(\kappa\to\nu\mid \lambda,\mu)$. We will call $\mathsf{U}_{v,u}^{\mathrm{fwd}}$ the \emph{forward transition probabilities} (\textit{on signatures}). \end{definition} The backward transition probabilities $\mathsf{U}_{v,u}^{\mathrm{bwd}} \left( \nu\to \kappa\mid \lambda,\mu \right)$ (where the signatures $\lambda\in\mathsf{Sign}_N$, $\nu,\mu\in\mathsf{Sign}_{N+1}$ with $\lambda\prec \nu\mathop{\dot\succ}\mu$ are given) are defined in a similar way, but now the cross vertex $\oiYBio{.25}{3.5}{10.5pt}$ is added to the right of $\nu_1$ and is dragged to the left using the backward Yang-Baxter moves having probabilities $P_{v,u}^{\mathrm{bwd}}$ given in \Cref{fig:bwd_YB}. The process of dragging the cross vertex to the left terminates at $\mu_{N+1}$ when the cross vertex has the state $\iiYBii{.25}{3.5}{10.5pt}$\,. This process does not affect the signatures $\lambda$ and $\mu$, and turns the fixed signature $\nu\in \mathsf{Sign}_{N+1}$ in the middle into a \emph{random} signature $\kappa\in \mathsf{Sign}_N$. \begin{definition} \label{def:Ubwd} The law of the random signature $\kappa\in \mathsf{Sign}_{N}$ just described will be denoted by $\mathsf{U}_{v,u}^{\mathrm{bwd}}(\nu\to\kappa\mid \lambda,\mu)$. We will call $\mathsf{U}_{v,u}^{\mathrm{bwd}}$ the \emph{backward transition probabilities} (\emph{on signatures}). \end{definition} Clearly, by the very construction, \begin{equation} \label{transition_probabilities_U_sum_to_one} \begin{split} \sum_{\nu\in \mathsf{Sign}_{N+1}} \mathsf{U}_{v,u}^{\mathrm{fwd}} \left( \kappa\to \nu\mid \lambda,\mu \right)&=1,\qquad \textnormal{for every $\lambda,\kappa,\mu$ with $\lambda\mathop{\dot\succ}\kappa\prec \mu$};\\ \sum_{\kappa\in \mathsf{Sign}_{N}} \mathsf{U}_{v,u}^{\mathrm{bwd}} \left( \nu\to \kappa\mid \lambda,\mu \right)&=1,\qquad \textnormal{for every $\lambda,\nu,\mu$ with $\lambda\prec \nu\mathop{\dot\succ}\mu$}. \end{split} \end{equation} The first of these sums is infinite and converges due to \eqref{weights_nonnegativity_region_v_u}. The second of the sums is finite. \begin{proposition} \label{prop:U_are_products_of_P} Let $N\in \mathbb{Z}_{\ge0}$, $\lambda,\kappa\in\mathsf{Sign}_N$, and $\mu,\nu\in\mathsf{Sign}_{N+1}$ be fixed. The forward transition probability $\mathsf{U}_{v,u}^{\mathrm{fwd}}(\kappa\to\nu\mid \lambda,\mu)$ on signatures is equal to the product of finitely many local forward transition probabilities $P_{v,u}^{\mathrm{fwd}}$ over columns with numbers from $\mu_{N+1}$ to $\nu_1$. Similarly, $\mathsf{U}_{v,u}^{\mathrm{bwd}}(\nu\to\kappa\mid \lambda,\mu)$ is the product of finitely many local backward transition probabilities $P_{v,u}^{\mathrm{bwd}}$ over columns from $\mu_{N+1}$ to $\nu_1$. \end{proposition} \begin{proof} We argue only about forward transition probabilities, the case of the backward ones is analogous. Let the multiplicative notations of the signatures $\lambda,\kappa,\nu,\mu$ be $\lambda=\ldots(-1)^{\ell_{-1}}0^{\ell_0}1^{\ell_1}2^{\ell_2}\ldots $, $\kappa=\ldots(-1)^{k_{-1}}0^{k_0}\ldots $, $\nu=\ldots(-1)^{n_{-1}}0^{n_0}\ldots $, and $\mu=\ldots(-1)^{m_{-1}}0^{m_0}\ldots $. Consider the situation in the definition of $\mathsf{U}_{v,u}^{\mathrm{fwd}}$ when the cross vertex is moved through the column number $r$ (for example, $r=0$ in \Cref{fig:U_transition_probabilities}). Assume that the following data is known before the move of the cross vertex: \begin{itemize} \item The state of the cross vertex (i.e., one of six states as in \eqref{cross_vertex_weights}); \item The numbers $\ell_r,k_r,m_r$ of vertical arrows at the $r$-th column before the move of the cross vertex; \item The numbers $\ell_r,n_r,m_r$ of vertical arrows at the $r$-th column after the move of the cross vertex; \item The numbers of horizontal arrows in both layers of the arrow configuration as in \Cref{fig:U_transition_probabilities} between the $(r-1)$-st column and the cross vertex, as well as between the $r$-th and the $(r+1)$-st columns. \end{itemize} One readily sees that the state of the cross vertex after the forward randomized Yang-Baxter move (placing the cross vertex one step to the right) is completely determined by the above data. The state of the cross vertex and all the above data at the far left is known. Therefore, by induction all the intermediate states of the cross vertex in the definition of $\mathsf{U}_{v,u}^{\mathrm{fwd}}(\kappa\to\nu\mid \lambda,\mu)$ are completely determined by the four signatures $\lambda,\kappa,\nu,\mu$. This implies that the transition probability $\mathsf{U}_{v,u}^{\mathrm{fwd}}(\kappa\to\nu\mid \lambda,\mu)$ on signatures is indeed equal to the product of the local transition probabilities $P_{v,u}^{\mathrm{fwd}}$ depending on these intermediate cross vertex states. This completes the proof. \end{proof} \subsection{Bijective proof of the skew Cauchy identity} \label{sub:bijective_proof_skew_Cauchy} The key observation leading to our bijective proof of \Cref{thm:skew_Cauchy_one} is the following \begin{proposition}[Reversibility on signatures] \label{prop:reversibility_on_signatures} Fix arbitrary $N\in \mathbb{Z}_{\ge0}$, $\lambda,\kappa\in\mathsf{Sign}_N$, and $\mu,\nu\in\mathsf{Sign}_{N+1}$. We have for any $(u,v)$ satisfying \eqref{weights_nonnegativity_region_v_u}: \begin{equation} \label{reversibility_on_signatures} \left[ \iiYBii{.35}{3}{14.5pt} \right]_{v,u} G_{\lambda/\kappa}^{c}(v^{-1})F_{\mu/\kappa}(u) \mathsf{U}_{v,u}^{\mathrm{fwd}}(\kappa\to\nu\mid \lambda,\mu) = \left[ \oiYBio{.35}{3}{14.5pt} \right]_{v,u} F_{\nu/\lambda}(u)\,G_{\nu/\mu}^{c}(v^{-1}) \mathsf{U}_{v,u}^{\mathrm{bwd}}(\nu\to\kappa\mid \lambda,\mu), \end{equation} where the weights of the cross vertices are given in \eqref{cross_vertex_weights} (modulo the swap, cf. \Cref{rmk:swap_u_v}). \end{proposition} \begin{remark} \label{rmk:reversibility_on_signatures} Both sides of \eqref{reversibility_on_signatures} are nonzero only if $\lambda\mathop{\dot\succ}\kappa\prec \mu$ and $\lambda\prec \nu\mathop{\dot\succ}\mu$. Indeed, if, say, the condition $\kappa\prec \mu$ is violated, then $F_{\mu/\kappa}(u)$ is zero by the very definition. At the same time $\mathsf{U}_{v,u}^{\mathrm{bwd}}(\nu\to\kappa\mid \lambda,\mu)$ also vanishes because $\kappa\not\prec\mu$ implies that $\kappa$ cannot arise as the middle signature in the two-layer arrow configuration after dragging the cross vertex from far right to the left. \end{remark} \begin{proof}[Proof of \Cref{prop:reversibility_on_signatures}] By \eqref{F_skew_one_variable_definition}, \eqref{G_skew_one_variable_definition} and \Cref{prop:U_are_products_of_P}, the skew functions $F, G^c$ as well as the transition probabilities $\mathsf{U}_{v,u}$ in both sides of \eqref{reversibility_on_signatures} can be expressed as products over the columns in the two-layer arrow configurations as in \Cref{fig:skew_Cauchy}. The desired identity \eqref{reversibility_on_signatures} then follows by repeatedly applying the local reversibility condition at each column for the probabilities of the Yang-Baxter moves $P_{v,u}^{\mathrm{fwd}}$ and $P_{v,u}^{\mathrm{bwd}}$. The local reversibility condition is satisfied by the very construction of the latter probabilities, see \Cref{def:bijectivisation} and \Cref{sub:YB_bijectivisation_new_label}. The quantities $\left[ \iiYBii{.25}{3.5}{10.5pt} \right]_{v,u} G_{\lambda/\kappa}^{c}(v^{-1})F_{\mu/\kappa}(u) $ and $\left[ \oiYBio{.25}{3.5}{10.5pt} \right]_{v,u} F_{\nu/\lambda}(u)\,G_{\nu/\mu}^{c}(v^{-1})$ collect the weights entering the local reversibility conditions, while the probabilities $\mathsf{U}_{v,u}^{\mathrm{fwd}},\mathsf{U}_{v,u}^{\mathrm{bwd}}$ collect the local probabilities $P_{v,u}^{\mathrm{fwd}},P_{v,u}^{\mathrm{bwd}}$. This implies \eqref{reversibility_on_signatures}. \end{proof} \begin{proof}[Proof of \Cref{thm:skew_Cauchy_one}] Summing \eqref{reversibility_on_signatures} over both $\kappa\in\mathsf{Sign}_{N}$ and $\nu\in \mathsf{Sign}_{N+1}$ and recalling that $\left[\iiYBii{.25}{3.5}{10.5pt} \right]_{v,u}=1$ and $\left[ \oiYBio{.25}{3.5}{10.5pt} \right]_{v,u}=\frac{v-u}{v-tu}$, we have \begin{multline} \label{skew_Cauchy_identity_bijective_proof} \sum_{\kappa\in \mathsf{Sign}_N} G_{\lambda/\kappa}^{c}(v^{-1})F_{\mu/\kappa}(u) \Bigg(\sum_{\nu\in \mathsf{Sign}_{N+1}} \mathsf{U}_{v,u}^{\mathrm{fwd}}(\kappa\to\nu\mid \lambda,\mu) \Bigg) \\ = \frac{v-u}{v-tu} \sum_{\nu\in \mathsf{Sign}_{N+1}} F_{\nu/\lambda}(u)\,G_{\nu/\mu}^{c}(v^{-1}) \Bigg( \sum_{\kappa\in \mathsf{Sign}_N} \mathsf{U}_{v,u}^{\mathrm{bwd}}(\nu\to\kappa\mid \lambda,\mu) \Bigg). \end{multline} By \eqref{transition_probabilities_U_sum_to_one}, the sums in the parentheses in both sides are equal to $1$, which implies the desired identity \eqref{skew_Cauchy_identity}. \end{proof} We call the above proof of the skew Cauchy identity \eqref{skew_Cauchy_identity} \emph{bijective} because \eqref{skew_Cauchy_identity_bijective_proof} provides a \emph{refinement} of \eqref{skew_Cauchy_identity} (involving summation over $\kappa,\nu$ in both sides), in which the terms in both sides are bijectively identified with each other with the help of the reversibility condition \eqref{reversibility_on_signatures}. Thus, the transition probabilities $\mathsf{U}_{v,u}^{\mathrm{fwd}}$ and $\mathsf{U}_{v,u}^{\mathrm{bwd}}$ show how to split terms in both sides of the original identity \eqref{skew_Cauchy_identity} into smaller ones, such that these smaller terms are identified with each other. \subsection{Markov projection of the forward transition onto first columns} \label{sub:properties_of_global_transitions} For notational convenience, in this subsection we assume that both $\lambda$ and $\mu$ are nonnegative signatures (i.e., whose parts are all nonnegative). Then the signatures $\kappa,\nu$ entering $\mathsf{U}_{v,u}^{\mathrm{fwd}}(\kappa\to \nu\mid \lambda,\mu)$ (as well as $\mathsf{U}_{v,u}^{\mathrm{bwd}}(\nu\to\kappa\mid \lambda,\mu)$) should also be nonnegative, otherwise these transition probabilities vanish for interlacing reasons. Fix any $h\in \mathbb{Z}_{\ge1}$. For any nonnegative signature $\rho$ having multiplicative notation $\rho=0^{r_0}1^{r_1}2^{r_2}\ldots $, let $\rho^{[] (-.5,0)--++(7,0); \draw[->] (0,-.5)--++(0,4); \foreach \ii in {1,2,3,4,5,6} { \node at (\ii-.5,-.4) {$v_\ii$}; \draw[dotted, thick] (\ii,-.5)--++(0,3.75); } \foreach \jj in {1,2,3} { \node at (-.4,\jj-.5) {$u_\jj$}; \draw[dotted, thick] (-.5,\jj)--++(6.75,0); } \draw [red, line width=2] (0,3)--++(3,0)--++(0,-1)--++(1,0)--++(0,-1)--++(2,0)--++(0,-1); \foreach \p in {(0,3),(3,3),(3,2),(4,2),(4,1),(6,1),(6,0)} { \draw[red,fill] \p circle(3pt); } \node at (0,-.9) {$x_1$}; \node at (3,-.9) {$x_2$}; \node at (4,-.9) {$x_3$}; \node at (6,-.9) {$x_4$}; \node at (-1,0) {$y_4$}; \node at (-1,1) {$y_3$}; \node at (-1,2) {$y_2$}; \node at (-1,3) {$y_1$}; \node at (1.5,3.35) {$G^c$}; \node at (3.7,2.35) {$G^c$}; \node at (5.4,1.35) {$G^c$}; \node at (2.7,2.5) {$F$}; \node at (3.7,1.5) {$F$}; \node at (5.7,0.5) {$F$}; \node at (0,3) (p1) {}; \node at (3,3) (p2) {}; \node at (3,2) (p3) {}; \node at (4,2) (p4) {}; \node at (4,1) (p5) {}; \node at (6,1) (p6) {}; \node at (6,0) (p7) {}; \node[anchor=east] (lp1) at (-.5,4) {$\lambda^{(0,3)}=(0^{3})$}; \node (lp2) at (3.6,3.3) {$\lambda^{(3,3)}$}; \node (lp3) at (2.6,1.7) {$\lambda^{(3,2)}$}; \node (lp4) at (4.6,2.3) {$\lambda^{(4,2)}$}; \node (lp5) at (4.6,1.3) {$\lambda^{(4,1)}$}; \node[anchor=west] (lp6) at (6.5,2) {$\lambda^{(6,1)}$}; \node[anchor=west] (lp7) at (6.5,1) {$\lambda^{(6,0)}=\varnothing$}; \draw[line width=.7,dashed] (p1)--(lp1); \draw[line width=.7,dashed] (p6)--(lp6); \draw[line width=.7,dashed] (p7)--(lp7); \end{tikzpicture} \caption{An illustration of the spin Hall-Littlewood process indexed by sequences $\vec{x}=(0,3,4,6)$ and $\vec{y}=(3,2,1,0)$. The second product (over in $(i,j)$) in \eqref{spin_HL_process_normalization} runs over all boxes inside the region bounded by the down-right path $\mathcal{P}_{\vec{x},\vec{y}}$. For this particular path the product contains 13 terms.} \label{fig:spin_HL_process} %\end{noindent} \end{figure} One of the properties of spin Hall-Littlewood processes is that the marginal distribution of each single signature $\lambda^{(x,y)}\in\mathsf{Sign}^+_y$, $(x,y)\in \mathcal{P}_{\vec{x},\vec{y}}$, under $\mathscr{HP}_{\vec{x},\vec{y}}$ \eqref{spin_HL_process} is given by the spin Hall-Littlewood measure $\mathscr{H}_{x,y}$ \eqref{spin_HL_measure}. More generally, take any subpath $\mathcal{Q}$ of $\mathcal{P}_{\vec{x},\vec{y}}$ such that $\mathcal{Q}$ is itself a down-right path. Then the marginal distribution of the signatures $\{\lambda^{q}\colon q\in \mathcal{Q}\}$ under the original spin Hall-Littlewood process $\mathscr{HP}_{\vec{x},\vec{y}}$ \eqref{spin_HL_process} is itself a spin Hall-Littlewood process corresponding to the path $\mathcal{Q}$. \subsection{Yang-Baxter field} \label{sub:YB_random_field} Let us now introduce the Yang-Baxter field with the help of the forward transition probabilities on signatures discussed in \Cref{sec:local_transition_probabilities}. The field depends on $t\in[0,1)$, $s\in(-1,0]$, and two sequences of spectral parameters $v_1,v_2,\ldots $, $u_1,u_2,\ldots $ such that $0\le u_i] (-.8,0)--++(7.6,0) node[above] {$\mathop{\dot\prec}$}; \draw[->] (0,-.8)--++(0,4.6) node[above] {$\prec$}; \foreach \ii in {1,2,3,4,5,6} { \node at (\ii-.5,-.4) {$v_\ii$}; \draw[dotted, thick] (\ii,-.7)--++(0,4.3); } \foreach \jj in {1,2,3} { \node at (-.4,\jj-.5) {$u_\jj$}; \draw[dotted, thick] (-.7,\jj)--++(7.3,0); } \foreach \ii in {0,...,5} {\node[rectangle,draw,fill=white] at (\ii,0) {$\varnothing$};} \node[rectangle,draw,fill=white] at (0,1) {$(0)$}; \node[rectangle,draw,fill=white] at (0,2) {$(00)$}; \node[rectangle,draw,fill=pink] at (0,3) {$(000)$}; \node[rectangle,draw,fill=pink] at (6,0) {$\varnothing$}; \node[rectangle,draw,fill=white] at (1,1) {$\boldsymbol\lambda^{(1,1)}$}; \node[rectangle,draw,fill=white] at (2,1) {$\boldsymbol\lambda^{(2,1)}$}; \node[rectangle,draw,fill=white] at (3,1) {$\boldsymbol\lambda^{(3,1)}$}; \node[rectangle,draw,fill=pink] at (4,1) {$\boldsymbol\lambda^{(4,1)}$}; \node[rectangle,draw,fill=pink] at (5,1) {$\boldsymbol\lambda^{(5,1)}$}; \node[rectangle,draw,fill=pink] at (6,1) {$\boldsymbol\lambda^{(6,1)}$}; \node[rectangle,draw,fill=white] at (1,2) {$\boldsymbol\lambda^{(1,2)}$}; \node[rectangle,draw,fill=white] at (2,2) {$\boldsymbol\lambda^{(2,2)}$}; \node[rectangle,draw,fill=pink] (kappa) at (3,2) {$\boldsymbol\lambda^{(3,2)}$}; \node[rectangle,draw,fill=pink] at (4,2) {$\boldsymbol\lambda^{(4,2)}$}; \node[rectangle,draw,fill=white] at (5,2) {$\boldsymbol\lambda^{(5,2)}$}; \node[rectangle,draw,fill=white] at (6,2) {$\boldsymbol\lambda^{(6,2)}$}; \node[rectangle,draw,fill=pink] at (1,3) {$\boldsymbol\lambda^{(1,3)}$}; \node[rectangle,draw,fill=pink] at (2,3) {$\boldsymbol\lambda^{(2,3)}$}; \node[rectangle,draw,fill=pink] at (3,3) {$\boldsymbol\lambda^{(3,3)}$}; \node[rectangle,draw,dashed,fill=gray!40!white] (nu) at (4,3) {$\boldsymbol\lambda^{(4,3)}$}; \node[rectangle,draw,fill=white] at (5,3) {$\boldsymbol\lambda^{(5,3)}$}; \node[rectangle,draw,fill=white] at (6,3) {$\boldsymbol\lambda^{(6,3)}$}; \draw[->,dotted,red,line width=2] (kappa)--(nu.south west); \end{tikzpicture} \caption{Yang-Baxter random field. Signatures along a down-right path (the extension of the path in \Cref{fig:spin_HL_process}) are highlighted in red. The signature $\boldsymbol\lambda^{(3,2)}$ is replaced by $\boldsymbol\lambda^{(4,3)}$ in this path with the help of the forward transition probability, cf. the proof of \Cref{thm:YB_field_spin_HL_process}.} \label{fig:YB_field} %\end{noindent} \end{figure} The discussion in \Cref{sub:properties_of_global_transitions} readily implies the following Markov projection property of the Yang-Baxter field: \begin{proposition} \label{prop:YB_Markov_projections} Fix any $h\in \mathbb{Z}_{\ge1}$. Under the Yang-Baxter field, the first $h$ columns of the signatures $\boldsymbol\lambda^{(x,y)}$ evolve in a marginally Markovian way (i.e., independently of the columns $h+1,h+2,\ldots $). \end{proposition} This evolution of the first $h$ columns defines a random field indexed by $\mathbb{Z}_{\ge0}^{2}$ with values in $\mathbb{Z}_{\ge0}^{h}$ which can be regarded as an $h$-layer stochastic vertex model. In \Cref{sec:dynamicS6V,sec:degenerations} we discuss the case $h=1$ in detail. Details on the two-layer case for $s=0$ may be found in \cite[Section 4.4]{BufetovMatveev2017}. The next theorem states a key property of the Yang-Baxter field $\boldsymbol\Lambda$: \begin{theorem} \label{thm:YB_field_spin_HL_process} Under the Yang-Baxter field, for any down-right path $\mathcal{P}_{\vec{x},\vec{y}}$ as in \eqref{spin_HL_down_right_path_sequences}--\eqref{spin_HL_down_right_path}, the joint distribution of the signatures $\{\boldsymbol\lambda^p\colon p\in \mathcal{P}_{\vec{x},\vec{y}}\}$ is given by the spin Hall-Littlewood process $\mathscr{HP}_{\vec{x},\vec{y}}$ \eqref{spin_HL_process}. \end{theorem} \begin{proof} Extend the path $\mathcal{P}_{\vec{x},\vec{y}}$ by adding to it all the intermediate vertices, so that the distance between each two consecutive vertices along the extended path is equal to $1$ (cf. \Cref{fig:YB_field}). Let us also add vertices $(0,y_1+1)$ and $(x_k+1,0)$ in the beginning and the end of the path, respectively. If we establish the claim for such extended paths, then the original claim will follow, cf. the remark in the end of \Cref{sub:spin_HL_measures_processes}. Using the inductive definition of $\boldsymbol\Lambda$, we establish the modified claim by induction on the down-right path. The base of the induction is the case when the path goes along the coordinate axes, i.e., has the form $\left\{ (0,y_1+1),(0,y_1),\ldots, (0,1),(0,0),(1,0),\ldots,(x_k+1,0)\right\}$. In this case the random signatures along this path are in fact deterministic, and coincide with the corresponding signatures under the spin Hall-Littlewood process corresponding to this path. In the induction step, we replace one down-right corner of the form $\{(x,y+1),(x,y),(x+1,y)\}$ by the right-down corner $\{(x,y+1),(x+1,y+1),(x+1,y)\}$ (see an example in \Cref{fig:YB_field} where $(x,y)=(3,2)$). Denote the old and the new paths by $\mathcal{P}$ and $\mathcal{P}'$, respectively. For shorter notation, set \begin{equation*} \kappa:=\boldsymbol\lambda^{(x,y)},\qquad \mu:=\boldsymbol\lambda^{(x,y+1)}, \qquad \lambda:=\boldsymbol\lambda^{(x+1,y)}, \qquad \nu:=\boldsymbol\lambda^{(x+1,y+1)}. \end{equation*} Assume that the joint distribution of the signatures along $\mathcal{P}$ is given by the corresponding spin Hall-Littlewood process. The joint distribution along $\mathcal{P}'$ can be obtained from the joint distribution along $\mathcal{P}$ with the help of the conditional distribution of $\nu$ given $\lambda,\kappa,\mu$. By \Cref{def:YB_field}, the latter conditional distribution is given by the forward transition probability. Thus, we see that the joint distribution of all four signatures $\lambda,\kappa,\mu,\nu$ is proportional to the left-hand side of \eqref{reversibility_on_signatures} (with $u=u_{y+1}$, $v=v_{x+1}$). Using this identity and summing over $\kappa$, we see from the right-hand side of \eqref{reversibility_on_signatures} that the joint distribution of $\lambda,\nu,\mu$ is proportional to $F_{\nu/\lambda}(u_{y+1})\,G_{\nu/\mu}^{c}(v^{-1}_{x+1})$, as it should be under the spin Hall-Littlewood process corresponding to the path $\mathcal{P}'$. This completes the induction step and the proof of the proposition. \end{proof} \Cref{thm:YB_field_spin_HL_process} and \Cref{prop:reversibility_on_signatures} readily imply a backward version of the conditional distribution \eqref{YB_field_definition_forward} in the Yang-Baxter field: \begin{corollary} \label{cor:YB_field_bwd_conditional_distr} Under the Yang-Baxter field, for any $(x,y)\in \mathbb{Z}_{\ge0}$ the conditional distribution of $\boldsymbol\lambda^{(x,y)}$ given the signatures to the right and above it is equal to the backward transition probability: \begin{equation*} \mathrm{Prob}(\boldsymbol\lambda^{(x,y)}=\kappa\mid \boldsymbol\lambda^{(x+1,y)},\boldsymbol\lambda^{(x,y+1)}, \boldsymbol\lambda^{(x+1,y+1)}) = \mathsf{U}^{\mathrm{bwd}}_{v_{x+1},u_{y+1}} \bigl( \boldsymbol\lambda^{(x+1,y+1)}\to\kappa \mid \boldsymbol\lambda^{(x+1,y)}, \boldsymbol\lambda^{(x,y+1)} \bigr). \end{equation*} \end{corollary} \section{A dynamic stochastic six vertex model} \label{sec:dynamicS6V} Here we consider the Markov projection of the Yang-Baxter field onto the column number zero. This produces a new dynamic version of the stochastic six vertex model. The original stochastic six vertex model was introduced in \cite{GwaSpohn1992}, and its asymptotic behavior was studied in various regimes in, e.g., \cite{BCG6V}, \cite{AmolBorodin2016Phase}, \cite{Amol2016Stationary}. We recall this model in \Cref{sub:degen_HL} below. \subsection{Dynamic vertex weights} \label{sub:dynamicS6V_subsection} Let $\boldsymbol\Lambda=\{\boldsymbol\lambda^{(x,y)}\}_{x,y\ge0}$ be the Yang-Baxter field constructed in \Cref{sec:YB_field}. Recall that each $\boldsymbol\lambda^{(x,y)}$ is a random nonnegative signature (of length $y$). For each $(x,y)\in \mathbb{Z}_{\ge0}^2$, let $\boldsymbol\ell^{(x,y)}:= (\boldsymbol\lambda^{(x,y)})^{[0]}\in \mathbb{Z}_{\ge0}$ denote the number of arrows in the zeroth column of the arrow configuration encoded by the signature $\boldsymbol\lambda^{(x,y)}$.\footnote{Equivalently, $(\boldsymbol\lambda^{(x,y)})^{[0]}$ is the number of zero parts in the signature $\boldsymbol\lambda^{(x,y)}$.} Since $\boldsymbol\lambda^{(x,y)}\in \mathsf{Sign}_y$, we have $\boldsymbol\ell^{(x,y)}\le y$. \Cref{prop:YB_Markov_projections} implies that the scalar random field $\mathbf{L}:=\{\boldsymbol\ell^{(x,y)}\}_{x,y\ge0}$ \emph{does not depend} on the rest of the Yang-Baxter field (i.e., of the numbers of arrows in $\boldsymbol\lambda^{(x,y)}$ in columns $\ge1$). In this way we say that $\mathbf{L}$ is a marginally Markovian projection of the Yang-Baxter field $\boldsymbol\Lambda$ onto the column number zero. Let us now present an independent description of $\mathbf{L}$. From the definition of the Yang-Baxter field via conditional probabilities \eqref{YB_field_definition_forward} it follows that for each $(x,y)\in \mathbb{Z}_{\ge0}^{2}$ the value of $\boldsymbol\ell^{(x+1,y+1)}$ is randomly determined using $\boldsymbol\ell^{(x+1,y)},\boldsymbol\ell^{(x,y)}$, and $\boldsymbol\ell^{(x,y+1)}$, and the corresponding conditional probabilities can be read from \eqref{U_0_dynamic_S6V_transitions}. In the language of values of the field $\mathbf{L}$ these conditional probabilities are given in \Cref{fig:L_dynamicS6V_probabilities}. The nature of the six possible configurations of the values of $\mathbf{L}$ at $2\times 2$ squares allow to \emph{identify} $\mathbf{L}$ with the height function in a dynamic version of the stochastic six vertex model. Let us describe this model in more detail. \begin{figure}[htpb] %\begin{noindent} \centering \begin{tabular}{c|c|c|c|c|c} \scalebox{.9}{\begin{tikzpicture} [scale=1.2, very thick] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell+1$}; \node[anchor=north west] at (.1,-.1) {$\ell-1$}; \node[anchor=south west] at (.1,.1) {$\ell$}; \draw[ultra thick,->] (-.6,0)--++(.55,0); \draw[ultra thick,->] (0,-.6)--++(0,.55); \draw[ultra thick,->] (0,0)--++(.55,0); \draw[ultra thick,->] (0,0)--++(0,.55); \end{tikzpicture}}& \scalebox{.9}{\begin{tikzpicture} [scale=1.2, very thick] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell+1$}; \node[anchor=north west] at (.1,-.1) {$\ell$}; \node[anchor=south west] at (.1,.1) {$\ell$}; \draw[ultra thick,->] (-.6,0)--++(.6,0); \draw[ultra thick,->] (0,0)--++(0,.6); \end{tikzpicture}}& \scalebox{.9}{\begin{tikzpicture} [scale=1.2, very thick] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell+1$}; \node[anchor=north west] at (.1,-.1) {$\ell$}; \node[anchor=south west] at (.1,.1) {$\ell+1$}; \draw[ultra thick,->] (-.6,0)--++(.6,0); \draw[ultra thick,->] (0,0)--++(.6,0); \end{tikzpicture}}& \scalebox{.9}{\begin{tikzpicture} [scale=1.2, very thick] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell$}; \node[anchor=north west] at (.1,-.1) {$\ell-1$}; \node[anchor=south west] at (.1,.1) {$\ell$}; \draw[ultra thick,->] (0,-.6)--++(0,.6); \draw[ultra thick,->] (0,0)--++(.6,0); \end{tikzpicture}}& \scalebox{.9}{\begin{tikzpicture} [scale=1.2, very thick] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell$}; \node[anchor=north west] at (.1,-.1) {$\ell-1$}; \node[anchor=south west] at (.1,.1) {$\ell-1$}; \draw[ultra thick,->] (0,-.6)--++(0,.6); \draw[ultra thick,->] (0,0)--++(0,.6); \end{tikzpicture}}& \scalebox{.9}{\begin{tikzpicture} [scale=1.2, very thick] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell$}; \node[anchor=north west] at (.1,-.1) {$\ell$}; \node[anchor=south west] at (.1,.1) {$\ell$}; \end{tikzpicture}} \\\hline \scalebox{.9}{1} & \scalebox{.9}{$\dfrac{(1-t)v}{v-t u}\dfrac{u-st^\ell}{v-st^\ell}$} & \scalebox{.9}{$\dfrac{v-u}{v-tu} \dfrac{v-st^{\ell+1}}{v-st^\ell}$} & \scalebox{.9}{$\dfrac{(1-t)u}{v-tu} \dfrac{v-st^{\ell}}{u-st^{\ell}}$} & \scalebox{.9}{$\dfrac{t(v-u)}{v-tu} \dfrac{u-st^{\ell-1}}{u-st^{\ell}}$} & \scalebox{.9}{1} \Bigg. \end{tabular} \caption{% Conditional probabilities in the random field $\mathbf{L}$ on $\mathbb{Z}_{\ge0}^2$. In the top row all possible values of the field in the square $\left\{ x,x+1 \right\}\times\left\{ y,y+1 \right\}$ are listed, where $\ell\in\mathbb{Z}_{\ge0}$ (and $\ell\ge1$ in the first, fourth, and fifth pictures). The bottom row contains the corresponding conditional probabilities to sample the top right value $\boldsymbol\ell^{(x+1,y+1)}$ of the field given the three other values. The spectral parameters are $v=v_{x+1}$ and $u=u_{y+1}$. The arrows represent identification with the six vertex configurations. } %\end{noindent} \label{fig:L_dynamicS6V_probabilities} \end{figure} First we define the space of configurations in our dynamic stochastic six vertex model. Consider an ensemble of infinite up-right paths in the positive integer quadrant with the following properties: \begin{itemize} \item Paths go along edges of the shifted lattice $\left( \mathbb{Z}_{\ge0}+\frac{1}{2} \right)^2$; \item Each edge of $\left( \mathbb{Z}_{\ge0}+\frac{1}{2} \right)^2$ is occupied by at most one path; \item Paths can touch each other at a vertex but cannot cross each other; \item On the boundary of the quadrant no paths enter from below, and at each height $n+\frac{1}{2}$, $n\ge0$, a new path enters through the left part of the boundary; \end{itemize} Fix such a configuration of up-right paths. At each $(x,y)$ in the original non-shifted lattice $\mathbb{Z}_{\ge0}^2$ define the value of the \emph{height function}, $\mathfrak{h}(x,y)$, to be the number of paths passing below $(x,y)$. See \Cref{fig:dyn_stoch6V} for an illustration. \begin{definition}[DS6V] \label{def:dyn_S6V} The \emph{dynamic stochastic six vertex model} (\emph{DS6V} for short) is a probability distribution on ensembles of up-right paths (depending on the parameters $t\in[0,1)$, $s\in(-1,0]$, and two sequences $v_1,v_2,\ldots $ and $u_1,u_2,\ldots $ such that $0\le u_i] (-.35,0)--++(6.85,0) node [right] {$x$}; \draw[->] (0,-.35)--++(0,4.85) node [left] {$y$}; \foreach \ii in {1,2,3,4,5,6} { \node at (\ii-.5,-1) {$v_\ii$}; \draw[dotted, thick] (\ii,-.3)--++(0,4.65); \node at (\ii,-.6) {$\ii$}; } \foreach \jj in {1,2,3,4} { \node at (-1,\jj-.5) {$u_\jj$}; \draw[dotted, thick] (-.3,\jj)--++(6.65,0); \node at (-.6,\jj) {$\jj$}; } \node at (-.6,0) {$0$}; \node at (0,-.6) {$0$}; \foreach \zz in {(0,0),(1,0),(2,0),(3,0),(4,0),(5,0),(6,0),(5,1),(6,1),(5,2),(6,2),(6,3)} {\node[rectangle,draw,fill=white] at \zz {0};} \foreach \zz in {(0,1),(1,1),(2,1),(3,1),(4,1),(3,2),(4,2),(5,3),(5,4),(6,4)} {\node[rectangle,draw,fill=white] at \zz {1};} \foreach \zz in {(0,2),(1,2),(2,2),(1,3),(2,3),(3,3),(4,3),(4,4)} {\node[rectangle,draw,fill=white] at \zz {2};} \foreach \zz in {(0,3),(1,4),(2,4),(3,4)} {\node[rectangle,draw,fill=white] at \zz {3};} \foreach \zz in {(0,4)} {\node[rectangle,draw,fill=white] at \zz {4};} \draw[line width=2.7,->] (-.65, .5)--(-.5,.5)--++(1,0)--++(1,0)--++(1,0)--++(1,0)--++(1,0)--++(0,1)--++(0,.92)--++(.08,.08)--++(.92,0)--++(0,1)--++(1,0); \draw[line width=2.7,->] (-.65,1.5)--(-.5,1.5)--++(1,0)--++(2,0)--++(0,1)--++(1,0)--++(0.92,0)--++(.08,.08)--++(0,.92)--++(0,1); \draw[line width=2.7,->] (-.65,2.5)--(-.5,2.5)--++(1,0)--++(0,.92)--++(.08,.08)--++(.92,0)--++(1,0)--++(1,0)--++(0,1); \draw[line width=2.7,->] (-.65,3.5)--(-.5,3.5)--++(0.92,0)--++(.08,.08)--++(0,.92); \end{tikzpicture} } \caption{Path configuration of six vertex type in a quadrant together with its height function.} %\end{noindent} \label{fig:dyn_stoch6V} \end{figure} \begin{remark} \label{rmk:dynS6V_not_the_same} The vertex model introduced in \Cref{def:dyn_S6V} differs from the dynamic stochastic six vertex model presented recently in \cite{borodin2017elliptic} as a degeneration of the stochastic Interaction-Round-a-Face model (introduced in the same work). A higher spin model following the approach of the latter paper was then developed in \cite{aggarwal2017dynamical}. All these dynamic stochastic vertex models are closely related to versions of the Yang-Baxter equation with dynamic parameters (see \Cref{sub:dynamic_YB} below for our dynamic Yang-Baxter exuation which seems to be simpler than the one in \cite{borodin2017elliptic}). Therefore, we regard the model from \Cref{def:dyn_S6V} as another dynamic version of the stochastic six vertex model, different from the ones in \cite{borodin2017elliptic}, \cite{aggarwal2017dynamical}. \end{remark} \begin{proposition} \label{prop:dyn6V_is_YB_field} Let $\mathfrak{H}:=\{\mathfrak{h}(x,y)\}_{x,y\ge0}$ be the random field of values of the height function of DS6V (\Cref{def:dyn_S6V}). Let $\mathbf{L}=\{\boldsymbol\ell^{(x,y)}\}_{x,y\ge0}$ be the random field obtained as the projection of the Yang-Baxter random field of \Cref{def:YB_field} onto the column number zero. Then these random fields $\mathfrak{H}$ and $\mathbf{L}$ have the same distribution. \end{proposition} \begin{proof} Straightforward from the identification of weights in $\mathfrak{H}$ and $\mathbf{L}$ in \Cref{fig:L_dynamicS6V_probabilities} together with the identification of the boundary conditions. \end{proof} From \Cref{thm:YB_field_spin_HL_process} and \Cref{prop:dyn6V_is_YB_field} we immediately get the following interpretation of the distribution of the height function in DS6V: \begin{corollary} \label{cor:dyn6V_spin_HL_process} Fix a down-right path $\mathcal{P}_{\vec{x},\vec{y}}$ as in \eqref{spin_HL_down_right_path_sequences}--\eqref{spin_HL_down_right_path}. The joint distribution of the random variables $\{\mathfrak{h}(p)\colon p\in\mathcal{P}_{\vec{x},\vec{y}}\}$ (i.e., the values of the height function of the dynamic stochastic six vertex model along this down-right path), coincides with the joint distribution of $\bigl\{(\lambda^{(p)})^{[0]}\colon p\in \mathcal{P}_{\vec{x},\vec{y}}\bigr\}$, the numbers of zero parts in the signatures $\lambda^{(p)}$ governed by the spin Hall-Littlewood process $\mathscr{HP}_{\vec{x},\vec{y}}$ corresponding to the down-right path $\mathcal{P}_{\vec{x},\vec{y}}$. \end{corollary} \subsection{A dynamic Yang-Baxter equation} \label{sub:dynamic_YB} The probabilities of vertex configurations in DS6V (given in \Cref{fig:L_dynamicS6V_probabilities}) satisfy a dynamic version of the Yang-Baxter equation. It is convenient to formulate it in terms of the values of the height function since the corresponding arrow configurations can be readily recovered as in \Cref{fig:L_dynamicS6V_probabilities}. Consider two three-line configurations as in \Cref{fig:dyn_YBE}. Fix the six boundary values $\ell_0,\ell_1,\ell_2,\ell_1',\ell_2',\ell_3\in \mathbb{Z}_{\ge0}$ of the height function. Clearly, these values can be arbitrary provided that they satisfy \begin{equation} \label{conditions_on_ell} \ell_1-\ell_0,\ell_2-\ell_1,\ell_3-\ell_2 \in \left\{ 0,1 \right\}, \qquad \ell_1'-\ell_0,\ell_2'-\ell_1',\ell_3-\ell_2'\in \left\{ 0,1 \right\}. \end{equation} Also fix spectral parameters $\mathsf{u}_1,\mathsf{u}_2,\mathsf{v}$. For the dynamic Yang-Baxter equation in \Cref{thm:dynamic_YB} below these parameters do not have to satisfy any conditions as in \Cref{def:dyn_S6V}. However, if $0\le \mathsf{u}_2<\mathsf{u}_1<\mathsf{v}<1$ and $0\le t<1$, $-1, line width=2] (0,1)--(0.882353,.5); \draw[->, line width=2] (0.882353,.5)--(2.2,1.24667); \draw[->, line width=2] (2.2,1.24667)--(3,1.7); \node at (2.6,-1) {$\ell$}; \node at (1.3,-.2) {$\ell$}; \node at (0,.5) {$\ell$}; \node at (2.6,0.5) {$\ell$}; \node at (2.65,2) {$\ell+1$}; \node at (1.3,1.3) {$\ell+1$}; \node at (1.75,.5) {$\ell$}; \end{tikzpicture} }\right] + \left[\scalebox{.7}{ \begin{tikzpicture}[scale=1.1, thick, baseline=12pt] \draw[densely dashed] (0,0) -- (3,1.7); \draw[densely dashed] (0,1) -- (3,-.7); \draw[densely dashed] (2.2,-1.2)--++(0,3.4); \draw[->, line width=2] (0,1)--(0.882353,.5); \draw[->, line width=2] (0.882353,.5)--(2.2,-0.246667); \draw[->, line width=2] (2.2,-0.246667)--(2.2,1.24667); \draw[->, line width=2] (2.2,1.24667)--(3,1.7); \node at (2.6,-1) {$\ell$}; \node at (1.3,-.2) {$\ell$}; \node at (0,.5) {$\ell$}; \node at (2.6,0.5) {$\ell$}; \node at (2.65,2) {$\ell+1$}; \node at (1.3,1.3) {$\ell+1$}; \node at (1.75,.5) {$\ell+1$}; \end{tikzpicture} }\right] = \left[\scalebox{.7}{ \begin{tikzpicture}[scale=1.1, thick,baseline=12pt] \draw[densely dashed] (0,1.7) -- (3,0); \draw[densely dashed] (0,-.7) -- (3,1); \draw[densely dashed] (.8,-1.2)--++(0,3.4); \draw[->, line width=2] (0,1.7)--(0.8,1.246667); \draw[->, line width=2] (.8,1.24667)--(2.11765,.5); \draw[->, line width=2] (2.11765,.5)--(3,1); \node at (1.8,-.2) {$\ell$}; \node at (.35,-1) {$\ell$}; \node at (.35,.5) {$\ell$}; \node at (3,0.5) {$\ell$}; \node at (1.8,1.3) {$\ell+1$}; \node at (.4,2) {$\ell+1$}; \node at (1.25,.5) {$\ell$}; \end{tikzpicture} }\right]. \end{equation*} This translates into the following identity between rational functions \begin{multline*} \tfrac{(1-t) \mathsf{u}_1 (\mathsf{u}_2-s t^\ell)}{(\mathsf{u}_1-t \mathsf{u}_2) (\mathsf{u}_1-s t^\ell)} \tfrac{(\mathsf{v}-\mathsf{u}_1) (s t^{\ell+1}-\mathsf{v})}{(\mathsf{v}-t \mathsf{u}_1) (s t^\ell-\mathsf{v})} + \tfrac{(\mathsf{u}_1-\mathsf{u}_2) (\mathsf{u}_1-s t^{\ell+1})}{(\mathsf{u}_1-t \mathsf{u}_2) (\mathsf{u}_1-s t^\ell)} \tfrac{(1-t) \mathsf{v} (\mathsf{u}_2-s t^\ell)}{(\mathsf{v}-t \mathsf{u}_2) (\mathsf{v}-s t^\ell)} \tfrac{(1-t) \mathsf{u}_1 (\mathsf{v}-s t^{\ell+1})}{(\mathsf{v}-t \mathsf{u}_1) (\mathsf{u}_1-s t^{\ell+1})} \\= \tfrac{(\mathsf{v}-\mathsf{u}_2) (s t^{\ell+1}-\mathsf{v})}{(\mathsf{v}-t \mathsf{u}_2) (s t^\ell-\mathsf{v})} \tfrac{(1-t) \mathsf{u}_1 (\mathsf{u}_2-s t^\ell)}{(\mathsf{u}_1-t \mathsf{u}_2) (\mathsf{u}_1-s t^\ell)}, \end{multline*} which is readily verified by hand. The remaining 19 identities comprising the dynamic Yang-Baxter equation are checked in a similar way, and the theorem follows. \end{proof} \begin{remark} \label{rmk:dyn_YB_from_static} The dynamic Yang-Baxter equation of \Cref{thm:dynamic_YB} can in fact be reduced to the usual Yang-Baxter equation for the stochastic six vertex model, but we do not use this fact here. \end{remark} The dynamic Yang-Baxter equation of \Cref{thm:dynamic_YB} satisfied by the probabilities in the dynamic stochastic six vertex model hints at the model's integrability (i.e., that certain observables of this model are computable in explicit form). We do not discuss these problems in the present work, though in \Cref{sec:degenerations} below we consider degenerations of DS6V for which certain observables indeed can be computed in explicit form. \appendix \section{Degenerations and limits} \label{sec:degenerations} %\begin{noindent} Here we discuss a number of degenerations of the dynamic stochastic six vertex model (DS6V) and its properties stated in \Cref{thm:YB_field_spin_HL_process} and \Cref{cor:dyn6V_spin_HL_process}. Some of these degenerations correspond to degenerations of the spin Hall-Littlewood symmetric functions outlined in \cite[Section 8]{Borodin2014vertex}. The tables in \Cref{fig:dynamic_vertex_weights_degen,fig:dynamic_vertex_weights_degen_half_cont} list various degenerations of the DS6V weights considered in Appendices \ref{sub:degen_HL} to \ref{sub:hc_degen_t0s0}. Additional (less direct) degenerations are discussed in \Cref{sub:degen_rational,sub:degen_ASEP,sub:degen_finite_spin}. We also discuss two degenerations of the full Yang-Baxter field in \Cref{sub:degen_HL,sub:hc_degen_Schur}, and compare them to known systems. \begin{remark} \label{rmk:first_k_columns} Every degeneration of the DS6V model we consider can be lifted to a $k$-layer model, where $k\ge2$ is arbitrary. Indeed, such a model would arise by taking the corresponding degeneration of the full Yang-Baxter field, and looking at its Markov projection onto the first $k$ columns as in \Cref{sub:properties_of_global_transitions}. Such multilayer models for $s=0$ were explicitly written down in \cite{BufetovMatveev2017}. For shortness, we will not address multilayer extensions in the present work. \end{remark} For simplicity we assume that the spectral parameters are constant, $u_i\equiv u$ and $v_j\equiv v$, but most constructions (except the ASEP type limit in \Cref{sub:degen_ASEP}) work for the inhomogeneous parameters $u_i,v_j$, too. %\end{noindent} \begin{figure}[htpb] %\begin{noindent} \centering \begin{tabular}{c|c|c|c|c|c} && \scalebox{.9}{\begin{tikzpicture} [scale=1.2, very thick] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell+1$}; \node[anchor=north west] at (.1,-.1) {$\ell$}; \node[anchor=south west] at (.1,.1) {$\ell$}; \draw[ultra thick,->] (-.6,0)--++(.6,0); \draw[ultra thick,->] (0,0)--++(0,.6); \end{tikzpicture}}& \scalebox{.9}{\begin{tikzpicture} [scale=1.2, very thick] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell+1$}; \node[anchor=north west] at (.1,-.1) {$\ell$}; \node[anchor=south west] at (.1,.1) {$\ell+1$}; \draw[ultra thick,->] (-.6,0)--++(.6,0); \draw[ultra thick,->] (0,0)--++(.6,0); \end{tikzpicture}}& \scalebox{.9}{\begin{tikzpicture} [scale=1.2, very thick] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell$}; \node[anchor=north west] at (.1,-.1) {$\ell-1$}; \node[anchor=south west] at (.1,.1) {$\ell$}; \draw[ultra thick,->] (0,-.6)--++(0,.6); \draw[ultra thick,->] (0,0)--++(.6,0); \end{tikzpicture}}& \scalebox{.9}{\begin{tikzpicture} [scale=1.2, very thick] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell$}; \node[anchor=north west] at (.1,-.1) {$\ell-1$}; \node[anchor=south west] at (.1,.1) {$\ell-1$}; \draw[ultra thick,->] (0,-.6)--++(0,.6); \draw[ultra thick,->] (0,0)--++(0,.6); \end{tikzpicture}} \\\hline &\scalebox{.9}{\parbox{.1\textwidth}{Original weights}} & \scalebox{.9}{$\dfrac{(1-t)v}{v-t u}\dfrac{u-st^\ell}{v-st^\ell}$} & \scalebox{.9}{$\dfrac{v-u}{v-tu} \dfrac{v-st^{\ell+1}}{v-st^\ell}$} & \scalebox{.9}{$\dfrac{(1-t)u}{v-tu} \dfrac{v-st^{\ell}}{u-st^{\ell}}$} & \scalebox{.9}{$\dfrac{t(v-u)}{v-tu} \dfrac{u-st^{\ell-1}}{u-st^{\ell}}$} \bigg. \\\hline (a)&\scalebox{.9}{\Cref{sub:degen_HL}} & \scalebox{.9}{$\dfrac{(1-t)u/v}{1-tu/v}$} & \scalebox{.9}{$\dfrac{1-u/v}{1-tu/v}$} & \scalebox{.9}{$\dfrac{1-t}{1-tu/v}$} & \scalebox{.9}{$\dfrac{t(1-u/v)}{1-tu/v}$} \bigg. \\\hline (b)&\scalebox{.9}{\Cref{sub:degen_t0}} & \scalebox{.9}{$\dfrac{u-s\mathbf{1}_{\ell=0}}{v-s\mathbf{1}_{\ell=0}}$} & \scalebox{.9}{$ \dfrac{v-u}{v-s\mathbf{1}_{\ell=0}}$} & \scalebox{.9}{$1$} & \scalebox{.9}{$0$} \bigg. \\\hline (c)&\scalebox{.9}{\Cref{sub:degen_Schur}} & \scalebox{.9}{$u/v$} & \scalebox{.9}{$1-u/v$} & \scalebox{.9}{$1$} & \scalebox{.9}{$0$} \big. \\\hline (d)&\scalebox{.9}{\Cref{sub:degen_IHL}} & \scalebox{.9}{$\dfrac{(1-t)v}{v-t u}\dfrac{u+t^\ell}{v+t^\ell}$} & \scalebox{.9}{$\dfrac{v-u}{v-tu} \dfrac{v+t^{\ell+1}}{v+t^\ell}$} & \scalebox{.9}{$\dfrac{(1-t)u}{v-tu} \dfrac{v+t^{\ell}}{u+t^{\ell}}$} & \scalebox{.9}{$\dfrac{t(v-u)}{v-tu} \dfrac{u+t^{\ell-1}}{u+t^{\ell}}$} \bigg. \\\hline (e)&\scalebox{.9}{\Cref{sub:degen_t0s0}} & \scalebox{.9}{$\dfrac{u+\mathbf{1}_{\ell=0}}{v+\mathbf{1}_{\ell=0}}$} & \scalebox{.9}{$\dfrac{v-u}{v+\mathbf{1}_{\ell=0}}$} & \scalebox{.9}{$1$} & \scalebox{.9}{$0$} \bigg. \end{tabular} \caption{% Direct degenerations of the dynamic stochastic six vertex weights from \Cref{sec:dynamicS6V} considered in the first part of \Cref{sec:degenerations}. Here $\ell\in \mathbb{Z}_{\ge0}$ (and $\ell\ge1$ in the last two cases) is the parameter corresponding to the height function, and $\mathbf{1}_{\cdots}$ denotes the indicator of an event. The vertices $(1,1;1,1)$ and $(0,0;0,0)$ always having weight $1$ are not shown.% } \label{fig:dynamic_vertex_weights_degen} %\end{noindent} \end{figure} \subsection{Hall-Littlewood degeneration and stochastic six vertex model} \label{sub:degen_HL} Setting $s=0$ and keeping other parameters makes the DS6V weights independent of the height function. Moreover, in this degeneration the weights depend only on the ratio $u/v$ and not on the individual parameters $u,v$. See \Cref{fig:dynamic_vertex_weights_degen}(a). Thus, in this limit the DS6V turns into the usual stochastic six vertex model introduced in \cite{GwaSpohn1992} and studied in Integrable Probability since \cite{BCG6V}. The spin Hall-Littlewood symmetric functions $F$ and $G^c$ turn (up to simple factors) into the Hall-Littlewood symmetric polynomials \cite[Ch. III]{Macdonald1995}. The correspondence between the stochastic six vertex model and Hall-Littlewood processes following from \Cref{cor:dyn6V_spin_HL_process} was obtained earlier in \cite{borodin2016stochastic_MM} (at the level of formulas), \cite{BorodinBufetovWheeler2016} (for a half-continuous degeneration, cf. \Cref{sub:hc_degen_HL} below), and in full form in \cite{BufetovMatveev2017}. The Yang-Baxter field $\boldsymbol \Lambda:=\{ \boldsymbol\lambda^{(x,y)} \}_{x,y\ge0}$ for $s=0$ becomes a certain field of random Young diagrams indexed by $\mathbb{Z}_{\ge0}^{2}$ related to Hall-Littlewood measures and processes. This random field \emph{differs} from the Hall-Littlewood RSK field introduced in \cite{BufetovMatveev2017}, despite that: \begin{itemize} \item In both fields, joint distributions along down-right paths are the same and are given by the Hall-Littlewood processes as in \Cref{cor:dyn6V_spin_HL_process}. \item The projection onto the first column in both fields produces the stochastic six vertex model. \end{itemize} The existence of two different random fields with these properties might seem surprising, but such non-uniqueness of 2-dimensional stochastic dynamics was observed before, e.g., in \cite{BorodinPetrov2013NN} or \cite[Section 4]{BorodinPetrov2013Lect}. The fact that the $s=0$ Yang-Baxter field and the Hall-Littlewood RSK field are indeed different will be evident in \Cref{sub:hc_degen_Schur} when we take further degenerations and obtain different objects. \begin{remark} \label{rmk:spinHL_not_RSK} The Hall-Littlewood RSK field of \cite{BufetovMatveev2017} has an additional structure coming from the fact that the skew Hall-Littlewood symmetric functions in one variable are proportional to a power of the variable. Using this fact, analogues of the probabilities $\mathsf{U}^{\mathrm{fwd}}_{v,u}$ and $\mathsf{U}^{\mathrm{bwd}}_{v,u}$ for the Hall-Littlewood RSK field lead to \emph{randomized RSK correspondences}: having Young diagrams $\mu,\kappa,\lambda$, and an integer $r\in \mathbb{Z}_{\ge0}$ (corresponding to the power of $u/v$), the randomized RSK produces a random output Young diagram $\nu$. See \cite[Section 3.6]{BufetovMatveev2017} for details on this reduction of a random field of Young diagrams to randomized RSK correspondences with input. However, for $s\ne 0$ the skew spin Hall-Littlewood symmetric functions in one variable are not simply proportional to powers of the variables. This presents a clear obstacle to a possible reduction of the Yang-Baxter field or another such random field of signatures to a randomized correspondence with integer input. Therefore, we do not address this issue in the present work. \end{remark} Observables of Hall-Littlewood processes pertaining to the projection onto first columns can be extracted using the action of Hall-Littlewood versions of Macdonald difference operators (e.g., see \cite{dimitrov2016kpz}). Thus, the connection between the stochastic six vertex model and Hall-Littlewood processes produces tools for the analysis of the former model alternative to the original approach of \cite{BCG6V}. See, e.g., \cite{borodin2016stochastic_MM} for an analysis via Hall-Littlewood measures. \subsection{Schur degeneration and modified discrete time PushTASEP} \label{sub:degen_t0} Setting $t=0$ and keeping all other parameters makes the DS6V weights look as in \Cref{fig:dynamic_vertex_weights_degen}(b). These weights are still dynamic in the sense that they retain dependence on the height function. However, this dependence only singles out the bottommost path: the behavior of all other paths follows the same weights. As noted in \cite[Section 8.3]{Borodin2014vertex}, the spin Hall-Littlewood functions $F$ and $G^c$ for $t=0$ turn into certain determinants generalizing Schur polynomials, thus making the spin Hall-Littlewood measures and processes in this degeneration potentially more tractable. Let us reinterpret the $t=0$ degeneration of DS6V as a discrete time particle system by regarding the horizontal direction as time (a similar interpretation is valid for the general DS6V model, too, only the corresponding particle system becomes more complicated.) \begin{definition} \label{def:discrete_modified_push} Consider a discrete time particle system living on half infinite particle configurations $x_1(\mathsf{t})0$, as it should be. Indeed, after time $\mathsf{t}] (-.6,0)--++(.6,0); \draw[ultra thick,->] (0,0)--++(0,.6); \node at (0,-.9) {Rate}; \end{tikzpicture} } & \scalebox{.9}{ \begin{tikzpicture}[scale=1.2, very thick] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell+1$}; \node[anchor=north west] at (.1,-.1) {$\ell$}; \node[anchor=south west] at (.1,.1) {$\ell+1$}; \draw[ultra thick,->] (-.6,0)--++(.6,0); \draw[ultra thick,->] (0,0)--++(.6,0); \node at (0,-.9) {Probability}; \end{tikzpicture} } & \scalebox{.9}{ \begin{tikzpicture}[scale=1.2, very thick] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell$}; \node[anchor=north west] at (.1,-.1) {$\ell-1$}; \node[anchor=south west] at (.1,.1) {$\ell$}; \draw[ultra thick,->] (0,-.6)--++(0,.6); \draw[ultra thick,->] (0,0)--++(.6,0); \node at (0,-.9) {Probability}; \end{tikzpicture} } & \scalebox{.9}{ \begin{tikzpicture}[scale=1.2, very thick] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell$}; \node[anchor=north west] at (.1,-.1) {$\ell-1$}; \node[anchor=south west] at (.1,.1) {$\ell-1$}; \draw[ultra thick,->] (0,-.6)--++(0,.6); \draw[ultra thick,->] (0,0)--++(0,.6); \node at (0,-.9) {Probability}; \end{tikzpicture} } \\\hline (a) & \scalebox{.9}{\Cref{sub:degen_half_continuous_DS6V}} & \scalebox{.9}{$(1-t)(u-st^\ell)$} & \scalebox{.9}{$1-O(v^{-1})$} & \scalebox{.9}{$\dfrac{(1-t)u}{u-st^\ell}$} & \scalebox{.9}{$\dfrac{tu-st^\ell}{u-st^\ell}$} \bigg. \\\hline (b) & \scalebox{.9}{\Cref{sub:hc_degen_HL}} & \scalebox{.9}{$(1-t)u$} & \scalebox{.9}{$1-O(v^{-1})$} & \scalebox{.9}{$1-t$} & \scalebox{.9}{$t$} \bigg. \\\hline (c) & \scalebox{.9}{\Cref{sub:hc_degen_t0}} & \scalebox{.9}{$u-s\mathbf{1}_{\ell=0}$} & \scalebox{.9}{$1-O(v^{-1})$} & \scalebox{.9}{$1$} & \scalebox{.9}{$0$} \bigg. \\\hline (d) & \scalebox{.9}{\Cref{sub:hc_degen_Schur}} & \scalebox{.9}{$u$} & \scalebox{.9}{$1-O(v^{-1})$} & \scalebox{.9}{$1$} & \scalebox{.9}{$0$} \bigg. \\\hline (e) & \scalebox{.9}{\Cref{sub:hc_degen_IHL}} & \scalebox{.9}{$(1-t)(u+t^\ell)$} & \scalebox{.9}{$1-O(v^{-1})$} & \scalebox{.9}{$\dfrac{(1-t)u}{u+t^\ell}$} & \scalebox{.9}{$\dfrac{tu-st^\ell}{u+t^\ell}$} \bigg. \\\hline (f) & \scalebox{.9}{\Cref{sub:hc_degen_t0s0}} & \scalebox{.9}{$u+\mathbf{1}_{\ell=0}$} & \scalebox{.9}{$1-O(v^{-1})$} & \scalebox{.9}{$1$} & \scalebox{.9}{$0$} \bigg. \end{tabular} \caption{The half-continuous DS6V model and its various degenerations. The vertices $(1,1;1,1)$ and $(0,0;0,0)$ always having weight $1$ are not shown.} \label{fig:dynamic_vertex_weights_degen_half_cont} \end{figure} Taking the expansion as $v\to+\infty$ of the DS6V vertex weights in \Cref{fig:dyn_stoch6V}, we see that \begin{align*} \Biggl[ \scalebox{.8}{ \begin{tikzpicture}[scale=1.2, very thick, baseline=-3pt] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell+1$}; \node[anchor=north west] at (.1,-.1) {$\ell$}; \node[anchor=south west] at (.1,.1) {$\ell$}; \draw[ultra thick,->] (-.6,0)--++(.6,0); \draw[ultra thick,->] (0,0)--++(0,.6); \end{tikzpicture} } \Biggr] & = v^{-1}(1-t)(u-st^{\ell})+O(v^{-2}) , \quad \Biggl[ \scalebox{.8}{ \begin{tikzpicture}[scale=1.2, very thick, baseline=-3pt] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell+1$}; \node[anchor=north west] at (.1,-.1) {$\ell$}; \node[anchor=south west] at (.1,.1) {$\ell+1$}; \draw[ultra thick,->] (-.6,0)--++(.6,0); \draw[ultra thick,->] (0,0)--++(.6,0); \end{tikzpicture} } \Biggr] = 1-O(v^{-1}) , \\ \Biggl[\scalebox{.8}{ \begin{tikzpicture}[scale=1.2, very thick, baseline=-3pt] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell$}; \node[anchor=north west] at (.1,-.1) {$\ell-1$}; \node[anchor=south west] at (.1,.1) {$\ell$}; \draw[ultra thick,->] (0,-.6)--++(0,.6); \draw[ultra thick,->] (0,0)--++(.6,0); \end{tikzpicture} } \Biggr] & = \frac{(1-t)u}{u-st^{\ell}}+O(v^{-2}) ,\qquad \qquad \qquad \Biggl[\scalebox{.8}{ \begin{tikzpicture}[scale=1.2, very thick, baseline=-3pt] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell$}; \node[anchor=north west] at (.1,-.1) {$\ell-1$}; \node[anchor=south west] at (.1,.1) {$\ell-1$}; \draw[ultra thick,->] (0,-.6)--++(0,.6); \draw[ultra thick,->] (0,0)--++(0,.6); \end{tikzpicture} } \Biggr] = \frac{tu-st^{\ell}}{u-st^{\ell}}+O(v^{-2}) , \\&\hspace{50pt} \Biggl[ \scalebox{.8}{ \begin{tikzpicture}[scale=1.2, very thick, baseline=-3pt] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell+1$}; \node[anchor=north west] at (.1,-.1) {$\ell-1$}; \node[anchor=south west] at (.1,.1) {$\ell$}; \draw[ultra thick,->] (-.6,0)--++(.55,0); \draw[ultra thick,->] (0,-.6)--++(0,.55); \draw[ultra thick,->] (0,0)--++(.55,0); \draw[ultra thick,->] (0,0)--++(0,.55); \end{tikzpicture} } \Biggr] = \Biggl[ \scalebox{.8}{ \begin{tikzpicture}[scale=1.2, very thick, baseline=-3pt] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell$}; \node[anchor=north west] at (.1,-.1) {$\ell$}; \node[anchor=south west] at (.1,.1) {$\ell$}; \end{tikzpicture} } \Biggr] =1. \end{align*} Thus, for $v\gg1$, taking into account the DS6V boundary conditions, we see that all up-right paths will go to the right most of the steps. Occasionally with probability proportional to $v^{-1}$, a path might turn up using the vertex $(0,1;1,0)$, move some random distance up using several vertices $(1,0;1,0)$, and either turn right using $(1,0;0,1)$, or hit a neighboring path above it using $(1,1;1,1)$ (recall that paths can touch each other at a vertex but cannot cross each other). In the latter case, this neighboring path now faces up, and in turn should make a number of upward steps and either eventually turn right, or hit the next path, and so on. The update in this vertical slice eventually terminates after some path decides to turn right, or after the infinite densely packed cluster of paths is pushed up by one. In the limit as $v\to+\infty$ we thus obtain a probability distribution on up-right paths in the half-continuous quadrant $\mathbb{R}_{\ge0}\times (\mathbb{Z}_{\ge0}+\tfrac12)$. All paths enter through the left boundary, and nothing enters from below. Each $\ell$-th path from below, $\ell\in \mathbb{Z}_{\ge1}$, carries an independent Poisson process of rate $(1-t)(u-st^{\ell-1})$. Outside arrivals of these Poisson processes\footnote{To rigorously define the system note that the behavior of the paths up to vertical coordinate $M$ does not depend on the behavior of the system above $M$, for any $M\ge1$. Thus, the evolution of any finite part of the system with vertical coordinate $\le M$ is well-defined, and for different $M$ these processes are compatible, thus defining the measure on the full half-continuous quadrant.} all paths go to the right. When there is an arrival in the $\ell$-th Poisson process, the corresponding path turns up, and then behaves as explained in the previous paragraph using probabilities of the vertices $(1,0;1,0)$, $(1,0;0,1)$, and $(1,1;1,1)$. Similarly to \Cref{def:discrete_modified_push}, one can interpret this half-continuous DS6V model as a continuous time particle system $x_1(\tau)\mu_{i+m+1}\quad \text{for some $m\ge0$}. \end{equation} \end{itemize} When $\lambda\ne \kappa$ and the difference is only in $\lambda_i=\kappa_i+1$, the transition probability $U(\mu\to\nu\mid \kappa\to\lambda)$ is in general equal to $\mathbf{1}_{\nu=\mu}$ (no move propagation), except: \begin{itemize} \item (mandatory pushing to restore interlacing) If $\lambda_i=\mu_i+1=\kappa_i+1$, this leads to $\nu_i=\mu_i+1$ with probability $1$; \item (special pushing) If \begin{equation} \label{YB_Schur_cont_field_exotic_push} \mu_i=\lambda_i=\mu_{i-1}=\lambda_{i-1} =\ldots=\mu_{i-m}=\lambda_{i-m}<\lambda _{i-m-1}\quad\text{for some $m\ge0$}, \end{equation} then together with $\lambda_i=\kappa_i+1$ this leads to $\nu_{i-m}=\mu_{i-m}+1$ with probability $1$. \end{itemize} In particular, in this dynamics the difference between $\lambda$ and $\kappa$, as well as between $\nu$ and $\mu$, is in the move of at most one particle to the right by one (in the language of Young diagrams, in adding one box). \end{definition} \begin{proposition} \label{prop:YB_Schur_cont_field} The half-continuous $t=s=0$ Yang-Baxter field is identified with the dynamics in \Cref{def:Schur_dyn_YB}. \end{proposition} \begin{proof} The Yang-Baxter field is determined using the forward transition probabilities $\mathsf{U}^{\mathrm{fwd}}_{v,u}$, which in turn are products of the local probabilities $P^{\mathrm{fwd}}_{v,u}$, cf. \Cref{def:Ufwd,def:YB_field}. Setting $s=t=0$ and expanding the latter as $v\to+\infty$ we get the quantities given in the table in \Cref{fig:fwd_Schur_cont}. Note that these quantities do not depend on the multiplicity $g$ of arrows in the middle as was the case for $s,t\ne 0$. Because of this, we can assume without loss of generality that all the multiplicities in the middle are $0$ or $1$. It remains to match the corresponding expansions as $v\to+\infty$ of $\mathsf{U}^{\mathrm{fwd}}_{v,u}$ to rates $w_j$ and update probabilities $U(\mu\to\nu\mid \kappa\to\lambda)$ given in \Cref{def:Schur_dyn_YB}. We do this in two steps, for $\lambda=\kappa$ (considering jump rates) and $\lambda\ne\kappa$ (dealing with move propagation). \begin{figure}[htpb] %\begin{noindent} \centering \scalebox{.9}{ $ \begin{array}{c||c||c|c||c|c||c} P^{\mathrm{fwd}}_{v,u}& \ooYBoo{.3}{3}{13.5pt} & \ioYBio{.3}{3}{13.5pt} & \oiYBio{.3}{3}{13.5pt} & \oiYBoi{.3}{3}{13.5pt} & \ioYBoi{.3}{3}{13.5pt} & \iiYBii{.3}{3}{13.5pt} \phantom{\bigg.} \\\hline \ooYBoo{.3}{3}{13.5pt} \phantom{\Big.} & 1 & v^{-1}u & \coli 1-O(v^{-1}) &1 &\colx 0 & 1 \\\hline \ioYBio{.3}{3}{13.5pt} \phantom{\Big.} & 1 & v^{-1}u & \coli 1-O(v^{-1}) & 1 & \colx0 & 1 \\\hline \oiYBio{.3}{3}{13.5pt} \phantom{\Big.} & \colx1 & \colx0 & 1 & \colx 1 & \colxx 0 & \colx 1 \\\hline \oiYBoi{.3}{3}{13.5pt} \phantom{\Big.} & 1 & 1 & \coli0 & 1 & \colx0 & 1 \\\hline \ioYBoi{.3}{3}{13.5pt} \phantom{\Big.} & \coli1 & \coli0 & \colii1 & \coli0 & 1 & \coli1 \\\hline \iiYBii{.3}{3}{13.5pt} \phantom{\Big.} & 1 & v^{-1}u & \coli 1-O(v^{-1}) & 1 & \colx 0 & 1 \end{array} $ } \caption{Behavior of the forward local Yang-Baxter transition probabilities for $s=t=0$ as $v\to+\infty$. The coloring of the table cells is explained in \Cref{fig:fwd_YB}. Note that the parameters $u,v$ are swapped compared to \Cref{fig:fwd_YB}, cf. \Cref{rmk:swap_u_v}.} \label{fig:fwd_Schur_cont} %\end{noindent} \end{figure} \smallskip\noindent \textit{Jump rates.} First consider the case $\lambda=\kappa$. Then the arrow configuration $\lambda\mathop{\dot\succ}\kappa\prec \mu$ (cf. \Cref{fig:skew_Cauchy,fig:U_transition_probabilities}) looks as in \Cref{fig:Schur_YB_proof}(a), and we need to drag the cross vertex through this configuration from left to right. The nonnegative integer line $\mathbb{Z}_{\ge0}$ is divided into segments of two types: type~I segments $[\lambda_i,\mu_i)$ and type~II segments $[\mu_{i+1},\lambda_i)$. When $\lambda_i,\mu_i$ are sufficiently apart, these types of segments interlace, but it can also happen that segments of the same type can be neighbors. The cross vertex starts in state $\iiYBii{.25}{3.5}{10.5pt}$ in type~I segment, and this state cannot change thoughout type~I segment. Observe that on the boundary from type~I to type~II segment (say, corresponding to the arrow at $\mu_i$), if the length of type~II segment is positive, the cross vertex transforms (while moving to the right) as: \begin{itemize} \item $\iiYBii{.25}{3.5}{10.5pt} \rightsquigarrow\ioYBio{.25}{3.5}{10.5pt} \rightsquigarrow\oiYBio{.25}{3.5}{10.5pt}$ with probability $v^{-1}u+O(v^{-2})$ (i.e., at rate $u$) if the length of the type~II segment is greater than $1$, or if the length of the type~II segment is $1$ and the following type~I segment has zero length; \item $\iiYBii{.25}{3.5}{10.5pt} \rightsquigarrow\ioYBio{.25}{3.5}{10.5pt} \rightsquigarrow\iiYBii{.25}{3.5}{10.5pt}$ with probability $v^{-1}u+O(v^{-2})$ (i.e., at rate $u$) if the length of the type~II segment is equal to $1$ and the next segment is type~I of positive length; \item $\iiYBii{.25}{3.5}{10.5pt} \rightsquigarrow\oiYBio{.25}{3.5}{10.5pt}$ with probability $1-O(v^{-1})$. \end{itemize} In the first two cases this move places an arrow in the middle at $\nu_i=\mu_i+1$, and in the second case an arrow is placed at $\nu_i=\mu_i$. As $v\to+\infty$, the move leading to $\nu_i=\mu_i+1$ can occur only once in the process of dragging the cross vertex, which proves the claim that in general the rates $w_i$ are equal to $u$. The cross vertex does not change throughout type~II segments and leaves such a segment as $\oiYBio{.25}{3.5}{10.5pt}$ (unless event of probability $v^{-1}u$ occurs and the length of type~II segment is $1$, but this is already considered above). When entering type~I segment of positive length (at, say, the boundary corresponding to $\lambda_j=\kappa_j$), the cross vertex transforms as $\oiYBio{.25}{3.5}{10.5pt} \rightsquigarrow\iiYBii{.25}{3.5}{10.5pt}$, and this removes an arrow in the middle at $\kappa_j$. A type~II segment of zero length corresponds to $\mu_{i+1}=\kappa_i=\lambda_i$ for some $i$, which blocks the independent jump of $\mu_{i+1}$. A type~I segment of zero length does not change the state of the cross from $\oiYBio{.25}{3.5}{10.5pt}$ which it has while traveling through type~II segment; this behavior corresponds to the special case \eqref{YB_Schur_cont_field_exotic_jump} in which the jump rate is zero. This establishes the claim about the jump rates. \begin{figure}[htpb] %\begin{noindent} \centering \begin{tikzpicture} [scale=.9] \def\h{.7} \draw[dashed] (0,0)--++(8.5,0); \draw[dashed] (0,\h)--++(8.5,0); \draw[->, ultra thick] (0,0)--++(2,0)--++(0,\h)--++(1,0)--++(0,.5); \draw[->, ultra thick] (0,\h)--++(1,0)--++(0,.5); \draw[->, ultra thick] (2,-.5)--++(0,.5)--++(1.5,0)--++(0,\h+.5); \draw[->, ultra thick] (3.5,-.5)--++(0,.5)--++(1,0)--++(0,\h)--++(1,0)--++(0,.5); \draw[->, ultra thick] (4.5,-.5)--++(0,.5)--++(1,0)--++(0,\h)--++(2,0)--++(0,.5); \draw[->, ultra thick] (5.5,-.5)--++(0,.5)--++(3,0); \node at (-.2,-.3) {$\lambda$}; \node at (-.2,\h/2) {$\kappa$}; \node at (-.2,\h+.3) {$\mu$}; \node at (8,-.4) {(a)}; \node at (.5,\h+.7) {I}; \node at (1.5,\h+.7) {II}; \node at (2.5,\h+.7) {I}; \node at (3.25,\h+.7) {II}; \node at (4,\h+.7) {II}; \node at (5,\h+.7) {I}; \node at (6.5,\h+.7) {I}; \node at (8,\h+.7) {II}; \end{tikzpicture} \\\vspace{10pt} \begin{tikzpicture} [scale=1] \def\h{.7} \draw[dashed] (0,0)--++(3,0); \draw[dashed] (0,\h)--++(3,0); \draw[->, ultra thick] (0,0)--++(1,0)--++(0,\h)--++(1.5,0)--++(0,.5); \draw[->, ultra thick] (1.5,-.5)--++(0,.5)--++(1.5,0); \node at (1.8,-.3) {$\lambda_i$}; \node at (1.3,\h/2) {$\kappa_i$}; \node at (2.85,\h+.3) {$\mu_i$}; \node at (2.7,-.4) {(b)}; \end{tikzpicture} % \qquad \qquad % \begin{tikzpicture} % [scale=1] % \def\h{.7} % \draw[dashed] (0,0)--++(3,0); % \draw[dashed] (0,\h)--++(3,0); % \draw[->, ultra thick] (0,0)--++(1,0)--++(0,\h)--++(1.5,0)--++(0,.5); % \draw[->, ultra thick] (0,\h)--++(1,0)--++(0,.5); % \draw[->, ultra thick] (1.5,-.5)--++(0,.5)--++(1.5,0); % \node at (1.8,-.3) {$\lambda_i$}; % \node at (1.3,\h/2) {$\kappa_i$}; % \node at (2.85,\h+.3) {$\mu_i$}; % \node at (.5,\h+.3) {$\mu_{i+1}$}; % \node at (2.7,-.4) {(c)}; % \end{tikzpicture} \qquad \begin{tikzpicture} [scale=1] \def\h{.7} \draw[dashed] (0,0)--++(3,0); \draw[dashed] (0,\h)--++(3,0); \draw[->, ultra thick] (0,0)--++(1,0)--++(0,\h)--++(.5,0)--++(0,.5); \draw[->, ultra thick] (1.5,-.5)--++(0,.5)--++(1.5,0); \node at (1.8,-.3) {$\lambda_i$}; \node at (1.3,\h/2) {$\kappa_i$}; \node at (1.8,\h+.3) {$\mu_i$}; \node at (2.7,-.4) {(c)}; \end{tikzpicture} \qquad \begin{tikzpicture} [scale=1] \def\h{.7} \draw[dashed] (0,0)--++(3,0); \draw[dashed] (0,\h)--++(3,0); \draw[->, ultra thick] (0,0)--++(1,0)--++(0,\h)--++(.5,0)--++(0,.5); \draw[->, ultra thick] (1.55,-.5)--++(0,.5)--++(1.45,0); \draw[->, ultra thick] (1.45,-.5)--++(0,.5)--++(0.05,0)--++(0,\h)--++(1.5,0); \node at (1.8,-.3) {$\lambda_i$}; \node at (1.99,\h/2) {$\kappa_{i-1}$}; \node at (.7,\h/2) {$\kappa_{i}$}; \node at (1.8,\h+.3) {$\mu_i$}; \node at (2.7,-.4) {(d)}; \end{tikzpicture} \qquad \begin{tikzpicture} [scale=1] \def\h{.7} \draw[dashed] (0,0)--++(3,0); \draw[dashed] (0,\h)--++(3,0); \draw[->, ultra thick] (0,0)--++(1,0)--++(0,\h)--++(0,0)--++(0,.5); \draw[->, ultra thick] (1.5,-.5)--++(0,.5)--++(1.5,0); \node at (1.8,-.3) {$\lambda_i$}; \node at (1.3,\h/2) {$\kappa_i$}; \node at (1.35,\h+.3) {$\mu_i$}; \node at (2.7,-.4) {(e)}; \end{tikzpicture} % \qquad \qquad % \begin{tikzpicture} % [scale=1] % \def\h{.7} % \draw[dashed] (0,0)--++(3,0); % \draw[dashed] (0,\h)--++(3,0); % \draw[->, ultra thick] (0,0)--++(1,0)--++(0,\h)--++(.05,0)--++(0,.5); % \draw[->, ultra thick] (0,\h)--++(.95,0)--++(0,.5); % \draw[->, ultra thick] (1.5,-.5)--++(0,.5)--++(1.5,0); % \node at (1.8,-.3) {$\lambda_i$}; % \node at (1.3,\h/2) {$\kappa_i$}; % \node at (2,\h+.3) {$\mu_i=\mu_{i+1}$}; % \node at (2.7,-.4) {(g)}; % \end{tikzpicture} \caption{Arrow configurations $\lambda\mathop{\dot\succ}\kappa\prec \mu$ in the proof of \Cref{prop:YB_Schur_cont_field}.} \label{fig:Schur_YB_proof} %\end{noindent} \end{figure} \smallskip \noindent \textit{Move propagation.} Assume now that $\lambda\ne \kappa$, and the difference between these two signatures at level $k-1$ can be only at one location, $\lambda_i=\kappa_i+1$. This fact would follow by induction on levels of the array after we show that the move propagation mechanism is as in \Cref{def:Schur_dyn_YB}. Indeed, this would imply that a single move of a particle by one cannot result in a move of a particle by more than one, or moves by more than one particle, at the level one higher. Then in the process of dragging the cross vertex through the arrow configuration $\lambda\mathop{\dot\succ}\kappa\prec \mu$ to obtain the signature $\nu$ all updates are deterministic: an event with probability $O(v^{-1})$ has already occured at level $k-1$ or below, and at a single time moment two or more such events cannot occur. Updates through the parts of the configuration where $\lambda_j=\kappa_j$ have been considered above: they all lead to setting $\nu_j=\mu_j$. Thus, it remains to consider the update coming from the passing of the cross vertex through the part of the configuration where $\lambda_i=\kappa_i+1$. There are four basic cases, see \Cref{fig:Schur_YB_proof}(b)-(e): \begin{itemize} \item (b) When $\mu_i>\lambda_i$ and $\mu_{i+1}<\kappa_{i}$, the cross vertex is updated as $\oiYBio{.25}{3.5}{10.5pt} \rightsquigarrow\oiYBoi{.25}{3.5}{10.5pt} \rightsquigarrow\iiYBii{.25}{3.5}{10.5pt}$. This removes the arrow at $\kappa_i$ and corresponds to $U(\mu\to\nu\mid \kappa\to\lambda)=\mathbf{1}_{\nu=\mu}$. \item (c) When $\mu_i=\lambda_i=\kappa_i+1<\lambda_{i-1}$, the update is $\oiYBio{.25}{3.5}{10.5pt} \rightsquigarrow\oiYBoi{.25}{3.5}{10.5pt} \rightsquigarrow\ioYBio{.25}{3.5}{10.5pt} \rightsquigarrow\oiYBio{.25}{3.5}{10.5pt}$, which removes the arrow at $\kappa_i$ and places a new arrow (corresponding to $\nu_i$ after the update) at $\lambda_i+1$, which corresponds to the push under conditions \eqref{YB_Schur_cont_field_exotic_push}. \item (d) When $\mu_i=\lambda_i=\kappa_i+1=\lambda_{i-1}$, the update is $\oiYBio{.25}{3.5}{10.5pt} \rightsquigarrow\oiYBoi{.25}{3.5}{10.5pt} \rightsquigarrow\iiYBii{.25}{3.5}{10.5pt}$, which removes the arrow at $\kappa_i$, and does not affect the arrow at $\kappa_{i-1}=\lambda_{i-1}$ which becomes $\nu_i=\mu_i$. This case violates of \eqref{YB_Schur_cont_field_exotic_push}, and thus the update rule is $U(\mu\to\nu\mid \kappa\to\lambda)=\mathbf{1}_{\nu=\mu}$. \item (e) When $\mu_i<\lambda_i$ (and necessarily $\mu_i=\lambda_i+1$), the update is $\oiYBio{.25}{3.5}{10.5pt} \rightsquigarrow\ooYBoo{.25}{3.5}{10.5pt} \rightsquigarrow\oiYBio{.25}{3.5}{10.5pt}$, which removes an arrow at $\kappa_i$ and adds a new arrow for $\nu_i$ at $\lambda_i$. This corresponds to the mandatory pushing to restore interlacing. \end{itemize} %\begin{noindent} Each of the cases (c)-(e) admits a slight variation when $\mu_{i+1}=\kappa_i>\kappa_{i+1}$. Then in the update of the cross vertex state the initial state is $\iiYBii{.25}{3.5}{10.5pt}$ instead of $\oiYBio{.25}{3.5}{10.5pt}$, but the rows of the table in \Cref{fig:fwd_Schur_cont} corresponding to these two states are the same up to $O(v^{-1})$. There is also another variation of (e) when $\lambda_{i-1}=\kappa_{i-1}=\lambda_i<\mu_{i-1}$, in which case the update is $\oiYBio{.25}{3.5}{10.5pt}\rightsquigarrow\ooYBoo{.25}{3.5}{10.5pt} \rightsquigarrow\iiYBii{.25}{3.5}{10.5pt}$. This does not remove an arrow at $\kappa_{i-1}$ which becomes $\nu_i$ after the passing of the cross, and this agrees with the mandatory pushing. This completes the proof. %\end{noindent} \end{proof} \begin{remark} \label{rmk:YB_Schur_directly_check} One can directly check that the Yang-Baxter dynamics on interlacing arrays described in \Cref{prop:YB_Schur_cont_field} in the language of interlacing arrays satisfies equation (2.20) of \cite{BorodinPetrov2013NN}. This equation implies that the dynamics acts nicely on Schur processes (i.e., in agreement with \Cref{thm:YB_field_spin_HL_process}). However, after establishing \Cref{prop:YB_Schur_cont_field}, this fact also follows as a degeneration of \Cref{thm:YB_field_spin_HL_process}. \end{remark} The dynamics of \Cref{prop:YB_Schur_cont_field} is very similar to the one constructed in \cite{BorFerr2008DF} using an idea of coupling Markov chains from \cite{DiaconisFill1990}. Namely, in the latter dynamics the absence of independent jumps and additional pushing in the special cases \eqref{YB_Schur_cont_field_exotic_jump}, \eqref{YB_Schur_cont_field_exotic_push} are eliminated. In other words, in the dynamics of \cite{BorFerr2008DF} every particle simply jumps to the right by one at rate $u$ while obeying the blocking and the mandatory pushing rules. On the other hand, the Hall-Littlewood RSK field introduced in \cite{BufetovMatveev2017} in the half-continuous $t=0$ limit turns into a continuous time dynamics on interlacing arrays related to the column insertion Robinson-Schensted-Knuth (RSK) correspondence. In this dynamics, only the leftmost particles $\lambda^{(j)}_j$ can independently jump. At the same time, each move (to the right by one) of a particle $\lambda^{(j)}_i$ triggers a move of a particle to the right of it on the upper level. Typically, this triggered particle is $\lambda^{(j+1)}_i$, but the move is donated to the right if it is blocked. We refer to \cite[Section 7]{BorodinPetrov2013NN} for a detailed description of this dynamics related to the (column) RSK. Since this RSK dynamics differs from the one coming from the Yang-Baxter field via \Cref{prop:YB_Schur_cont_field}, we see that the Hall-Littlewood RSK field of \cite{BufetovMatveev2017} also differs from the $s=0$ Yang-Baxter field of \Cref{sub:degen_HL}. \subsection{Half-continuous Hall-Littlewood degeneration with rescaling} \label{sub:hc_degen_IHL} Renaming $u=-su$ and slowing the continuous time (equivalently, rescaling the continuous horizontal direction in the vertex model language) by the factor $(-s)$ makes the rates and probabilities in the half-continuous DS6V independent of $s$. Then we can send $s\to0$ and obtain a well-defined dynamic half-continuous vertex model. This model can be also obtained as a half-continuous limit $v\to+\infty$ of the one described in \Cref{sub:degen_IHL}. The resulting rates and probabilities for this model are listed in \Cref{fig:dynamic_vertex_weights_degen_half_cont}(e). \subsection{Half-continuous Schur degeneration with rescaling} \label{sub:hc_degen_t0s0} Further setting $t=0$ in the model of \Cref{sub:hc_degen_IHL} turns the rates and probabilities into the ones in \Cref{fig:dynamic_vertex_weights_degen_half_cont}(f). Via a simple time rescaling, this model becomes the same as the modified continuous time PushTASEP considered in \Cref{sub:hc_degen_t0}. \subsection{Rational limit $t\to1$} \label{sub:degen_rational} In this and the following subsections we return to the original DS6V weights as in \Cref{fig:L_dynamicS6V_probabilities}. Let us take limit $t\to1$ in these weights, simultaneously rescaling all other parameters: \begin{equation*} t=e^{\varepsilon},\qquad s=e^{\varepsilon\zeta}, \qquad u=e^{x\varepsilon},\qquad v=e^{-y\varepsilon}, \qquad \varepsilon\to0. \end{equation*} In this limit the vertex weights turn into the following: \begin{align} \nonumber %\begin{noindent} \biggl[\scalebox{.7}{ \begin{tikzpicture}[scale=1.2, very thick, baseline=-4pt] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell+1$}; \node[anchor=north west] at (.1,-.1) {$\ell-1$}; \node[anchor=south west] at (.1,.1) {$\ell$}; \draw[ultra thick,->] (-.6,0)--++(.55,0); \draw[ultra thick,->] (0,-.6)--++(0,.55); \draw[ultra thick,->] (0,0)--++(.55,0); \draw[ultra thick,->] (0,0)--++(0,.55); \end{tikzpicture} }\biggr] & =1, \qquad \biggl[\scalebox{.7}{ \begin{tikzpicture}[scale=1.2, very thick, baseline=-4pt] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell$}; \node[anchor=north west] at (.1,-.1) {$\ell$}; \node[anchor=south west] at (.1,.1) {$\ell$}; \end{tikzpicture} }\biggr]=1, \\ \biggl[\scalebox{.7}{ \begin{tikzpicture}[scale=1.2, very thick, baseline=-4pt] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell+1$}; \node[anchor=north west] at (.1,-.1) {$\ell$}; \node[anchor=south west] at (.1,.1) {$\ell$}; \draw[ultra thick,->] (-.6,0)--++(.6,0); \draw[ultra thick,->] (0,0)--++(0,.6); \end{tikzpicture} }\biggr] & =\frac{\ell-x+\zeta}{(x+y+1)(\ell+y+\zeta)}, \qquad \qquad \biggl[\scalebox{.7}{ \begin{tikzpicture}[scale=1.2, very thick, baseline=-4pt] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell+1$}; \node[anchor=north west] at (.1,-.1) {$\ell$}; \node[anchor=south west] at (.1,.1) {$\ell+1$}; \draw[ultra thick,->] (-.6,0)--++(.6,0); \draw[ultra thick,->] (0,0)--++(.6,0); \end{tikzpicture} }\biggr]= \frac{(x+y)(\ell+y+\zeta+1)}{(x+y+1)(\ell+y+\zeta)}, \label{rational_limit} \\ \biggl[\scalebox{.7}{ \begin{tikzpicture}[scale=1.2, very thick, baseline=-4pt] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell$}; \node[anchor=north west] at (.1,-.1) {$\ell-1$}; \node[anchor=south west] at (.1,.1) {$\ell$}; \draw[ultra thick,->] (0,-.6)--++(0,.6); \draw[ultra thick,->] (0,0)--++(.6,0); \end{tikzpicture} }\biggr] & = \frac{\ell+y+\zeta}{(x+y+1)(\ell-x+\zeta)}\,\mathbf{1}_{\ell\ge1},\qquad \biggl[\scalebox{.7}{ \begin{tikzpicture}[scale=1.2, very thick, baseline=-4pt] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell$}; \node[anchor=north west] at (.1,-.1) {$\ell-1$}; \node[anchor=south west] at (.1,.1) {$\ell-1$}; \draw[ultra thick,->] (0,-.6)--++(0,.6); \draw[ultra thick,->] (0,0)--++(0,.6); \end{tikzpicture} }\biggr]= \frac{(x+y)(\ell-x+\zeta-1)}{(x+y+1)(\ell-x+\zeta)}\,\mathbf{1}_{\ell\ge1} . %\end{noindent} \nonumber \end{align} These weights are dynamic in the sense that they depend on the height function $\ell$. Moreover, under certain restrictions on the parameters (for example, if $x,y>0$ and $\zeta>x$), these weights are between $0$ and $1$ for all $\ell\in \mathbb{Z}_{\ge0}$. Thus, the weights \eqref{rational_limit} define a dynamic stochastic vertex model. Its height function is identified via \Cref{cor:dyn6V_spin_HL_process} with an observable of a measure constructed out of rational symmetric functions of \cite[Section 8.5]{Borodin2014vertex}. The Hall-Littlewood case $(s=0)$ corresponds to setting $\zeta\to+\infty$ in the weights \eqref{rational_limit}. This vertex model is no longer dynamic, it has symmetric vertex weights (i.e., the probabilities for a path to turn right or left are both equal to $1/(1+x+y)$) and can be regarded as a discrete time version of the symmetric simple exclusion process (SSEP). One can thus say that the limit $t\to1$ for $s=0$ corresponds to the transition from the XXZ to the XXX model, and the model \eqref{rational_limit} can be regarded as a dynamic version of SSEP/XXX. \subsection{Limit to a dynamic version of ASEP} \label{sub:degen_ASEP} Here we consider a limit of DS6V to a continuous time particle system generalizing the ASEP (Asymmetric Simple Exclusion Process). For the stochastic six vertex model a limit to the usual ASEP was observed in \cite{GwaSpohn1992} (see also \cite{BCG6V} for details). Recall that the spectral parameters of the DS6V weights satisfy $0\le u] (-.6,0)--++(.6,0); \draw[ultra thick,->] (0,0)--++(0,.6); \end{tikzpicture} }\biggr] & = 1+O(v-u), \qquad \biggl[\scalebox{.7}{ \begin{tikzpicture}[scale=1.2, very thick, baseline=-4pt] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell+1$}; \node[anchor=north west] at (.1,-.1) {$\ell$}; \node[anchor=south west] at (.1,.1) {$\ell+1$}; \draw[ultra thick,->] (-.6,0)--++(.6,0); \draw[ultra thick,->] (0,0)--++(.6,0); \end{tikzpicture} }\biggr]= \frac{(v-u) (u-s t^{\ell+1})}{(1-t) u (u-s t^\ell)}+O(v-u)^2, \\ \biggl[\scalebox{.7}{ \begin{tikzpicture}[scale=1.2, very thick, baseline=-4pt] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell$}; \node[anchor=north west] at (.1,-.1) {$\ell-1$}; \node[anchor=south west] at (.1,.1) {$\ell$}; \draw[ultra thick,->] (0,-.6)--++(0,.6); \draw[ultra thick,->] (0,0)--++(.6,0); \end{tikzpicture} }\biggr] & = 1+O(v-u) , \qquad \biggl[\scalebox{.7}{ \begin{tikzpicture}[scale=1.2, very thick, baseline=-4pt] \draw[dashed] (0,-.6)--++(0,1.2); \draw[dashed] (-.6,0)--++(1.2,0); \node[anchor=north east] at (-.1,-.1) {$\ell$}; \node[anchor=south east] at (-.1,.1) {$\ell$}; \node[anchor=north west] at (.1,-.1) {$\ell-1$}; \node[anchor=south west] at (.1,.1) {$\ell-1$}; \draw[ultra thick,->] (0,-.6)--++(0,.6); \draw[ultra thick,->] (0,0)--++(0,.6); \end{tikzpicture} }\biggr]= \frac{(v-u)t (u-s t^{\ell-1})}{(1-t) u (u-s t^\ell)} +O(v-u)^2 . %\end{noindent} \end{align*} We thus see that the up-right lattice paths perform staircase like movements most of the time. Occasionally, however, these staircases move up or down according to the weights of the vertices $(1,0;1,0)$ and $(0,1;0,1)$, respectively. Subtracting the staircase movement, rescaling the vertical direction by the factor of $\frac{v-u}{u(1-t)}$, and interpreting it as time leads to the following continuous time particle system on $\mathbb{Z}$. The particles are ordered as $y_1(\tau)>y_2(\tau)>\ldots $, and at most one particle per site is allowed. The six vertex boundary condition translates into the step initial condition $y_i(0)=-i$, $i\ge1$. In continuous time, each particle $y_{\ell}$ tries to jump to the right by one at rate $\dfrac{u-st^{\ell}}{u-st^{\ell-1}}$, and to the left by one at rate $t\,\dfrac{u-st^{\ell-1}}{u-st^{\ell}}$. If the destination is occupied, the corresponding jump is blocked and $y_{\ell}$ does not move. See \Cref{fig:dyn_ASEP} in the Introduction. Thus, one can say that our dynamic ASEP is a generalization of the ASEP with certain particle-dependent jump rates. The connection to spin Hall-Littlewood measures might provide tools for asymptotic analysis of this model. The dynamic version of the ASEP obtained above is somewhat similar to the one of \cite{borodin2017elliptic}, \cite{BorodinCorwin2017dynamic} coming from vertex models at elliptic level. However, these two models are different. In particular, in our model the dynamic dependence on the height function is via the quantities $h_x=\#\{\textnormal{number of particles to the right of $x$}\}$, while in \cite{borodin2017elliptic}, \cite{BorodinCorwin2017dynamic} the dynamic parameter is $s_x=2h_x+x$ which incorporates both the particle's number and location. \subsection{Finite vertical spin} \label{sub:degen_finite_spin} Setting $s=t^{-I/2}$, where $I\in \mathbb{Z}_{\ge1}$, turns the vertical representation giving rise to the vertex weights \eqref{vertex_weights} into a spin $\frac{I}{2}$ one. This gives rise to a vertex model with at most $I$ vertical arrows per edge allowed. Let us briefly discuss what this means for the main constructions of the present paper. For simplicity, we only consider the case $I=1$ when the higher spin six vertex model turns into the six vertex model. Call a signature $\lambda\in \mathsf{Sign}_N$ \emph{strict} if $\lambda_1>\lambda_2>\ldots>\lambda_N $. Observe that for $s=t^{-\frac{1}{2}}$ the weight \begin{equation*} \Big[ \Voi{.6}g{g-1}{-4.5pt} \Big]_{u}= \dfrac{(1-t^{g-2})u}{1-ut^{-\frac{1}{2}}} \end{equation*} vanishes for $g=2$. Thus, $G_{\mu/\nu}^c(v)$ also vanishes if $\nu$ is strict and $\mu$ is not, see \Cref{ssub:G_definition}. At the same time the function $G^{c}_{\lambda/(0^N)}$ entering the spin Hall-Littlewood measure \eqref{spin_HL_measure} is not well-defined since $(0^{N})$ is not strict. This presents an obstacle in degenerating spin Hall-Littlewood measures and processes to $s=t^{-\frac{1}{2}}$ in a straightforward way. On the other hand, the vertex weights for $s=t^{-\frac{1}{2}}$ satisfy a Yang-Baxter equation, and bijectivisation can be applied to it, too. Following the lines of \Cref{sec:local_transition_probabilities}, one can define forward transition probabilities $\mathsf{U}^{\mathrm{fwd}}(\kappa\to\nu\mid\lambda,\mu)$, where $\kappa,\lambda\in\mathsf{Sign}_{N-1}$ and $\mu,\nu\in\mathsf{Sign}_{N}$ are strict. Using these probabilities, it is possible to define an analogue of the Yang-Baxter field $\boldsymbol\lambda^{(x,y)}$, $x,y\in\mathbb{Z}_{\ge0}$, with boundary conditions $\boldsymbol\lambda^{(x,0)}=\varnothing$, $\boldsymbol\lambda^{(0,y)}=(-1,-2,\ldots,-y )$. It is not clear whether this version of the Yang-Baxter field leads via Markov projections to an analogue of the dynamic stochastic six vertex model of \Cref{sec:dynamicS6V}, and we do not discuss this issue here. \section{Yang-Baxter equation} \label{app:YB_equation} Here we write out all the explicit identities between rational functions which comprise the Yang-Baxter equation. This equation states that certain combinations of vertex weights \eqref{vertex_weights}, \eqref{cross_vertex_weights} are equal to each other. Writing all possible cases out we arrive at the following 16 identities. For better notation, in the vertex weights we put cross vertices together with pairs of vertices, and use the shorthand \begin{equation} \label{YB_spectral_parameters_convention} \left[ \cdots \right]:=\left[ \cdots \right]_{u,v}, \qquad \left[ \cdots \right]':=\left[ \cdots \right]_{v,u} \end{equation} for the vertex weights. Moreover, by agreement, the weights of the cross vertices are not affected by the swapping of spectral parameters, and are given by \eqref{cross_vertex_weights} in both sides of each of the identities. Below are all the 16 identities comprising the Yang-Baxter equation. They depend on an arbitrary nonnegative integer $g$ subject to the agreement that once an arrow configuration in either side of a formula contains $g-1$ or $g-2$, we assume that $g\ge1$ or $g\ge2$, respectively. Each of the identities below is readily verified by hand: \begin{align} \label{YB1.1}\tag{YB1.1} & \biggl[ \ooYBoo{.3}{3}{13.5pt}\ooWoo{.6}ggg{6.75pt} \biggr] = \tfrac{(1-st^gu)(1-st^gv)}{(1-su)(1-sv)} = \biggl[ \ooWoo{.6}ggg{6.75pt} \ooYBoo{.3}{3}{13.5pt} \biggr]' ; \\ \label{YB1.2}\tag{YB1.2} \begin{split} &\biggl[ \ooYBoo{.3}{3}{13.5pt}\ooWio{.6}g{g-1}{g-1}{6.75pt} \biggr] = \tfrac{(1-s^2t^{g-1})u(1-st^{g-1}v)}{(1-su)(1-sv)} \\[-8pt]&\hspace{5pt}= \tfrac{(1-s^2t^{g-1})v(1-st^{g-1}u)}{(1-sv)(1-su)} \tfrac{(1-t)u}{u-tv}+ \tfrac{(1-st^gv)(1-s^2t^{g-1})u}{(1-sv)(1-su)} \tfrac{u-v}{u-tv} = \biggl[ \ooWio{.6}g{g-1}{g-1}{6.75pt}\ioYBio{.3}{3}{13.5pt} \biggr]'+ \biggl[ \ooWoi{.6}g{g}{g-1}{6.75pt}\oiYBio{.3}{3}{13.5pt} \biggr]'; \end{split} \\ \label{YB1.3}\tag{YB1.3} \begin{split} &\biggl[ \ooYBoo{.3}{3}{13.5pt}\ooWoi{.6}g{g}{g-1}{6.75pt} \biggr] = \tfrac{(1-st^gu)(1-s^2t^{g-1})v}{(1-su)(1-sv)} \\[-8pt]&\hspace{5pt}= \tfrac{(1-st^gv)(1-s^2t^{g-1})u}{(1-sv)(1-su)} \tfrac{(1-t)v}{u-tv}+ \tfrac{(1-s^2t^{g-1})v(1-st^{g-1}u)}{(1-sv)(1-su)} \tfrac{t(u-v)}{u-tv} = \biggl[ \ooWoi{.6}g{g}{g-1}{6.75pt}\oiYBoi{.3}{3}{13.5pt} \biggr]'+ \biggl[ \ooWio{.6}g{g-1}{g-1}{6.75pt}\ioYBoi{.3}{3}{13.5pt} \biggr]' ; \end{split} \\ \label{YB1.4}\tag{YB1.4} & \biggl[ \ooYBoo{.3}{3}{13.5pt}\ooWii{.6}g{g-1}{g-2}{6.75pt} \biggr] = \tfrac{(1-s^2t^{g-1})u(1-s^2t^{g-2})v}{(1-su)(1-sv)} = \biggl[ \ooWii{.6}g{g-1}{g-2}{6.75pt}\iiYBii{.3}{3}{13.5pt} \biggr]' ; \\ \label{YB2.1}\tag{YB2.1} \begin{split} & \biggl[ \ioYBio{.3}{3}{13.5pt}\oiWoo{.6}g{g+1}{g+1}{6.75pt} \biggr] + \biggl[ \ioYBoi{.3}{3}{13.5pt}\ioWoo{.6}gg{g+1}{6.75pt} \biggr] = \tfrac{(1-t)u}{u-tv} \tfrac{(1-t^{g+1})(1-st^{g+1}v)}{(1-su)(1-sv)} + \tfrac{t(u-v)}{u-tv} \tfrac{(1-st^gu)(1-t^{g+1})}{(1-su)(1-sv)} \\[-8pt]& \hspace{5pt} = \tfrac{(1-t^{g+1})(1-st^{g+1}u)}{(1-sv)(1-su)} = \biggl[ \oiWoo{.6}g{g+1}{g+1}{6.75pt}\ooYBoo{.3}{3}{13.5pt} \biggr]' ; \end{split} \\ \label{YB2.2}\tag{YB2.2} \begin{split} & \biggl[ \ioYBio{.3}{3}{13.5pt}\oiWio{.6}ggg{6.75pt} \biggr] + \biggl[ \ioYBoi{.3}{3}{13.5pt}\ioWio{.6}g{g-1}g{6.75pt} \biggr] = \tfrac{(1-t)u}{u-tv} \tfrac{(u-st^g)(1-st^g v)}{(1-su)(1-sv)} + \tfrac{t(u-v)}{u-tv} \tfrac{(1-s^2t^{g-1})u(1-t^g)}{(1-su)(1-sv)} \\[-8pt]& \hspace{5pt} = \tfrac{(v-st^g)(1-st^gu)}{(1-sv)(1-su)} \tfrac{(1-t)u}{u-tv} + \tfrac{(1-t^{g+1})(1-s^2t^g)u}{(1-sv)(1-su)} \tfrac{u-v}{u-tv} = \biggl[ \oiWio{.6}g{g}{g}{6.75pt}\ioYBio{.3}{3}{13.5pt} \biggr]' + \biggl[ \oiWoi{.6}g{g+1}{g}{6.75pt}\oiYBio{.3}{3}{13.5pt} \biggr]' ; \end{split} \\ \label{YB2.3}\tag{YB2.3} \begin{split} & \biggl[ \ioYBio{.3}{3}{13.5pt}\oiWoi{.6}g{g+1}g{6.75pt} \biggr] + \biggl[ \ioYBoi{.3}{3}{13.5pt}\ioWoi{.6}ggg{6.75pt} \biggr] = \tfrac{(1-t)u}{u-tv} \tfrac{(1-t^{g+1})(1-s^2t^g)v}{(1-su)(1-sv)} + \tfrac{t(u-v)}{u-tv} \tfrac{(1-st^gu)(v-st^g)}{(1-su)(1-sv)} \\[-8pt]& \hspace{5pt} = \tfrac{(1-t^{g+1})(1-s^2t^g)u}{(1-sv)(1-su)} \tfrac{(1-t)v}{u-tv} + \tfrac{(v-st^g)(1-st^gu)}{(1-sv)(1-su)} \tfrac{t(u-v)}{u-tv} = \biggl[ \oiWoi{.6}g{g+1}{g}{6.75pt}\oiYBoi{.3}{3}{13.5pt} \biggr]' + \biggl[ \oiWio{.6}g{g}{g}{6.75pt}\ioYBoi{.3}{3}{13.5pt} \biggr]' ; \end{split} \\ \label{YB2.4}\tag{YB2.4} \begin{split} & \biggl[ \ioYBio{.3}{3}{13.5pt}\oiWii{.6}gg{g-1}{6.75pt} \biggr] + \biggl[ \ioYBoi{.3}{3}{13.5pt}\ioWii{.6}g{g-1}{g-1}{6.75pt} \biggr] = \tfrac{(1-t)u}{u-tv} \tfrac{(u-st^g)(1-s^2t^{g-1})v}{(1-su)(1-sv)} + \tfrac{t(u-v)}{u-tv} \tfrac{(1-s^2t^{g-1})u(v-st^{g-1})}{(1-su)(1-sv)} \\[-8pt]& \hspace{5pt} = \tfrac{(v-st^g)(1-s^2t^{g-1})u}{(1-sv)(1-su)} = \biggl[ \oiWii{.6}g{g}{g-1}{6.75pt}\iiYBii{.3}{3}{13.5pt} \biggr]' ; \end{split} \\ \label{YB3.1}\tag{YB3.1} \begin{split} & \biggl[ \oiYBoi{.3}{3}{13.5pt}\ioWoo{.6}gg{g+1}{6.75pt} \biggr] + \biggl[ \oiYBio{.3}{3}{13.5pt}\oiWoo{.6}g{g+1}{g+1}{6.75pt} \biggr] = \tfrac{(1-t)v}{u-tv} \tfrac{(1-st^gu)(1-t^{g+1})}{(1-su)(1-sv)} + \tfrac{u-v}{u-tv} \tfrac{(1-t^{g+1})(1-st^{g+1}v)}{(1-su)(1-sv)} \\[-8pt]&\hspace{5pt}= \tfrac{(1-st^gv)(1-t^{g+1})}{(1-sv)(1-su)} = \biggl[ \ioWoo{.6}gg{g+1}{6.75pt}\ooYBoo{.3}{3}{13.5pt} \biggr]' ; \end{split} \\ \label{YB3.2}\tag{YB3.2} \begin{split} & \biggl[ \oiYBoi{.3}{3}{13.5pt}\ioWio{.6}g{g-1}g{6.75pt} \biggr] + \biggl[ \oiYBio{.3}{3}{13.5pt}\oiWio{.6}ggg{6.75pt} \biggr] = \tfrac{(1-t)v}{u-tv} \tfrac{(1-s^2t^{g-1})u(1-t^g)}{(1-su)(1-sv)} + \tfrac{u-v}{u-tv} \tfrac{(u-st^g)(1-st^gv)}{(1-su)(1-sv)} \\[-8pt]&\hspace{5pt}= \tfrac{(1-s^2t^{g-1})v(1-t^g)}{(1-sv)(1-su)} \tfrac{(1-t)u}{u-tv} + \tfrac{(1-st^gv)(u-st^g)}{(1-sv)(1-su)} \tfrac{u-v}{u-tv} = \biggl[ \ioWio{.6}g{g-1}g{6.75pt}\ioYBio{.3}{3}{13.5pt} \biggr]' + \biggl[ \ioWoi{.6}ggg{6.75pt}\oiYBio{.3}{3}{13.5pt} \biggr]' ; \end{split} \\ \label{YB3.3}\tag{YB3.3} \begin{split} & \biggl[ \oiYBoi{.3}{3}{13.5pt}\ioWoi{.6}ggg{6.75pt} \biggr] + \biggl[ \oiYBio{.3}{3}{13.5pt}\oiWoi{.6}g{g+1}g{6.75pt} \biggr] = \tfrac{(1-t)v}{u-tv} \tfrac{(1-st^gu)(v-st^g)}{(1-su)(1-sv)} + \tfrac{u-v}{u-tv} \tfrac{(1-t^{g+1})(1-s^2t^g)v}{(1-su)(1-sv)} \\[-8pt]&\hspace{5pt}= \tfrac{(1-st^gv)(u-st^g)}{(1-sv)(1-su)} \tfrac{(1-t)v}{u-tv} + \tfrac{(1-s^2t^{g-1})v(1-t^g)}{(1-sv)(1-su)} \tfrac{t(u-v)}{u-tv} = \biggl[ \ioWoi{.6}g{g}g{6.75pt}\oiYBoi{.3}{3}{13.5pt} \biggr]' + \biggl[ \ioWio{.6}g{g-1}g{6.75pt}\ioYBoi{.3}{3}{13.5pt} \biggr]' ; \end{split} \\ \label{YB3.4}\tag{YB3.4} \begin{split} & \biggl[ \oiYBoi{.3}{3}{13.5pt}\ioWii{.6}g{g-1}{g-1}{6.75pt} \biggr] + \biggl[ \oiYBio{.3}{3}{13.5pt}\oiWii{.6}gg{g-1}{6.75pt} \biggr] = \tfrac{(1-t)v}{u-tv} \tfrac{(1-s^2t^{g-1})u(v-st^{g-1})}{(1-su)(1-sv)} + \tfrac{u-v}{u-tv} \tfrac{(u-st^g)(1-s^2t^{g-1})v}{(1-su)(1-sv)} \\[-8pt]&\hspace{5pt}= \tfrac{(1-s^2t^{g-1})v(u-st^{g-1})}{(1-sv)(1-su)} = \biggl[ \ioWii{.6}g{g-1}{g-1}{6.75pt}\iiYBii{.3}{3}{13.5pt} \biggr]' ; \end{split} \\ \label{YB4.1}\tag{YB4.1} & \biggl[ \iiYBii{.3}{3}{13.5pt}\iiWoo{.6}g{g+1}{g+2}{6.75pt} \biggr] = \tfrac{(1-t^{g+1})(1-t^{g+2})}{(1-su)(1-sv)} = \biggl[ \iiWoo{.6}g{g+1}{g+2}{6.75pt}\ooYBoo{.3}{3}{13.5pt} \biggr]' ; \\ \label{YB4.2}\tag{YB4.2} \begin{split} & \biggl[ \iiYBii{.3}{3}{13.5pt}\iiWio{.6}gg{g+1}{6.75pt} \biggr] = \tfrac{(u-st^g)(1-t^{g+1})}{(1-su)(1-sv)} \\[-8pt]&\hspace{5pt} = \tfrac{(v-st^g)(1-t^{g+1})}{(1-sv)(1-su)} \tfrac{(1-t)u}{u-tv} + \tfrac{(1-t^{g+1})(u-st^{g+1})}{(1-sv)(1-su)} \tfrac{u-v}{u-tv} = \biggl[ \iiWio{.6}g{g}{g+1}{6.75pt}\ioYBio{.3}{3}{13.5pt} \biggr]' + \biggl[ \iiWoi{.6}g{g+1}{g+1}{6.75pt}\oiYBio{.3}{3}{13.5pt} \biggr]' ; \end{split} \\ \label{YB4.3}\tag{YB4.3} \begin{split} & \biggl[ \iiYBii{.3}{3}{13.5pt}\iiWoi{.6}g{g+1}{g+1}{6.75pt} \biggr] = \tfrac{(1-t^{g+1})(v-st^{g+1})}{(1-su)(1-sv)} \\[-8pt]&\hspace{5pt} = \tfrac{(1-t^{g+1})(u-st^{g+1})}{(1-sv)(1-su)} \tfrac{(1-t)v}{u-tv} + \tfrac{(v-st^g)(1-t^{g+1})}{(1-sv)(1-su)} \tfrac{t(u-v)}{u-tv} = \biggl[ \iiWoi{.6}g{g+1}{g+1}{6.75pt}\oiYBoi{.3}{3}{13.5pt} \biggr]' + \biggl[ \iiWio{.6}g{g}{g+1}{6.75pt}\ioYBoi{.3}{3}{13.5pt} \biggr]' ; \end{split} \\ \label{YB4.4}\tag{YB4.4} & \biggl[ \iiYBii{.3}{3}{13.5pt}\iiWii{.6}g{g}{g}{6.75pt} \biggr] = \tfrac{(u-st^g)(v-st^g)}{(1-su)(1-sv)} = \biggl[ \iiWii{.6}g{g}{g}{6.75pt}\iiYBii{.3}{3}{13.5pt} \biggr] . \end{align} \section{Probabilities of forward and backward Yang-Baxter moves} \label{app:YB_probabilities} Here we list in full detail the probabilities of forward and backward Yang-Baxter moves discussed in \Cref{sub:YB_bijectivisation_new_label}. These probabilities (coming from identities \eqref{YB1.1}--\eqref{YB4.4} listed in \Cref{app:YB_equation}) depend on the spectral parameters $u,v$ and on an arbitrary nonnegative integer $g$ (which is required to be $\ge1$ if the corresponding arrow configurations contain $g-1$ vertical arrows). Equation numbers \eqref{F1.1}--\eqref{F4.4} and \eqref{B1.1}--\eqref{B4.4} below correspond to numbers of the Yang-Baxter identities in \Cref{app:YB_equation} whose bijectivisation gives these forward and backward transition probabilities. The forward transition probabilities look as follows (we do not write down transitions whose probabilities are identically zero): \begin{align} %\begin{noindent} \label{F1.1}\tag{F1.1} & P^{\mathrm{fwd}}_{u,v}\biggl( \ooYBoo{.3}{3}{13.5pt}\ooWoo{.6}ggg{6.75pt}\ ,\ \ooWoo{.6}ggg{6.75pt}\ooYBoo{.3}{3}{13.5pt} \biggr) = 1; \\ \label{F1.2}\tag{F1.2} & P^{\mathrm{fwd}}_{u,v}\biggl( \ooYBoo{.3}{3}{13.5pt}\ooWio{.6}{g+1}gg{6.75pt}\ ,\ \ooWio{.6}{g+1}gg{6.75pt}\ioYBio{.3}{3}{13.5pt} \biggr) = 1-P^{\mathrm{fwd}}_{u,v}\biggl( \ooYBoo{.3}{3}{13.5pt}\ooWio{.6}{g+1}gg{6.75pt}\ ,\ \ooWoi{.6}{g+1}{g+1}{g}{6.75pt}\oiYBio{.3}{3}{13.5pt} \biggr) = \dfrac{(1-t)v}{u-tv}\dfrac{1-st^{g}u}{1-st^{g}v} ; \\ \label{F1.3}\tag{F1.3} & P^{\mathrm{fwd}}_{u,v} \biggl( \ooYBoo{.3}{3}{13.5pt}\ooWoi{.6}{g}g{g-1}{6.75pt}\ ,\ \ooWoi{.6}{g}g{g-1}{6.75pt}\oiYBoi{.3}{3}{13.5pt} \biggr) = 1- P^{\mathrm{fwd}}_{u,v} \biggl( \ooYBoo{.3}{3}{13.5pt}\ooWoi{.6}{g}g{g-1}{6.75pt}\ ,\ \ooWio{.6}{g}{g-1}{g-1}{6.75pt}\ioYBoi{.3}{3}{13.5pt} \biggr) = \dfrac{(1-t)u}{u-tv}\dfrac{1-st^gv}{1-st^gu} ; \\ \label{F1.4}\tag{F1.4} & P^{\mathrm{fwd}}_{u,v} \biggl( \ooYBoo{.3}{3}{13.5pt}\ooWii{.6}{g+1}g{g-1}{6.75pt}\ ,\ \ooWii{.6}{g+1}{g}{g-1}{6.75pt}\iiYBii{.3}{3}{13.5pt} \biggr) =1 ; \\ \label{F2.1}\tag{F2.1} & P^{\mathrm{fwd}}_{u,v} \biggl( \ioYBio{.3}{3}{13.5pt}\oiWoo{.6}{g-1}g{g}{6.75pt}\ ,\ \oiWoo{.6}{g-1}{g}{g}{6.75pt}\ooYBoo{.3}{3}{13.5pt} \biggr) =P^{\mathrm{fwd}}_{u,v} \biggl( \ioYBoi{.3}{3}{13.5pt}\ioWoo{.6}{g}g{g+1}{6.75pt}\ ,\ \oiWoo{.6}{g}{g+1}{g+1}{6.75pt}\ooYBoo{.3}{3}{13.5pt} \biggr) =1 ; \\ \label{F2.2}\tag{F2.2} \begin{split} & P^{\mathrm{fwd}}_{u,v} \biggl( \ioYBio{.3}{3}{13.5pt}\oiWio{.6}{g}g{g}{6.75pt}\ ,\ \oiWio{.6}{g}{g}{g}{6.75pt}\ioYBio{.3}{3}{13.5pt} \biggr) = 1- P^{\mathrm{fwd}}_{u,v} \biggl( \ioYBio{.3}{3}{13.5pt}\oiWio{.6}{g}g{g}{6.75pt}\ ,\ \oiWoi{.6}{g}{g+1}{g}{6.75pt}\oiYBio{.3}{3}{13.5pt} \biggr) = \dfrac{v-st^g}{u-st^g} \dfrac{1-st^gu}{1-st^gv} ; \\ & P^{\mathrm{fwd}}_{u,v} \biggl( \ioYBoi{.3}{3}{13.5pt}\ioWio{.6}{g+1}g{g+1}{6.75pt}\ ,\ \oiWoi{.6}{g+1}{g+2}{g+1}{6.75pt}\oiYBio{.3}{3}{13.5pt} \biggr) = 1 ; \end{split} \\ \label{F2.3}\tag{F2.3} & P^{\mathrm{fwd}}_{u,v} \biggl( \ioYBio{.3}{3}{13.5pt}\oiWoi{.6}{g-1}g{g-1}{6.75pt}\ ,\ \oiWoi{.6}{g-1}{g}{g-1}{6.75pt}\oiYBoi{.3}{3}{13.5pt} \biggr) =P^{\mathrm{fwd}}_{u,v} \biggl( \ioYBoi{.3}{3}{13.5pt}\ioWoi{.6}{g}g{g}{6.75pt}\ ,\ \oiWio{.6}{g}{g}{g}{6.75pt}\ioYBoi{.3}{3}{13.5pt} \biggr) = 1 ; \\ \label{F2.4}\tag{F2.4} & P^{\mathrm{fwd}}_{u,v} \biggl( \ioYBio{.3}{3}{13.5pt}\oiWii{.6}{g}g{g-1}{6.75pt}\ ,\ \oiWii{.6}{g}{g}{g-1}{6.75pt}\iiYBii{.3}{3}{13.5pt} \biggr) =P^{\mathrm{fwd}}_{u,v} \biggl( \ioYBoi{.3}{3}{13.5pt}\ioWii{.6}{g+1}g{g}{6.75pt}\ ,\ \oiWii{.6}{g+1}{g+1}{g}{6.75pt}\iiYBii{.3}{3}{13.5pt} \biggr) =1 ; \\ \label{F3.1}\tag{F3.1} & P^{\mathrm{fwd}}_{u,v} \biggl( \oiYBio{.3}{3}{13.5pt}\oiWoo{.6}{g-1}g{g}{6.75pt}\ ,\ \ioWoo{.6}{g-1}{g-1}{g}{6.75pt}\ooYBoo{.3}{3}{13.5pt} \biggr) =P^{\mathrm{fwd}}_{u,v} \biggl( \oiYBoi{.3}{3}{13.5pt}\ioWoo{.6}{g}g{g+1}{6.75pt}\ ,\ \ioWoo{.6}{g}{g}{g+1}{6.75pt}\ooYBoo{.3}{3}{13.5pt} \biggr) =1 ; \\ \label{F3.2}\tag{F3.2} & P^{\mathrm{fwd}}_{u,v} \biggl( \oiYBoi{.3}{3}{13.5pt}\ioWio{.6}{g+1}g{g+1}{6.75pt}\ ,\ \ioWio{.6}{g+1}{g}{g+1}{6.75pt}\ioYBio{.3}{3}{13.5pt} \biggr) = P^{\mathrm{fwd}}_{u,v} \biggl( \oiYBio{.3}{3}{13.5pt}\oiWio{.6}{g}g{g}{6.75pt}\ ,\ \ioWoi{.6}{g}{g}{g}{6.75pt}\oiYBio{.3}{3}{13.5pt} \biggr) = 1 ; \\ \label{F3.3}\tag{F3.3} \begin{split} & P^{\mathrm{fwd}}_{u,v} \biggl( \oiYBio{.3}{3}{13.5pt}\oiWoi{.6}{g-1}g{g-1}{6.75pt}\ ,\ \ioWoi{.6}{g-1}{g-1}{g-1}{6.75pt}\oiYBoi{.3}{3}{13.5pt} \biggr) = 1- P^{\mathrm{fwd}}_{u,v} \biggl( \oiYBio{.3}{3}{13.5pt}\oiWoi{.6}{g-1}g{g-1}{6.75pt}\ ,\ \ioWio{.6}{g-1}{g-2}{g-1}{6.75pt}\ioYBoi{.3}{3}{13.5pt} \biggr) = \dfrac{1-t}{1-t^{g}} \dfrac{1-s^2t^{2g-2}}{1-s^2t^{g-1}} ; \\ & P^{\mathrm{fwd}}_{u,v} \biggl( \oiYBoi{.3}{3}{13.5pt}\ioWoi{.6}{g}g{g}{6.75pt}\ ,\ \ioWoi{.6}{g}{g}{g}{6.75pt}\oiYBoi{.3}{3}{13.5pt} \biggr) = 1 ; \end{split} \\ \label{F3.4}\tag{F3.4} & P^{\mathrm{fwd}}_{u,v} \biggl( \oiYBio{.3}{3}{13.5pt}\oiWii{.6}{g}g{g-1}{6.75pt}\ ,\ \ioWii{.6}{g}{g-1}{g-1}{6.75pt}\iiYBii{.3}{3}{13.5pt} \biggr) = P^{\mathrm{fwd}}_{u,v} \biggl( \oiYBoi{.3}{3}{13.5pt}\ioWii{.6}{g+1}g{g}{6.75pt}\ ,\ \ioWii{.6}{g+1}{g}{g}{6.75pt}\iiYBii{.3}{3}{13.5pt} \biggr) =1 ; \\ \label{F4.1}\tag{F4.1} & P^{\mathrm{fwd}}_{u,v} \biggl( \iiYBii{.3}{3}{13.5pt}\iiWoo{.6}{g-1}g{g+1}{6.75pt}\ ,\ \iiWoo{.6}{g-1}{g}{g+1}{6.75pt}\ooYBoo{.3}{3}{13.5pt} \biggr) =1 ; \\ \label{F4.2}\tag{F4.2} & P^{\mathrm{fwd}}_{u,v} \biggl( \iiYBii{.3}{3}{13.5pt}\iiWio{.6}{g}g{g+1}{6.75pt}\ ,\ \iiWio{.6}{g}{g}{g+1}{6.75pt}\ioYBio{.3}{3}{13.5pt} \biggr) = 1- P^{\mathrm{fwd}}_{u,v} \biggl( \iiYBii{.3}{3}{13.5pt}\iiWio{.6}{g}g{g+1}{6.75pt}\ ,\ \iiWoi{.6}{g}{g+1}{g+1}{6.75pt}\oiYBio{.3}{3}{13.5pt} \biggr) = \dfrac{(1-t)u}{u-tv}\dfrac{v-st^g}{u-st^g} ; \\ \label{F4.3}\tag{F4.3} & P^{\mathrm{fwd}}_{u,v} \biggl( \iiYBii{.3}{3}{13.5pt}\iiWoi{.6}{g-1}g{g}{6.75pt}\ ,\ \iiWoi{.6}{g-1}{g}{g}{6.75pt}\oiYBoi{.3}{3}{13.5pt} \biggr) = 1- P^{\mathrm{fwd}}_{u,v} \biggl( \iiYBii{.3}{3}{13.5pt}\iiWoi{.6}{g-1}g{g}{6.75pt}\ ,\ \iiWio{.6}{g-1}{g-1}{g}{6.75pt}\ioYBoi{.3}{3}{13.5pt} \biggr) = \dfrac{(1-t)v}{u-tv} \dfrac{u-st^{g}}{v-st^{g}}; \\ \label{F4.4}\tag{F4.4} & P^{\mathrm{fwd}}_{u,v} \biggl( \iiYBii{.3}{3}{13.5pt}\iiWii{.6}{g}g{g}{6.75pt}\ ,\ \iiWii{.6}{g}{g}{g}{6.75pt}\iiYBii{.3}{3}{13.5pt} \biggr) =1. %\end{noindent} \end{align} The backward transition probabilities have the following form (again, we omit transitions having zero probability): \begin{align} %\begin{noindent} \label{B1.1}\tag{B1.1} & P^{\mathrm{bwd}}_{u,v}\biggl( \ooWoo{.6}ggg{6.75pt}\ooYBoo{.3}{3}{13.5pt} \ ,\ \ooYBoo{.3}{3}{13.5pt}\ooWoo{.6}ggg{6.75pt} \biggr) = 1 ; \\ \label{B1.2}\tag{B1.2} & P^{\mathrm{bwd}}_{u,v}\biggl( \ooWio{.6}{g+1}gg{6.75pt}\ioYBio{.3}{3}{13.5pt} \ ,\ \ooYBoo{.3}{3}{13.5pt}\ooWio{.6}{g+1}gg{6.75pt} \biggr) = P^{\mathrm{bwd}}_{u,v}\biggl( \ooWoi{.6}{g}{g}{g-1}{6.75pt}\oiYBio{.3}{3}{13.5pt} \ ,\ \ooYBoo{.3}{3}{13.5pt}\ooWio{.6}{g}{g-1}{g-1}{6.75pt} \biggr) = 1 ; \\ \label{B1.3}\tag{B1.3} & P^{\mathrm{bwd}}_{u,v} \biggl( \ooWoi{.6}{g}g{g-1}{6.75pt}\oiYBoi{.3}{3}{13.5pt} \ ,\ \ooYBoo{.3}{3}{13.5pt}\ooWoi{.6}{g}g{g-1}{6.75pt} \biggr) = P^{\mathrm{bwd}}_{u,v} \biggl( \ooWio{.6}{g+1}{g}{g}{6.75pt}\ioYBoi{.3}{3}{13.5pt} \ ,\ \ooYBoo{.3}{3}{13.5pt}\ooWoi{.6}{g+1}{g+1}{g}{6.75pt} \biggr) =1 ; \\ \label{B1.4}\tag{B1.4} & P^{\mathrm{bwd}}_{u,v} \biggl( \ooWii{.6}{g+1}{g}{g-1}{6.75pt}\iiYBii{.3}{3}{13.5pt} \ ,\ \ooYBoo{.3}{3}{13.5pt}\ooWii{.6}{g+1}g{g-1}{6.75pt} \biggr) =1 ; \\ \label{B2.1}\tag{B2.1} & P^{\mathrm{bwd}}_{u,v} \biggl( \oiWoo{.6}{g-1}{g}{g}{6.75pt}\ooYBoo{.3}{3}{13.5pt} \ ,\ \ioYBio{.3}{3}{13.5pt}\oiWoo{.6}{g-1}g{g}{6.75pt} \biggr) = 1- P^{\mathrm{bwd}}_{u,v} \biggl( \oiWoo{.6}{g-1}{g}{g}{6.75pt}\ooYBoo{.3}{3}{13.5pt} \ ,\ \ioYBoi{.3}{3}{13.5pt}\ioWoo{.6}{g-1}{g-1}{g}{6.75pt} \biggr) = \dfrac{(1-t)u}{u-tv} \dfrac{1-st^gv}{1-st^gu} ; \\ \label{B2.2}\tag{B2.2} \begin{split} & P^{\mathrm{bwd}}_{u,v} \biggl( \oiWio{.6}{g}{g}{g}{6.75pt}\ioYBio{.3}{3}{13.5pt} \ ,\ \ioYBio{.3}{3}{13.5pt}\oiWio{.6}{g}g{g}{6.75pt} \biggr) =1 ; \\ & P^{\mathrm{bwd}}_{u,v} \biggl( \oiWoi{.6}{g-1}{g}{g-1}{6.75pt}\oiYBio{.3}{3}{13.5pt} \ ,\ \ioYBio{.3}{3}{13.5pt}\oiWio{.6}{g-1}{g-1}{g-1}{6.75pt} \biggr) = 1 - P^{\mathrm{bwd}}_{u,v} \biggl( \oiWoi{.6}{g-1}{g}{g-1}{6.75pt}\oiYBio{.3}{3}{13.5pt} \ ,\ \ioYBoi{.3}{3}{13.5pt}\ioWio{.6}{g-1}{g-2}{g-1}{6.75pt} \biggr) = \frac{1-t}{1-t^g}\frac{1-s^2t^{2g-2}}{1-s^2t^{g-1}} ; \end{split} \\ \label{B2.3}\tag{B2.3} & P^{\mathrm{bwd}}_{u,v} \biggl( \oiWoi{.6}{g-1}{g}{g-1}{6.75pt}\oiYBoi{.3}{3}{13.5pt} \ ,\ \ioYBio{.3}{3}{13.5pt}\oiWoi{.6}{g-1}g{g-1}{6.75pt} \biggr) =P^{\mathrm{bwd}}_{u,v} \biggl( \oiWio{.6}{g}{g}{g}{6.75pt}\ioYBoi{.3}{3}{13.5pt} \ ,\ \ioYBoi{.3}{3}{13.5pt}\ioWoi{.6}{g}g{g}{6.75pt} \biggr) =1 ; \\ \label{B2.4}\tag{B2.4} & P^{\mathrm{bwd}}_{u,v} \biggl( \oiWii{.6}{g}{g}{g-1}{6.75pt}\iiYBii{.3}{3}{13.5pt} \ ,\ \ioYBio{.3}{3}{13.5pt}\oiWii{.6}{g}g{g-1}{6.75pt} \biggr) = 1- P^{\mathrm{bwd}}_{u,v} \biggl( \oiWii{.6}{g}{g}{g-1}{6.75pt}\iiYBii{.3}{3}{13.5pt} \ ,\ \ioYBoi{.3}{3}{13.5pt}\ioWii{.6}{g}{g-1}{g-1}{6.75pt} \biggr) =\frac{(1-t)v}{u-tv} \frac{u-st^g}{v-st^g} ; \\ \label{B3.1}\tag{B3.1} & P^{\mathrm{bwd}}_{u,v} \biggl( \ioWoo{.6}{g}{g}{g+1}{6.75pt}\ooYBoo{.3}{3}{13.5pt} \ ,\ \oiYBoi{.3}{3}{13.5pt}\ioWoo{.6}{g}g{g+1}{6.75pt} \biggr) =1- P^{\mathrm{bwd}}_{u,v} \biggl( \ioWoo{.6}{g}{g}{g+1}{6.75pt}\ooYBoo{.3}{3}{13.5pt} \ ,\ \oiYBio{.3}{3}{13.5pt}\oiWoo{.6}{g}{g+1}{g+1}{6.75pt} \biggr) =\frac{(1-t)v}{u-tv} \frac{1-st^gu}{1-st^gv} ; \\ \label{B3.2}\tag{B3.2} & P^{\mathrm{bwd}}_{u,v} \biggl( \ioWio{.6}{g+1}{g}{g+1}{6.75pt}\ioYBio{.3}{3}{13.5pt} \ ,\ \oiYBoi{.3}{3}{13.5pt}\ioWio{.6}{g+1}g{g+1}{6.75pt} \biggr) = P^{\mathrm{bwd}}_{u,v} \biggl( \ioWoi{.6}{g}{g}{g}{6.75pt}\oiYBio{.3}{3}{13.5pt} \ ,\ \oiYBio{.3}{3}{13.5pt}\oiWio{.6}{g}g{g}{6.75pt} \biggr) = 1 ; \\ \label{B3.3}\tag{B3.3} \begin{split} & P^{\mathrm{bwd}}_{u,v} \biggl( \ioWio{.6}{g+1}{g}{g+1}{6.75pt}\ioYBoi{.3}{3}{13.5pt} \ ,\ \oiYBio{.3}{3}{13.5pt}\oiWoi{.6}{g+1}{g+2}{g+1}{6.75pt} \biggr) =1 ; \\ & P^{\mathrm{bwd}}_{u,v} \biggl( \ioWoi{.6}{g}{g}{g}{6.75pt}\oiYBoi{.3}{3}{13.5pt} \ ,\ \oiYBoi{.3}{3}{13.5pt}\ioWoi{.6}{g}g{g}{6.75pt} \biggr) = 1- P^{\mathrm{bwd}}_{u,v} \biggl( \ioWoi{.6}{g}{g}{g}{6.75pt}\oiYBoi{.3}{3}{13.5pt} \ ,\ \oiYBio{.3}{3}{13.5pt}\oiWoi{.6}{g}{g+1}{g}{6.75pt} \biggr) = \frac{v-st^g}{u-st^g} \frac{1-st^gu}{1-st^gv} ; \end{split} \\ \label{B3.4}\tag{B3.4} & P^{\mathrm{bwd}}_{u,v} \biggl( \ioWii{.6}{g+1}{g}{g}{6.75pt}\iiYBii{.3}{3}{13.5pt} \ ,\ \oiYBoi{.3}{3}{13.5pt}\ioWii{.6}{g+1}g{g}{6.75pt} \biggr) = 1- P^{\mathrm{bwd}}_{u,v} \biggl( \ioWii{.6}{g+1}{g}{g}{6.75pt}\iiYBii{.3}{3}{13.5pt} \ ,\ \oiYBio{.3}{3}{13.5pt}\oiWii{.6}{g+1}{g+1}{g}{6.75pt} \biggr) = \frac{(1-t)u}{u-tv}\frac{v-st^g}{u-st^g} ; \\ \label{B4.1}\tag{B4.1} & P^{\mathrm{bwd}}_{u,v} \biggl( \iiWoo{.6}{g-1}{g}{g+1}{6.75pt}\ooYBoo{.3}{3}{13.5pt} \ ,\ \iiYBii{.3}{3}{13.5pt}\iiWoo{.6}{g-1}g{g+1}{6.75pt} \biggr) =1 ; \\ \label{B4.2}\tag{B4.2} & P^{\mathrm{bwd}}_{u,v} \biggl( \iiWio{.6}{g}{g}{g+1}{6.75pt}\ioYBio{.3}{3}{13.5pt} \ ,\ \iiYBii{.3}{3}{13.5pt}\iiWio{.6}{g}g{g+1}{6.75pt} \biggr) = P^{\mathrm{bwd}}_{u,v} \biggl( \iiWoi{.6}{g-1}{g}{g}{6.75pt}\oiYBio{.3}{3}{13.5pt} \ ,\ \iiYBii{.3}{3}{13.5pt}\iiWio{.6}{g-1}{g-1}{g}{6.75pt} \biggr) =1 ; \\ \label{B4.3}\tag{B4.3} & P^{\mathrm{bwd}}_{u,v} \biggl( \iiWoi{.6}{g-1}{g}{g}{6.75pt}\oiYBoi{.3}{3}{13.5pt} \ ,\ \iiYBii{.3}{3}{13.5pt}\iiWoi{.6}{g-1}g{g}{6.75pt} \biggr) = P^{\mathrm{bwd}}_{u,v} \biggl( \iiWio{.6}{g}{g}{g+1}{6.75pt}\ioYBoi{.3}{3}{13.5pt} \ ,\ \iiYBii{.3}{3}{13.5pt}\iiWoi{.6}{g}{g+1}{g+1}{6.75pt} \biggr) =1 ; \\ \label{B4.4}\tag{B4.4} & P^{\mathrm{bwd}}_{u,v} \biggl( \iiWii{.6}{g}{g}{g}{6.75pt}\iiYBii{.3}{3}{13.5pt} \ ,\ \iiYBii{.3}{3}{13.5pt}\iiWii{.6}{g}g{g}{6.75pt} \biggr) =1. %\end{noindent} \end{align} \section{Another form of the skew Cauchy identity} \label{sub:another_Cauchy} The spin Hall-Littlewood symmetric functions satisfy another form of Cauchy identities which is worth mentioning. These identities involve the functions $G^c$ (\Cref{ssub:G_definition}) along with the functions $G$. The latter are variations of the $F$ functions (\Cref{ssub:F_definition}), the only difference is that the boundary condition on the left as in in \Cref{fig:connecting_interlacing} (left) is also empty. We refer to \cite[Section 3]{Borodin2014vertex} for a detailed definition of the functions $G$. Let us focus on the variant of the skew Cauchy identity with single variables (analogue of \Cref{thm:skew_Cauchy_one}): \begin{proposition} \label{prop:anotForm} Under assumption \eqref{condition_on_convergence}, let $\lambda, \mu \in \mathsf{Sign}_N$. We have \begin{equation} \label{eq:anotForm} \sum_{\kappa \in \mathsf{Sign}_N} G^c_{\lambda/\kappa}(v^{-1})G_{\mu/\kappa} (u) =\sum_{\nu \in \mathsf{Sign}_N} G^c_{\nu/\mu} (v^{-1}) G_{\nu/\lambda} (u). \end{equation} \end{proposition} \begin{proof} The proof is analogous to out proof of \Cref{thm:skew_Cauchy_one} presented in \Cref{sub:bijective_proof_skew_Cauchy}. The only difference is that we consider boundary conditions as in \Cref{fig:another_skew_Cauchy} instead of \Cref{fig:skew_Cauchy}. Namely, one defines the modified transition probabilities on signatures $\bar{\mathsf{U}}^{\mathrm{fwd}}_{v,u} (\kappa \to\nu \mid \lambda, \mu)$ ($\bar{\mathsf{U}}^{\mathrm{bwd}}_{v,u} (\nu\to\kappa \mid \lambda, \mu)$) obtained by dragging the cross vertex $\oiYBio{.25}{3.5}{10.5pt}$ from $-\infty$ to $+\infty$ (from $+\infty$ to $-\infty$, respectively), proves an analog of \Cref{prop:reversibility_on_signatures} and obtains a bijective proof of \eqref{eq:anotForm}. \end{proof} \begin{figure}[htpb] \centering \begin{tikzpicture}[scale=.6, thick] %\begin{noindent} \draw (-4.5,0)--++(10.5,0) node[below right] {$v$}; \draw (-4.5,1)--++(10.5,0) node[above right] {$u$}; \draw[densely dotted, line width=.4] (-4.5,2)--++(10.2,0); \draw[densely dotted, line width=.4] (-4.5,-1)--++(10.2,0); \foreach \ii in {-4,...,5} { \draw (\ii,1.1)--++(0,-.2); \draw (\ii,.1)--++(0,-.2) node[below, yshift=-25] {$\ii$}; \draw[densely dotted, line width=.4] (\ii,-1.3)--++(0,3.6); } \node at (0.5,-1.1) {$\lambda$}; \draw [line width=2,->] (-2,-1)--++(0,1); \draw [line width=2,->] (.9,-1)--++(0,1); \draw [line width=2,->] (1.1,-1)--++(0,1); \draw [line width=2,->] (3,-1)--++(0,1); \node at (0.5,2.1) {$\mu$}; \draw [line width=2,->] (-1,1)--++(0,1); \draw [line width=2,->] (0,1)--++(0,1); \draw [line width=2,->] (1,1)--++(0,1); \draw [line width=2,->] (4,1)--++(0,1); \node at (.5,0.5) {$\kappa$}; \draw [line width=2,->] (-3,0)--++(0,.9); \draw [line width=2,->] (0,0)--++(0,1); \draw [line width=2,->] (2,0)--++(0,1); \draw [line width=2,->] (-5,0)--++(1,0); \draw [line width=2,->] (-4,0)--++(1,0); \draw [line width=2,->] (-3,.9)--++(.1,.1)--++(.9,0); \draw [line width=2,->] (-2,1)--++(1,0); \draw [line width=2,->] (-2,0)--++(1,0); \draw [line width=2,->] (-1,0)--++(1,0); \draw [line width=2,->] (.9,0)--++(.1,.1)--++(0,.9); \draw [line width=2,->] (1.1,0)--++(.9,0); \draw [line width=2,->] (2,1)--++(1,0); \draw [line width=2,->] (3,1)--++(1,0); \draw [line width=2,->] (3,0)--++(1,0); \draw [line width=2,->] (4,0)--++(1,0); \draw [line width=2,->] (5,0)--++(1,0); \draw [line width=2,->] (6,0)--++(1,0); \end{tikzpicture}\qquad \quad \begin{tikzpicture}[scale=.6, thick] \draw (-4.5,0)--++(10.5,0) node[below right] {$u$}; \draw (-4.5,1)--++(10.5,0) node[above right] {$v$}; \draw[densely dotted, line width=.4] (-4.5,2)--++(10.2,0); \draw[densely dotted, line width=.4] (-4.5,-1)--++(10.2,0); \foreach \ii in {-4,...,5} { \draw (\ii,1.1)--++(0,-.2); \draw (\ii,.1)--++(0,-.2) node[below, yshift=-25] {$\ii$}; \draw[densely dotted, line width=.4] (\ii,-1.3)--++(0,3.6); } \node at (0.5,-1.1) {$\lambda$}; \draw [line width=2,->] (-2,-1)--++(0,.9); \draw [line width=2,->] (.9,-1)--++(0,1); \draw [line width=2,->] (1.1,-1)--++(0,1); \draw [line width=2,->] (3,-1)--++(0,1); \node at (0.5,2.1) {$\mu$}; \draw [line width=2,->] (-1.1,1)--++(.1,.1)--++(0,.9); \draw [line width=2,->] (0,1)--++(0,1); \draw [line width=2,->] (1,1)--++(0,1); \draw [line width=2,->] (4,1)--++(0,1); \node at (.5,0.5) {$\nu$}; \draw [line width=2,->] (-5,1)--++(1,0); \draw [line width=2,->] (-4,1)--++(1,0); \draw [line width=2,->] (-3,1)--++(1,0); \draw [line width=2,->] (-2,0)--++(1,0); \draw [line width=2,->] (-2,1)--++(1,0); \draw [line width=2,->] (-1,0)--++(0,.9); \draw [line width=2,->] (-1,.9)--++(.1,.1)--++(.9,0); \draw [line width=2,->] (.9,0)--++(.1,.1)--++(0,.9); \draw [line width=2,->] (1.1,0)--++(.9,0); \draw [line width=2,->] (2,0)--++(0,1); \draw [line width=2,->] (2,1)--++(1,0); \draw [line width=2,->] (3,1)--++(1,0); \draw [line width=2,->] (3,0)--++(1,0); \draw [line width=2,->] (4,0)--++(1,0); \draw [line width=2,->] (5,0)--++(0,1); \draw [line width=2,->] (5,1)--++(1,0); \draw [line width=2,->] (6,1)--++(1,0); \end{tikzpicture} \caption{Illustration of the sums in both sides of identity \eqref{eq:anotForm}.} \label{fig:another_skew_Cauchy} %\end{noindent} \end{figure} \Cref{prop:anotForm} is new. Its $s=0$ degeneration was mentioned in \cite[Sections 3.1 and 3.7]{BufetovMatveev2017}. The significance of this variation of the skew Cauchy identity is in the fact that it does not have any prefactors, which is neat from the combinatorial point of view. Another property which is better visible in this variation is a symmetry between $\lambda$ and $\mu$: \begin{proposition} \label{prop:symmetrF} Let $\bar{\mathsf{U}}^{\mathrm{fwd}}_{v,u}(\kappa \to \nu \mid \lambda, \mu)$ and $\bar{\mathsf{U}}^{\mathrm{bwd}}_{v,u}(\nu\to\kappa\mid\lambda,\mu)$ be transition probabilities defined in the proof of Proposition \ref{prop:anotForm}. We have \begin{equation*} \bar{\mathsf{U}}^{\mathrm{fwd}}_{v,u} (\kappa \to \nu \mid \lambda, \mu) = \bar{\mathsf{U}}^{\mathrm{fwd}}_{u^{-1},v^{-1}}(\kappa \to \nu \mid \mu, \lambda), \qquad \bar{\mathsf{U}}^{\mathrm{bwd}}_{v,u} (\nu \to \kappa \mid \lambda, \mu) = \bar{\mathsf{U}}^{\mathrm{bwd}}_{u^{-1},v^{-1}} (\nu \to \kappa \mid \mu, \lambda) . \end{equation*} \end{proposition} \begin{proof} Readily follows from Proposition 3.4. \end{proof} Note also that in the Hall-Littlewood case ($s=0$) \Cref{prop:symmetrF} becomes fully symmetric: \begin{equation*} \bar{\mathsf{U}}^{\mathrm{fwd}}_{v,u} (\kappa \to \nu \mid \lambda, \mu) = \bar{\mathsf{U}}^{\mathrm{fwd}}_{v,u}(\kappa \to \nu \mid \mu, \lambda), \qquad \bar{\mathsf{U}}^{\mathrm{bwd}}_{v,u} (\nu \to \kappa \mid \lambda, \mu) = \bar{\mathsf{U}}^{\mathrm{bwd}}_{v,u} (\nu \to \kappa \mid \mu, \lambda) . \end{equation*} Indeed, this is because the local transition probabilities (\Cref{fig:fwd_YB,fig:bwd_YB}) are invariant under the swap $(u,v) \to (v^{-1},u^{-1})$ if $s=0$. \section{Inhomogeneous modifications} \label{app:inhomogeneous_construction} Most constructions and results of the present paper can be generalized to allow the spectral parameter $u$ and the spin parameter $s$ in the higher spin weights \eqref{vertex_weights} vary along columns. Versions of the spin Hall-Littlewood functions $F$ and $G^c$ with this type of inhomogeneity, as well as Cauchy summation identities for these functions, are discussed in detail in \cite{BorodinPetrov2016inhom}. Such Cauchy identities were employed in that work to compute observables of the inhomogeneous stochastic higher spin six vertex model which are amenable to asymptotic analysis (performed in, e.g., \cite{BorodinPetrov2016Exp}). Let us briefly discuss the modifications needed to introduce inhomogeneity parameters into our constructions. These parameters form two families, $\{\xi_i\}_{i\in\mathbb{Z}}$ and $\{s_i\}_{i\in\mathbb{Z}}$. The vertex weights \eqref{vertex_weights} in the column number $i$ now depend on the parameters $\xi_i u$ (replacing $u$) and $s_i$. These parameters $\xi_i,s_i$ do not enter the cross vertex weights \eqref{cross_vertex_weights} involved in the Yang-Baxter equation. However, they do enter the local transition probabilities $P^{\mathrm{fwd}},P^{\mathrm{bwd}}$: in the tables in \Cref{fig:fwd_YB,fig:bwd_YB} one should replace the parameters $u,v,s$ with $\xi_iu,\xi_iv,s_i$, respectively, where $i\in \mathbb{Z}$ is the location through which the cross vertex is dragged. Next, the definitions of the functions $F$ and $G^c$ should be modified as in \cite{BorodinPetrov2016inhom}, by first replacing $(u,s)\to(\xi_m u,s_m)$ in \eqref{F_skew_one_variable_definition} and $(u^{-1},s)\to(\xi_m^{-1}u^{-1},s_m)$ in \eqref{G_skew_one_variable_definition}, and then defining the multivariable functions as in \Cref{ssub:F_G_multivar_definition}. Note that in Cauchy identities (e.g., in \eqref{nonskew_multi_Cauchy}) the parameters in the functions $F$ and $G^c$ should be $u_i\xi_m$ and $v_j^{-1}\xi_m^{-1}$, respectively. Remarkably, the double product $\prod\prod\frac{v_j-u_i}{v_j-tu_i}$ entering \eqref{nonskew_multi_Cauchy} remains the same in the inhomogeneous setting. Having inhomogeneous versions of the spin Hall-Littlewood functions $F$ and $G^c$, one can define the corresponding measures and processes as in \Cref{sub:spin_HL_measures_processes}. The local transition probabilities assembled into $\mathsf{U}^{\mathrm{fwd}}_{v,u}$ and $\mathsf{U}^{\mathrm{bwd}}_{v,u}$ thus give rise to an inhomogeneous version of the Yang-Baxter field depending on $t$, the parameters $\{u_i\}$, $\{v_j\}$ as in \Cref{fig:YB_field}, and two series of inhomogeneous parameters $\{\xi_m\}$ and $\{s_m\}$. The latter parameters may be thought of as belonging to the third dimension in \Cref{fig:YB_field}, the one where the signatures $\boldsymbol\lambda^{(x,y)}$ live. The dynamic stochastic six vertex model (DS6V) arising in \Cref{sec:dynamicS6V} as a Markov projection of the Yang-Baxter field onto the column number zero does not feel the inhomogeneous parameters $\{\xi_m\}$ and $\{s_m\}$ for $m\ge1$. This follows by the very construction of the Yang-Baxter field using the probabilities $\mathsf{U}^{\mathrm{fwd}}_{v,u}$. In other words: \begin{corollary} \label{cor:independence_on_inhom_parameters} The distribution of the number of zero parts $\lambda^{[0]}$ under the inhomogeneous version of the spin Hall-Littlewood measure described above does not depend on the inhomogeneity parameters $\xi_m,s_m$ for $m\ge1$. A similar statement holds for spin Hall-Littlewood processes. \end{corollary} On the other hand, the parameters $\{u_i\}$, $\{v_j\}$ entering the Yang-Baxter field, carry over to the DS6V model. The height function in this inhomogeneous DS6V model is identified with $\lambda^{[0]}$ under a spin Hall-Littlewood measure, in which the inhomogeneous parameters $u_i,v_j$ serve as variables in the functions $F$ and $G^c$. See \Cref{cor:dyn6V_spin_HL_process}. The presence of the inhomogeneous parameters $\{u_i\}$ and $\{v_j\}$ carries over to most of the degenerations of the DS6V model considered in \Cref{sec:degenerations}. An exception is the ASEP type limit of \Cref{sub:degen_ASEP} since this limit is performed along the diagonal of the quadrant. \printbibliography \end{document}