\documentclass[12pt]{amsart} \usepackage{amsfonts} \usepackage{amsmath} \usepackage{amssymb} \usepackage[mathscr]{eucal} \usepackage{graphicx} \usepackage[headings]{fullpage} \usepackage{times} %\usepackage{bbm} \usepackage{epsfig} \usepackage{color} \usepackage{hyperref} %\usepackage[dvipsnames]{xcolor} %\definecolor{ao}{rgb}{0.0, 0.5, 0.0} %\definecolor{darkmagenta}{rgb}{0.55, 0.0, 0.55} \newcommand{\overbar}[1]{\mkern 1.5mu\overline{\mkern-1.5mu#1\mkern-1.5mu}\mkern 1.5mu} \def\eod{\vrule height 6pt width 5pt depth 0pt} \newtheorem{theorem}{Theorem} \newtheorem{lemma}[theorem]{Lemma} \newtheorem{definition}[theorem]{Definition} \newtheorem{corollary}[theorem]{Corollary} \newtheorem{proposition}[theorem]{Proposition} \newtheorem{fact}[theorem]{Fact} \newtheorem{observation}[theorem]{Observation} \newtheorem{claim}[theorem]{Claim} \newtheorem{example}[theorem]{Example} \newtheorem{remark}[theorem]{Remark} \newtheorem{problem}[theorem]{Problem} \newcommand{\sB}{\mathcal{B}} \newcommand{\sT}{\mathcal{T}} \newcommand{\pfaff}{ \mathsf{Pfaff}} \newcommand{\diag}{ \mathsf{Diag}} \newcommand{\ED}{ \mathsf{ED}} \newcommand{\MED}{ \mathsf{MED}} \newcommand{\oS}{ \overline{ S}} \newcommand{\od}{ \overline{ d}} \newcommand{\of}{ \overline{ f}} \newcommand{\oc}{ \overline{c}} \newcommand{\sD}{ \mathcal{ D}} \newcommand{\sL}{\mathcal{ L}} \newcommand{\sM}{ \mathcal{ M}} \newcommand{\sO}{ \mathcal{ O}} \newcommand{\bone}{ \mathbbm{1}} \newcommand{\bz}{ \mathbb{ 0}} \newcommand{\RR}{ \mathbb{R}} \newcommand{\ZZ}{ \mathbb{Z}} \newcommand{\QQ}{ \mathbb{Q}} \newcommand{\vv}{ \mathbbm{v}} \newcommand{\lm}{ \mathsf{lexmin}} \newcommand{\revpath}{ \mathsf{reverse\_on\_path}} \newcommand{\vsp}{\vskip 1em} \newcommand{\OB}{ \mathsf{OB}} \newcommand{\ch}{ \mathsf{ch}} \newcommand{\cv}{ \mathsf{cv}} \newcommand{\CF}{ \mathsf{CF}} \newcommand{\vol}{ \mathsf{vol}} \newcommand{\uu}{ \mathbbm{u}} \newcommand{\ob}{ \mathbbm{b}} \newcommand{\rhalf}{\lfloor r/2 \rfloor} \newcommand{\nhalf}{\lfloor n/2 \rfloor} \newcommand{\Or}{\mathsf{Orient}} \newcommand{\AOr}{\mathsf{AwayOrient}} \newcommand{\rows}{ \mathsf{rows}} \newcommand{\cyc}{ \mathsf{cycles}} \newcommand{\cols}{ \mathsf{cols}} \newcommand{\Row}{ \mathsf{Row}} \newcommand{\Col}{ \mathsf{Col}} \newcommand{\comment}[1]{} \newcommand{\floor}[1]{\lfloor #1 \rfloor} \newcommand{\cofsum}{\mathsf{cofsum}} \newcommand{\fp}{ \mathsf{fp}} \newcommand{\rar}{ \rightarrow} \newcommand{\SSS}{\mathfrak{S}} \newcommand{\DD}{ \mathfrak{D}} \newcommand{\CC}{ \mathbb{C}} \newcommand{\NN}{ \mathbb{N}} \newcommand{\EDb}{ \overline{ \mathsf{ED}}} \newcommand{\dt}{\mathsf{det2}} \newcommand{\aw}{\mathsf{away}} \newcommand{\GTS}{\mathsf{GTS}} \newcommand{\law}{\mathsf{Lexaway}} \newcommand{\awy}{\mathsf{Aw}} \newcommand{\BT}{\mathsf{BT}} \newcommand{\wt}{\mathsf{wt}} \newcommand{\dtt}{\mathsf{det3}} \newcommand{\id}{\mathsf{id}} \newcommand{\bd}{\mathsf{bidir}} \newcommand{\fr}{\mathsf{free}} \newcommand{\sgn}{ \mathsf{sgn}} \newcommand{\perm}{ \mathsf{perm}} \newcommand{\Tr}{ \mathsf{Trace}} \newcommand{\la}{ \langle} \newcommand{\ra}{ \rangle} \newcommand{\row}{ \mathsf{Row}} \newcommand{\type}{ \mathsf{CycleType}} \newcommand{\col}{ \mathsf{Col}} \newcommand{\obr}[1]{\overbar{#1}} \newcommand{\otalpha}{\widetilde{\alpha} } %\DeclarePairedDelimiter{\ceil}{\lceil}{\rceil} \newcommand{\sF}{ \mathcal{ F}} \newcommand{\quotes}[1]{``#1''} \newcommand{\hsp}{\hspace{10mm}} \newcommand{\red}[1]{\textcolor{red}{#1}} %\newcommand{\magenta}[1]{\textcolor{magenta}{#1}} \newcommand{\magenta}[1]{#1} \newcommand{\green}[1]{\textcolor{green}{#1}} \newcommand{\purple}[1]{\textcolor{purple}{#1}} \newcommand{\cyan}[1]{\textcolor{cyan}{#1}} %\newcommand{\blue}[1]{\textcolor{blue}{#1}} \newcommand{\blue}[1]{#1} \newcommand{\ao}[1]{\textcolor{ao}{#1}} \newcommand{\darkmagenta}[1]{\textcolor{darkmagenta}{#1}} \begin{document} \title[Generalized Matrix polynomials of Tree Laplacians]{Generalized Matrix polynomials of Tree Laplacians indexed by Symmetric functions and the $\GTS$ poset} \author{Mukesh Kumar Nagar} \address {Department of Mathematics and Statistics\newline \indent Indian Institute of Technology Kanpur \newline \indent Kanpur 208 016, India} \email{mukesh.kr.nagar@gmail.com} \author{Sivaramakrishnan Sivasubramanian} \address {Department of Mathematics \newline \indent Indian Institute of Technology Bombay \newline \indent Mumbai 400 076, India} \email{krishnan@math.iitb.ac.in} \begin{abstract} Let $T$ be a tree on $n$ vertices with Laplacian matrix $L_T$ and $q$-Laplacian $\sL_T^q$. Let $\GTS_n$ be the generalized tree shift poset on the set of unlabelled trees on $n$ vertices. Inequalities are known between coefficients of the immanantal polynomial of $L_T$ and $\sL_T^q$ as one moves up the poset $\GTS_n$. Using the Frobenius characteristic, this can be thought as a result involving the Schur symmetric function $s_{\lambda}$. In this paper, we use an arbitrary symmetric function to define a {\it generalized matrix function} of an $n \times n$ matrix. When the symmetric function is the monomial and the forgotten symmetric function, we generalize such inequalities among coefficients of the generalized matrix polynomial of $\sL_T^q$ as one moves up the $\GTS_n$ poset. \end{abstract} \keywords {Tree, $\GTS_n$ poset, Laplacian, monomial symmetric function, generalized matrix polynomial} \subjclass[2010]{05C05, 06A06, 15A15} \maketitle \thispagestyle{myheadings} \font\rms=cmr8 \font\its=cmti8 \font\bfs=cmbx8 \markright{\its S\'eminaire Lotharingien de Combinatoire \bfs 83 \rms (2021), Article~B83a\hfill} %\def\thepage{} \section{Introduction} \label{sec:intro} For a positive integer $n$, let $[n]=\{1,2,\ldots,n\}$ and let $\SSS_n$ denote the symmetric group on the set $[n]$. We denote partitions $\lambda$ of the number $n$ as $\lambda \vdash n$. We write partitions using the exponential notation, with multiplicities of parts written as exponents. For $\lambda \vdash n$, let $\chi_{\lambda}^{}$ be the irreducible character of $\SSS_n$ over $\CC$ indexed by $\lambda$. We think of $\chi_{\lambda}^{}$ as a function $\chi_{\lambda}^{} : \SSS_n \mapsto \ZZ$. With respect to an irreducible character $\chi_{\lambda}^{}$, define the immanant of the $n \times n$ matrix $A = (a_{i,j})_{1 \leq i,j \leq n}$ as \begin{equation} \label{eqn:immanant} d_{\lambda}(A) = \sum_{\psi \in \SSS_n} \chi_{\lambda}^{}(\psi) \prod_{i=1}^n a_{i,\psi(i)}. \end{equation} Let $\Lambda_{\QQ}^n$ denote the vector space of degree $n$ symmetric functions with coefficients from $\QQ$. The set of monomial symmetric functions $\{m_{\lambda} \}_{\lambda\vdash n}$ is one of the well known bases of $\Lambda_{\QQ}^n$. We refer the reader to the books by Stanley \cite{EC2} and by Mendes and Remmel \cite{mendes-remmel-book} for background on symmetric functions. $\Lambda_{\QQ}^n$ actually has an inner product structure as well. Another inner product space often studied is $\CF_n$, the space of class functions from $\SSS_n \mapsto \QQ$. Further, there is a well known isometry between these two spaces called the Frobenius characteristic, denoted $\ch: \CF_n \rightarrow \Lambda_{\QQ}^n$ (see Stanley \cite{EC2} for more details). Let $\gamma \in \Lambda_{\QQ}^n$ and consider $\Gamma_{\gamma} = \ch^{-1}(\gamma)$ its inverse Frobenius image. Clearly, $\Gamma_{\gamma} \in \CF_n$ is a class function indexed by $\gamma$. Define the generalized matrix function (GMF henceforth) of an $n \times n$ matrix $A$ with respect to $\gamma$ as \begin{equation} \label{eqn:gen_matrix_fn} d_{\gamma}(A) = \sum_{\psi \in \SSS_n} \Gamma_{\gamma}(\psi) \prod_{i=1}^n a_{i,\psi(i)}. \end{equation} Define the generalized matrix polynomial $\zeta_{\gamma}^A(x)$ of $A$ in a new variable $x$ with respect to a symmetric function $\gamma$ as follows: $\zeta_{\gamma}^A(x)=d_{\gamma}(xI-A)$. It is well known that the inverse Frobenius image $\ch^{-1}(s_{\lambda})$ of the Schur symmetric function $s_{\lambda}$ is $\chi_{\lambda}^{}$, the irreducible character of $\SSS_n$ over $\CC$ indexed by $\lambda$ (see \cite{EC2}). Thus, from \eqref{eqn:immanant} and \eqref{eqn:gen_matrix_fn}, we can see that $d_{s_{\lambda}}(A)=d_{\lambda}(A)$ and $\zeta_{s_{\lambda}}^A(x)=d_{\lambda}(xI-A)$. %First page headline in AmS-LaTeX for S\'eminaire Lotharingien de Combinatoire %--restoring the headers and pagenumbering \pagenumbering{arabic} \addtocounter{page}{1} \markboth{\SMALL MUKESH KUMAR NAGAR AND SIVARAMAKRISHNAN SIVASUBRAMANIAN}{\SMALL GENERALIZED MATRIX POLYNOMIALS OF TREE LAPLACIANS} Csikv{\'a}ri in \cite{csikvari-poset1} defined a poset on the set of unlabelled trees with $n$ vertices that we denote in this paper as $\GTS_n$. Among other results, he showed that if one moves up along the $\GTS_n$ poset, all coefficients of the characteristic polynomial of the Laplacian matrix $L_T$ of a tree $T$ decrease in absolute value. This result was generalized by Nagar and Sivasubramanian in \cite{mukesh-siva-gts} to immanantal polynomials of $\sL_T^q$ indexed by any $\lambda \vdash n$, where $\sL_T^q$ is the $q$-Laplacian matrix of $T$ (see Theorem~\ref{thm:main_earlier}). Let $\RR^+$ denote the set of non-negative real numbers and $\RR^+[q^2]$ denote the set of polynomials in $q^2$ with coefficients in $\RR^+.$ Let $m_{\lambda}\in \Lambda_{\QQ}^n$ be the monomial symmetric function indexed by $\lambda \vdash n$. In Section~\ref{sec:gmfs_monomial} of this paper, we prove the following result which shows monotonicity of the coefficient of $(-1)^rx^{n-r}$ for $0 \leq r \leq n$ in $\zeta_{m_{\lambda}}^{\sL_{T}^q}(x)$ when we move up along $\GTS_n$. \begin{theorem} \label{thm:main} Let $T_1$ and $T_2$ be two trees with $n$ vertices such that $T_2$ covers $T_1$ in $\GTS_n$. Let $\sL_{T_1}^q$ and $\sL_{T_2}^q$ be the $q$-Laplacians of $T_1$ and $T_2$ respectively. For $\lambda \vdash n$, let \begin{align*} \zeta^{\sL_{T_1}^q}_{m_{\lambda}}(x) &= d_{m_{\lambda}}(xI -\sL_{T_1}^q) = \sum_{r=0}^n (-1)^r c_{m_{\lambda},r}^{\sL_{T_1}^q}(q) x^{n-r} \mbox{ and} \\ \zeta^{\sL_{T_2}^q}_{m_{\lambda}}(x) & = d_{m_{\lambda}}(xI -\sL_{T_2}^q) = \sum_{r=0}^n (-1)^r c_{m_{\lambda},r}^{\sL_{T_2}^q}(q) x^{n-r}. \end{align*} Then for all $\lambda \vdash n$, we have $c_{m_{\lambda},r}^{\sL_{T_1}^q}(q) - c_{m_{\lambda},r}^{\sL_{T_2}^q}(q) \in \RR^+[q^2]$, where $0 \leq r \leq n$. Further if $\lambda \neq 2^k,1^{n-2k}\vdash n$, then, $d_{m_{\lambda}}(xI -\sL_{T_1}^q)=d_{m_{\lambda}}(xI -\sL_{T_2}^q)=0.$ \end{theorem} Recall that for $\gamma \in \Lambda_{\QQ}^n$, $\Gamma_{\gamma} = \ch^{-1}(\gamma)$. For the proof of Theorem~\ref{thm:main}, we show the following lemma involving $\Gamma_{m_{\lambda}}=\ch^{-1}(m_{\lambda})$ and binomial coefficients. Let $\Gamma_{\gamma}(j)$ denote the class function $\Gamma_{\gamma}(\cdot)$ evaluated at a permutation $\psi \in \SSS_n$ with cycle type $2^j,1^{n-2j}$. For $0 \leq i \leq \nhalf$, define \begin{equation} \label{eqn:main} \alpha_i(\gamma) = \sum_{j=0}^i \binom{i}{j} \Gamma_{\gamma}(j). \end{equation} In Section~\ref{sec:gmfs_monomial}, we prove the following lemma which we believe is of independent interest. \begin{lemma} \label{lem:frob_inv_monomial} For all $\lambda \vdash n$ and for $0\leq i \leq \nhalf$, %the quantity %$\alpha_{i}^{}(m_{\lambda})$ is a non-negative integral multiple of $2^i$. %Moreover for $0\leq i \leq \nhalf$, $\alpha_{i}^{}(m_{\lambda}) = 2^i$ if $\lambda=2^i,1^{n-2i}$ and $0$ otherwise. \end{lemma} In Section~\ref{sec:gmfs_forgotten_symm_fn} we consider the generalized matrix polynomial of $\sL_T^q$ with respect to the {\it forgotten symmetric function}. Our main result of that section is Theorem~\ref{thm:main_f_lambda}, where we show similar monotonicity results as we move on the $\GTS_n$ poset. \section{Preliminaries} \label{sec:prelims} We now give some motivational background for our results and place it in its context. The normalized immanant of a matrix $A$ corresponding to a partition $\lambda$ is defined as $\displaystyle \od_{\lambda}(A) = \frac{d_{\lambda}(A)}{ \chi_{\lambda}^{}(\id)}$. Here, $\chi_{\lambda}^{}(\id)$ equals the dimension of the irreducible representation of $\SSS_n$ over $\CC$ indexed by $\lambda$. Schur \cite{schur-immanant-ineqs} showed that among normalized immanants, the determinant is the smallest normalized immanant for any positive semidefinite Hermitian matrix. This shows that for any positive semidefinite Hermitian matrix $A$, all its normalized immanants (and hence immanants) are non-negative. Though the immanant $d_{\lambda}(A)$ is non-negative, \eqref{eqn:immanant} is not a non-negative expression for $d_{\lambda}(A)$ as $\chi_{\lambda}^{}(\psi) \prod_{i=1}^n a_{i,\psi(i)}$ is not necessarily non-negative for all $\psi \in \SSS_n$ that contribute to $d_{\lambda}(A)$. Recall that the Laplacian matrix $L_G$ of a graph $G$ is defined as $L_G = D-A$, where $D$ is the diagonal matrix with degrees of $G$ on the diagonal and $A$ is the adjacency matrix of $G$. It is well known that $L_G$ is positive semidefinite for all graphs $G$ (see \cite{godsil-royle}). When the matrix $A$ is the Laplacian $L_T$ of a tree $T$, then Chan and Lam in \cite{hook_immanant_explained-chan_lam} gave the following two results which give an alternate positive expression for the immanant $d_{\lambda}(L_T)$. We denote the number of parts of a partition $\lambda$ of $n$ as $l(\lambda)$. For $\lambda \vdash n$, let $\chi_{\lambda}^{}(j)$ denote the character value $\chi_{\lambda}^{}(\cdot)$ evaluated at a permutation $\psi \in \SSS_n$ with cycle type $2^j, 1^{n-2j}$. For $0 \leq i \leq \nhalf$ and $\lambda \vdash n$, define \begin{equation} \alpha_{i,\lambda} = \sum_{j=0}^i \binom{i}{j} \chi_{\lambda}^{}(j). \label{eqn:main_char} \end{equation} \begin{lemma}[\sc Chan and Lam, \cite{chan-lam-binom-coeffs-char}] \label{lem:chan_lam_bino_char} For all $\lambda \vdash n$ and for $0\leq i \leq \nhalf$, the quantity $\alpha_{i, \lambda}$ is a non-negative integral multiple of $2^i$. Moreover for $1\leq i \leq \nhalf$, $\alpha_{i,\lambda}=0$ if and only if $l(\lambda)>n-i$. \end{lemma} \begin{theorem}[\sc Chan and Lam, \cite{hook_immanant_explained-chan_lam}] \label{thm:positive_immanant} Let $L_T$ be the Laplacian matrix of a tree $T$ on $n$ vertices. Then, for all $\lambda \vdash n$, $d_{\lambda}(L_T) = \sum_{i=0}^{\nhalf} \alpha_{i, \lambda} a_i(T)$, where $a_i(T)$ equals the number of vertex orientations with exactly $i$ bidirected edges (and is hence a non-negative integer for all $i$). \end{theorem} Lemma~\ref{lem:chan_lam_bino_char} and Theorem~\ref{thm:positive_immanant} make it clear that all immanants of $L_T$ are non-negative. Similar results are known for the $q$-Laplacian of $T$ (see Theorem~\ref{thm:mukesh-siva-positive_immanant}). Define $\sL_G^q = I+q^2(D-I) -qA$ as the $q$-Laplacian of a graph $G$, where $D$ and $A$ are as before and $I$ is the identity matrix. Here $q$ is a variable. It is easy to see that setting $q=1$ in $\sL_G^q$ gives us the usual combinatorial Laplacian $L_G$. The matrix $\sL_G^q$ has appeared in the contexts of Ihara--Selberg zeta functions of graphs $G$ (see Bass \cite{bass} and Foata and Zeilberger \cite{foata-zeilberger-bass-trams}). When the graph $G$ is a tree $T$, $\sL_T^q$ has connections with the inverse of the exponential distance matrix of $T$ (see Bapat, Lal and Pati \cite{bapat-lal-pati} and Nagar \cite{nagar}). Nagar and Sivasubramanian in \cite{mukesh-siva-hook} gave the following $q$-analogue of Theorem~\ref{thm:positive_immanant} involving %hereby generalising Theorem %\ref{thm:positive_immanant} to $\sL_T^q$. To state it, we need a $q$-analogue of the term $a_i(T)$. We describe it briefly and refer the reader to \cite{mukesh-siva-hook} for more details. Let $\sO_i(T)$ denote the set of vertex orientations in $T$ with $i$ bidirected edges, and let $a_i(T) = |\sO_i(T)|$. A statistic $\law: \sO_i(T) \mapsto \ZZ_{\geq 0}$ was defined in \cite{mukesh-siva-hook}. Using it, when $i \geq 1$, the following $q$-analogue $a_i(T,q)$ of $a_i(T)$ was defined (see \cite[Corollary~9]{mukesh-siva-hook}): \begin{equation} \label{eqn:q-analog-vertex-orient} a_i(T,q) = \sum_{O \in \sO_i} q^{\law(O)}. \end{equation} In \cite{mukesh-siva-hook}, we had also defined $a_0(T,q) = 1-q^2$. It turns out that for all $i \geq 0$, the $a_i(T,q)$ is a polynomial in $q^2$ (see \cite[Remark~15]{mukesh-siva-hook}). With this definition, we can state a $q$-analogue of Theorem~\ref{thm:positive_immanant}. \begin{theorem}[\sc Nagar and Sivasubramanian] \label{thm:mukesh-siva-positive_immanant} Let $\sL_T^q$ be the $q$-Laplacian matrix of a tree $T$ on $n$ vertices. Then, for all $\lambda \vdash n$, $d_{\lambda}(\sL_T^q) = \sum_{i=0}^{\nhalf} \alpha_{i, \lambda} a_i(T,q)$, where the $a_i(T,q)$ is defined in \eqref{eqn:q-analog-vertex-orient}. %is a polynomial in $q^2$, counting vertex % orientations with exactly $i$ bidirected edges with respect to a % statistic $\law(\cdot)$. \end{theorem} Clearly, setting $q=1$ in $\sL_T^q$ gives $L_T$ and setting $q=1$ in $a_i(T,q)$ gives $a_i(T)$ for $i$ with $0\leq i \leq \nhalf$. %(see \cite{mukesh-siva-hook} for the definitions). %Further, since $a_i(T,q)$ is a polynomial %in $q^2$, Theorem %\ref{thm:mukesh-siva-positive_immanant} holds for %all $q \in \RR$. Extensions of Theorem~\ref{thm:mukesh-siva-positive_immanant} to the bivariate $q,t$-Laplacian denoted as $\sL_T^{q,t}$ were also given in \cite{mukesh-siva-hook}. %and a counterpart of Theorem %\ref{thm:mukesh-siva-positive_immanant} holds even when %$q,t \in \RR$ with $qt > 0$ and when $q,t \in \CC$ with $qt > 0$. Later, in \cite{mukesh-siva-gts}, the following more general result about the coefficient of $(-1)^rx^{n-r}$ in the immanantal polynomial $d_{\lambda}(xI-\sL_T^q)$ was proved. To state it, we need polynomials $a_{i,r}^{}(T,q)$ which generalize $a_i(T,q)$. Let $B \subset V(T)$ be a subset of the vertex set of $T$ with $|B| = r$, where $r \leq n$. %is a polynomial in $q^2$, Let $\sO_{B,i}^T$ be the set of vertex orientations of vertices in $B$ that have $i$ bidirected edges. There is a statistic $\awy_B^T: \sO_{B,i}^T \mapsto \ZZ_{\geq 0}$ with respect to which we define \begin{equation} \label{eqn:defn_a_riT} a_{i,r}^{}(T,q) = \sum_{B \subseteq V(T), |B|=r} \sum_{O \in \sO_{B,i}^T}q^{\awy(_B^T(O)}. \end{equation} With this definition, we have the following \cite[Lemma~7, Corollary~12]{mukesh-siva-gts}. %which counts vertex orientations summed over vertex subsets $B$ having %exactly $r$ vertices in $T$ and with exactly $i$ bidirected %edges with respect to a statistic $\awy_B^T(\cdot)$. \begin{lemma}[\sc Nagar and Sivasubramanian] \label{lem:coeff_imm_poly} Let $T$ be a tree on $n$ vertices with $q$-Laplacian $\sL_T^q$. For $\lambda \vdash n$, let $d_{\lambda}(xI-\sL_T^q)=\sum_{r=0}^n (-1)^r c_{\lambda,r}^{\sL_{T}^q}(q) x^{n-r}$. Then for $0\leq r \leq n$, we have $c_{\lambda,r}^{\sL_{T}^q}(q) =\sum_{i=0}^{\rhalf}\alpha_{i,\lambda}^{} a_{i,r}^{}(T,q)$. %where $a_{i,r}^{}(T,q)$ is a polynomial in $q^2$, counting vertex %orientations on all vertex subsets $B$ having exactly $r$ vertices %in $T$ and with exactly $i$ bidirected edges with respect to a %statistic $\awy_B^T(\cdot)$. \end{lemma} In Lemma~\ref{lem:coeff_imm_poly}, one can get the immanant $d_{\lambda}(\sL_q^T)$ as the constant term of the immanantal polynomial (that is, by looking at the special case when $r=n$). %in $ a_{i,r}^{}(T,q)$ clearly %gives $ a_{i}^{}(T,q)$ for all $i$ (see \cite{mukesh-siva-gts} %for the definition of $ a_{i,r}^{}(T,q)$). Thus, Lemma~\ref{lem:coeff_imm_poly} generalizes Theorem~\ref{thm:mukesh-siva-positive_immanant}. Using Lemma~\ref{lem:coeff_imm_poly}, the following result about monotonicity for the coefficients of all immanantal polynomials of $\sL_{T}^q$ along $\GTS_n$ was proved in \cite[Theorem~1]{mukesh-siva-gts}. \begin{theorem}[\sc Nagar and Sivasubramanian] \label{thm:main_earlier} Let $T_1$ and $T_2$ be two trees with $n$ vertices such that $T_2$ covers $T_1$ in $\GTS_n$. Let $\sL_{T_1}^q$ and $\sL_{T_2}^q$ be the $q$-Laplacians of $T_1$ and $T_2$ respectively. For $\lambda \vdash n$, let \begin{align*} \zeta^{\sL_{T_1}^q}_{s_{\lambda}}(x) & = d_{\lambda}(xI -\sL_{T_1}^q) = \sum_{r=0}^n (-1)^r c_{\lambda,r}^{\sL_{T_1}^q}(q) x^{n-r} \mbox{ and} \\ \zeta^{\sL_{T_2}^q}_{s_{\lambda}}(x) & = d_{\lambda}(xI -\sL_{T_2}^q) = \sum_{r=0}^n (-1)^r c_{\lambda,r}^{\sL_{T_2}^q}(q) x^{n-r}. \end{align*} Then for all $\lambda \vdash n$, we have $c_{\lambda,r}^{\sL_{T_1}^q}(q) - c_{\lambda,r}^{\sL_{T_2}^q}(q) \in \RR^+[q^2]$, where $0 \leq r \leq n$. \end{theorem} The proof of Theorem~\ref{thm:main_earlier} required to give a combinatorial expression for the coefficients $c_{\lambda,r}^{\sL_{T_1}^q}(q)$ and $c_{\lambda,r}^{\sL_{T_2}^q}(q)$ (see Lemma~\ref{lem:coeff_imm_poly}) and then to give an injection between the objects counting $c_{\lambda,r}^{\sL_{T_2}^q}(q)$ and those counting $c_{\lambda,r}^{\sL_{T_1}^q}(q)$ (see Lemma~\ref{lem:diff_a_k,r(T,q)}). \section{GMF of $q$-Laplacians arising from symmetric functions} \label{sec:gmfs_symm_fns} Let $\Gamma_{\gamma}=\ch^{-1}(\gamma)$ be the inverse Frobenius image of the symmetric function $\gamma \in \Lambda_{\QQ}^n$. Note that $\Gamma_{\gamma}$ is a class function of $\SSS_n$ over $\CC$ indexed by $\gamma$. Plugging the matrix $\sL_T^q=(l_{i,j}^q)_{1\leq i,j \leq n }$ in \eqref{eqn:gen_matrix_fn}, we get \begin{equation} \label{eqn:gen_matrix_fn_sL} d_{\gamma}(\sL_T^q) = \sum_{\psi \in \SSS_n} \Gamma_{\gamma}(\psi) \prod_{i=1}^n l_{i,\psi(i)}^q. \end{equation} %Recall the polynomial $a_i(T,q)$ that appears in Theorem %\ref{thm:mukesh-siva-positive_immanant} and that For the matrix $\sL_T^q$, it is very easy to show the following counterpart of Theorem~\ref{thm:mukesh-siva-positive_immanant}, where we change the Frobenius inverse of the Schur symmetric function to the Frobenius inverse of an arbitrary symmetric function. We recall the polynomial $a_i(T,q)$ from \eqref{eqn:q-analog-vertex-orient} %Lemma \ref{lem:coeff_imm_poly} and $\alpha_i(\gamma)$ from \eqref{eqn:main}. We start with the following result. As its proof is a verbatim copy of the proof of Theorem~\ref{thm:mukesh-siva-positive_immanant}, with the only change being that we replace $\alpha_{i,\lambda}$ by $\alpha_i(\gamma)$, we omit it. \begin{theorem} \label{thm:mukesh-siva-gen_matrix_fn} Let $\sL_T^q$ be the $q$-Laplacian matrix of a tree $T$ on $n$ vertices. Then, for all $\gamma \in \Lambda_{\QQ}^n$, $d_{\gamma}(\sL_T^q) = \sum_{i=0}^{\nhalf} \alpha_i (\gamma) a_i(T,q)$. \end{theorem} It is very simple to show the following counterpart of Lemma~\ref{lem:coeff_imm_poly}. Recall the polynomial $a_{i,r}(T,q)$ from \eqref{eqn:defn_a_riT}. Since the proof is identical, we omit it and merely state the result. \begin{lemma} \label{lem:coeff_gmf_poly} Let $T$ be a tree on $n$ vertices with $q$-Laplacian $\sL_T^q$. Then for all $\gamma \in \Lambda_{\QQ}^n$, we have $d_{\gamma}(xI-\sL_T^q)=\sum_{r=0}^n (-1)^r c_{\gamma,r}^{\sL_{T}^q}(q) x^{n-r}$, where $c_{\gamma,r}^{\sL_{T}^q}(q) =\sum_{i=0}^{\rhalf}\alpha_{i}(\gamma)a_{i,r}^{}(T,q)$ for $0 \leq r \leq n$. \end{lemma} Every element $\gamma \in \Lambda_{\QQ}^n$ is written as a linear combination of basis vectors of $\Lambda_{\QQ}^n$. The vector space $\Lambda_{\QQ}^n$ has six standard bases. We use standard terminology to denote each of the usual bases of $\Lambda_{\QQ}^n$. Thus, $s_{\lambda}$, $p_{\lambda}$, $e_{\lambda}$, $h_{\lambda}$, $m_{\lambda}$ and $f_{\lambda}$ denote the Schur, power sum, elementary, homogeneous, monomial and forgotten symmetric functions respectively. The inverse Frobenius map of each of these will be denoted by the same letter, but in capital font and with the partition $\lambda$ as a superscript rather than a subscript. Thus, $E^{\lambda} = \ch^{-1}(e_{\lambda})$, $\chi_{\lambda}^{} = S^{\lambda} = \ch^{-1}(s_{\lambda})$ and so on. As mentioned earlier, the inverse Frobenius image $\ch^{-1}(s_{\lambda})$ of the Schur symmetric function $s_{\lambda}$ is $\chi_{\lambda}^{}$, the irreducible character of $\SSS_n$ over $\CC$ indexed by $\lambda$. Therefore the identity given in \eqref{eqn:main_char} is a special case of the identity given in \eqref{eqn:main} when $\gamma=s_{\lambda}$. Thus, if any symmetric function $\gamma \in \Lambda_{\QQ}^n$ is Schur-positive (that is, $\gamma = \sum_{\lambda \vdash n} a_{\lambda} s_{\lambda}$ where $a_{\lambda} \in \RR^{+}$ for all $\lambda \vdash n$), then, by linearity, Lemma~\ref{lem:chan_lam_bino_char} will be true with $\chi_{\lambda}^{}$ replaced by $\ch^{-1}(\gamma)$ in \eqref{eqn:main_char}. Unfortunately, the monomial symmetric function $m_{\lambda}$ indexed by $\lambda$ is not Schur-positive. Thus, if $M^{\lambda} = \ch^{-1}(m_{\lambda})$, it is not clear that Lemma \ref{lem:chan_lam_bino_char} with $\chi_{\lambda}^{}$ replaced by $M^{\lambda}$ in \eqref{eqn:main_char} is true. For $\gamma \in \Lambda_{\QQ}^n$, let $\Gamma_{\gamma} = \ch^{-1}(\gamma)$. %In $\alpha_i^{}(\gamma)$, since we only need to evaluate $\Gamma_{\gamma}(\psi)=\Gamma_{\gamma}(j)$ for %a permutation $\psi \in \SSS_n$ with cycle type $2^j, 1^{n-2j}$. Thus, if $\Gamma_{\gamma}(j) \geq 0$ for all $j$, then from \eqref{eqn:main}, we get $\alpha_i(\gamma) \geq 0$. Since the inverse Frobenius image $\Gamma_{p_{\lambda}}$ of $p_{\lambda}$ is a scalar multiple of the indicator function of the conjugacy class $C_{\lambda}$ indexed by $\lambda$ (see \cite{EC2}), it follows that $\alpha_i^{}(p_{\lambda}) \geq 0$ for all $\lambda \vdash n$ and for all $i=0,1,\ldots,\nhalf$. As mentioned earlier, if $\gamma \in \Lambda_{\QQ}^n$ is Schur-positive, then $\alpha_i^{}(\gamma) \geq 0$. Since $h_{\lambda}$ is Schur-positive (see \cite[Corollary~7.12.4]{EC2}), it follows from Lemma~\ref{lem:chan_lam_bino_char} that $\alpha_i^{}(h_{\lambda}) \geq 0$ for all $\lambda \vdash n$. Similarly, it is well known that $e_{\lambda}$ is also Schur-positive. Thus $\alpha_i^{}(e_{\lambda}) \geq 0$ for all $\lambda \vdash n$. We record these facts below for future use. \begin{lemma} \label{lem:p_h_e_sum} For all $\lambda \vdash n$ and for $0 \leq i \leq \nhalf$, we have $\alpha_i^{}(\gamma) \geq 0$ for $\gamma \in \{p_{\lambda}, h_{\lambda}, e_{\lambda} \}.$ \end{lemma} Let the tree $T$ have $n$ vertices and, for $0\leq i,r\leq n$, let $a_{i,r}^{}(T,q)=a_{i,r}^T(q)$ be the polynomial in $q^2$ with non-negative real coefficients, defined by Nagar and Sivasubramanian in \cite[page~7]{mukesh-siva-gts}. They showed the following result (see \cite[Lemma~19, 23 and Corollary~22]{mukesh-siva-gts}). \begin{lemma}[\sc Nagar and Sivasubramanian] \label{lem:diff_a_k,r(T,q)} Let $T_1$ and $T_2$ be two trees with $n$ vertices such that $T_2$ covers $T_1$ in $\GTS_n$. Then for all $0\leq i,r\leq n$, we have $a_{i,r}^{}(T_1,q)-a_{i,r}^{}(T_2,q) \in \RR^+[q^2]$. \end{lemma} With Lemma~\ref{lem:p_h_e_sum} and Lemma~\ref{lem:diff_a_k,r(T,q)} in place, we can get the following Corollary. Its proof mimics the proof of Theorem~\ref{thm:main_earlier}. \begin{corollary} \label{cor:three} Theorem~\ref{thm:main_earlier} is true when we replace the immanantal polynomial $d_{\lambda}(xI - \sL_{T_j}^q)$ by the generalized matrix polynomials $d_{\gamma}(xI - \sL_{T_j}^q)$ for $\gamma \in \{p_{\lambda}, h_{\lambda}, e_{\lambda} \}$ and for $j=1,2$. \end{corollary} By Theorem~\ref{thm:main_earlier} and Corollary~\ref{cor:three}, we thus have monotonicity results about four of the six standard bases, as we move up on $\GTS_n$. This paper plugs the gaps left by considering the generalized matrix polynomials of $\sL_T^q$ arising from the last two bases $m_{\lambda}$ and $f_{\lambda}$. In other words, we consider the cases when we replace the immanantal polynomial $d_{\lambda}(xI - \sL_{T_j}^q)$ by $d_{g}(xI - \sL_{T_j}^q)$ when $g \in \{m_{\lambda}, f_{\lambda} \}$ in Theorem~\ref{thm:main_earlier}. \section{GMF of tree Laplacians arising from $m_{\lambda}$, the monomial symmetric function} \label{sec:gmfs_monomial} We prove Theorem~\ref{thm:main} in this section. We need the notion of $\lambda$-brick tabloids of shape $\mu$ which is used to give a combinatorial interpretation of the quantity $\Gamma_{m_{\lambda}}(\psi)=M^{\lambda}(\psi)$, where $\lambda, \mu\vdash n$ and $\psi\in\SSS_n.$ We recall the following definition of $\lambda$-brick tabloid of shape $\mu$ defined by E\u{g}ecio\u{g}lu and Remmel in \cite{remmel-monomial-sym-frob}. \begin{definition} Let $\lambda \vdash n$ with $\lambda =\lambda_1\geq \lambda_2\geq\cdots\geq \lambda_{l(\lambda)}^{}$. Let $\mu \vdash n$ and let $F_{\mu}$ be its Ferrers diagram. A $\lambda$-brick tabloid $\sB_{\lambda,\mu}$ of shape $\mu$ is a filling of $F_{\mu}$ with bricks $b_1,b_2,\ldots,b_{l(\lambda)}^{}$ of size $\lambda_1, \lambda_2,\ldots, \lambda_{l(\lambda)}^{}$ respectively such that brick $b_i$ covers exactly $\lambda_i$ squares of $F_{\mu}$ all of which lie in a single row of $F_{\mu}$ and no two bricks overlap. Here, bricks of the same size are indistinguishable. \end{definition} We refer the reader to the book by Mendes and Remmel \cite[Chapter~2]{mendes-remmel-book} for an introduction to $\lambda$-brick tabloids. Given a brick $b$ in $\sB_{\lambda,\mu}$, let $|b|$ denote the length of $b$. Define $\wt_{\sB_{\lambda,\mu}}^{}(b)$ to be $|b|$ if $b$ is at the end of a row in $\sB_{\lambda,\mu}$ and $1$ otherwise. We next define a weight $w(\sB_{\lambda,\mu})$ for each $\lambda$-brick tabloid $\sB_{\lambda,\mu}$ of shape $\mu$ as follows: \begin{equation*} \label{eqn:def_wt_bricks} w(\sB_{\lambda,\mu})=\prod_{b\in \sB_{\lambda,\mu}} \wt_{\sB_{\lambda,\mu}}^{}(b). \end{equation*} In other words, $w(\sB_{\lambda,\mu})$ is the product of the lengths of the rightmost brick in each row of $\sB_{\lambda,\mu}$. For $\lambda, \mu \vdash n$, let $\BT_{\lambda,\mu}$ be the set of all $\lambda$-brick tabloids of shape $\mu$. For $\lambda \vdash n$ and for $\psi \in \SSS_n$, E\u{g}ecio\u{g}lu and Remmel in \cite[Theorem~1]{remmel-monomial-sym-frob} gave the following combinatorial interpretation of $M^{\lambda}(\psi)$ which we need. \begin{theorem}[\sc E\u{g}ecio\u{g}lu and Remmel] \label{thm:main_thm_remmel_egec} For $\lambda \vdash n$, let $M^{\lambda}=\ch^{-1}(m_{\lambda})$. When a permutation $\psi \in \SSS_n$ has cycle type $\mu$, then $$M^{\lambda}(\psi)=(-1)^{l(\lambda)-l(\mu)}\sum_{\sB_{\lambda,\mu} \in \BT_{\lambda,\mu}}w(\sB_{\lambda,\mu}).$$ \end{theorem} We say $\lambda$ is a refinement of $\mu$ if the parts $\mu_i$ of $\mu$ can be obtained as a disjoint union of the parts $\lambda_j$ of $\lambda$. From the definition of $\sB_{\lambda,\mu}$, it is easy to see that if $\lambda$ is not a refinement of $\mu$, then $\BT_{\lambda,\mu}=\emptyset$. Thus $M^{\lambda}(\psi)=0$, unless when $\lambda$ refines the cycle type $\mu$ of $\psi$. Let $M^{\lambda}(j)$ denote the value of the function $M^{\lambda}(\cdot)$ evaluated at a permutation $\psi \in \SSS_n$ with cycle type $2^j,1^{n-2j}$. We start with the following simple consequence of Theorem~\ref{thm:main_thm_remmel_egec}, \begin{corollary} \label{cor:remmel_mono} Let $\lambda \vdash n$ and let $\psi \in \SSS_n$ be a permutation with cycle type $\mu=2^j,1^{n-2j}$. Then, $$M^{\lambda}(j)= \begin{cases} (-1)^{j-k}2^k\binom{j}{k}, & \mbox{if } \lambda=2^k,1^{n-2k} \mbox{ and } k\leq j, \\ 0, & \mbox{otherwise.} \end{cases} $$ \end{corollary} \begin{proof} We divide the proof into two cases when $\lambda$ is a refinement of $\mu$ and when it is not. If $\lambda$ is a refinement of $\mu$, then we must have $\lambda=2^k,1^{n-2k}$ for some $k$ with $0 \leq k\leq j.$ In this case, all the $k$ bricks of length $2$ must be placed in some of the $k$ rows from the first $j$ rows of $F_{\mu}$. Thus, the number of such $\lambda$-brick tabloids $\sB_{\lambda,\mu}$ of shape $\mu$ is $\binom{j}{k}$, and each $\sB_{\lambda,\mu}$ contributes the weight $2^k$ in the summation of Theorem~\ref{thm:main_thm_remmel_egec}. Thus, the total contribution in $M^{\lambda}(j)$ is $2^k\binom{j}{k}$ and $l(\lambda)-l(\mu)=j-k$. Hence $M^{\lambda}(j)= (-1)^{j-k}2^k\binom{j}{k}$. When $\lambda$ is not a refinement of $\mu$, by Theorem~\ref{thm:main_thm_remmel_egec}, $M^{\lambda}(j)=0$, completing the proof. \end{proof} %\vspace{2mm} We next prove Lemma~\ref{lem:frob_inv_monomial} which says that for all $\lambda \vdash n$ and for $0\leq i \leq \nhalf$, the quantity $\alpha_{i}^{}(m_{\lambda})$ is a non-negative integral multiple of $2^i$. We will need the following very easy identity involving binomial coefficients $\binom{j}{k}\binom{i}{j}=\binom{i-k}{j-k}\binom{i}{k}$. \vspace{2mm} \begin{proof}[Proof of Lemma 2] By Corollary~\ref{cor:remmel_mono}, when $\lambda\neq 2^k,1^{n-2k}$, $M^{\lambda}(j)= 0$ for all $j$. Thus, from \eqref{eqn:main}, $\alpha_{i}^{}(m_{\lambda})=0$ for $ 0 \leq i \leq \nhalf.$ When $\lambda= 2^k,1^{n-2k}\vdash n$ for some $k$ with $0\leq k \leq \nhalf$, by Corollary~\ref{cor:remmel_mono}, we see that \begin{align*} \alpha_{i}^{}(m_{\lambda}) &= \sum_{j=0}^{i}M^{\lambda}(j)\binom{i}{j} = \sum\limits_{j=0}^{i}(-1)^{j-k}2^k\binom{j}{k}\binom{i}{j} \\ &= 2^k \binom{i}{k}\sum\limits_{j=k}^{i}(-1)^{j-k}\binom{i-k}{j-k} = \begin{cases} 2^i, & \mbox{if } i=k, \\ 0, & \mbox{otherwise.} \end{cases} \end{align*} The proof is complete. \end{proof} \vspace{2mm} Using Lemmas~\ref{lem:frob_inv_monomial} and \ref{lem:diff_a_k,r(T,q)}, we can now prove Theorem~\ref{thm:main}. \vspace{2mm} \begin{proof}[Proof of Theorem 1] From Lemmas~\ref{lem:frob_inv_monomial} and \ref{lem:coeff_gmf_poly}, for $0 \leq r \leq n$ and for $j=1,2$, it is simple to see that the coefficient of $(-1)^rx^{n-r}$ in $d_{m_{\lambda}}(xI-\sL_{T_j}^q)$ is given by \begin{equation} \label{eqn:coeff_m} c_{m_{\lambda},r}^{\sL_{T_j}^q}(q) =\sum_{i=0}^{\rhalf}\alpha_{i}(m_{\lambda})a_{i,r}^{}(T_j,q) = \begin{cases} 2^k a_{k,r}^{}(T_j,q),& \mbox{if } \lambda=2^k,1^{n-2k} \mbox{ and } k\leq \rhalf, \\ 0, & \mbox{otherwise.} \end{cases} \end{equation} By Lemma~\ref{lem:diff_a_k,r(T,q)} and \eqref{eqn:coeff_m}, we get $c_{m_{\lambda},r}^{\sL_{T_1}^q}(q)-c_{m_{\lambda},r}^{\sL_{T_2}^q}(q)\in \RR^+[q^2]$ for all $\lambda\vdash n$ and for $0 \leq r \leq n.$ Furthermore, if $\lambda\neq 2^k,1^{n-2k} \vdash n$ for all $k$, then from \eqref{eqn:coeff_m}, the coefficient of $(-1)^rx^{n-r}$ in $d_{m_{\lambda}}(xI-\sL_{T_j}^q)$ is zero for $0 \leq r \leq n$ and $j=1,2$. The proof is complete. \end{proof} \vspace{2mm} Thus, for all $\lambda\vdash n$ and for all $q\in \RR$, moving up on $\GTS_n$ decreases the coefficient of $(-1)^rx^{n-r}$ in $d_{m_{\lambda}}(xI-\sL_{T}^q)$, where $0 \leq r \leq n$. Consequently, for all positive integers $n$, this monotonicity result on $\GTS_n$ shows that the max-min pair of these coefficients is $(P_n,S_n)$, where $P_n$ and $S_n$ are the path tree and the star tree on $n$ vertices respectively. The following example illustrates Lemma~\ref{lem:frob_inv_monomial}. \begin{example} Let $n=15$. For all $\lambda\vdash n$ and for all $i\in \{0,1,2,3,4,5,6,7\}$, the values of the quantity $\alpha_i(m_{\lambda})$ are tabulated in Table \ref{tab:alpha_r_m_lambda}. \begin{table}[h] $$ \begin{array}{|l||c|c|c|c|c|c|c|c|} \hline \lambda \vdash n=15 & i=0 & i=1 & i=2 & i=3 & i=4 & i=5 & i=6 & i=7 \\ \hline \hline \lambda= 1^{15} & 1 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\ \hline \lambda= 2,1^{13} & 0 & 2 & 0 & 0 & 0 & 0 & 0 & 0\\ \hline \lambda= 2^2,1^{11} & 0 & 0 & 4 & 0 & 0 & 0 & 0 & 0\\ \hline \lambda= 2^3,1^{9} & 0 & 0 & 0 & 8 & 0 & 0 & 0 & 0\\ \hline \lambda= 2^4,1^{7} & 0 & 0 & 0 & 0 & 16 & 0 & 0 & 0\\ \hline \lambda= 2^5,1^{5} & 0 & 0 & 0 & 0 & 0 & 32 & 0 & 0\\ \hline \lambda= 2^6,1^{3} & 0 & 0 & 0 & 0 & 0 & 0 & 64 & 0\\ \hline \lambda= 2^7,1 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 128\\ \hline \lambda\neq 2^k,1^{n-2k} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0\\ \hline \end{array} $$ \caption{The value of $\alpha_i(m_{\lambda})$.} \label{tab:alpha_r_m_lambda} \end{table} \end{example} \section{GMF of tree Laplacians arising from $f_{\lambda}$, the forgotten symmetric function} \label{sec:gmfs_forgotten_symm_fn} Let $f_{\lambda}$ denote the forgotten symmetric function indexed by $\lambda \vdash n$ and let $F^{\lambda}=\ch^{-1}(f_{\lambda})$ denote its inverse Frobenius image. This section is devoted to show monotonicity of the coefficient of $(-1)^rx^{n-r}$ in $\zeta_{f_{\lambda}}^{\sL_T^q}(x)=d_{f_{\lambda}}(xI-\sL_T^q)$ when we move up along $\GTS_n$. For this, we need the following combinatorial interpretation of $F^{\lambda}(\psi)$ by E\u{g}ecio\u{g}lu and Remmel in \cite[Theorem~8]{remmel-monomial-sym-frob}. \begin{theorem}[\sc E\u{g}ecio\u{g}lu and Remmel] \label{thm:remmel_forgottan_inte} Let $\lambda \vdash n$ and let $\psi \in \SSS_n$ be a permutation with cycle type $\mu$. Then, $$F^{\lambda}(\psi)=(-1)^{n-l(\mu)}\sum_{\sB_{\lambda,\mu} \in \BT_{\lambda,\mu}}w(\sB_{\lambda,\mu}).$$ \end{theorem} For $\lambda \vdash n$, let $F^{\lambda}(j)$ denote the value of the function $F^{\lambda}(\cdot)$ evaluated at a permutation $\psi \in \SSS_n$ with cycle type $2^j,1^{n-2j}$. By Theorem~\ref{thm:remmel_forgottan_inte}, the proof of the following corollary is identical to the proof of Corollary~\ref{cor:remmel_mono}, we omit it and merely state the result. \begin{corollary} \label{cor:remmel_forgottan} For $\lambda \vdash n$ and for $0\leq j \leq \nhalf$, we have $$F^{\lambda}(j)= \begin{cases} (-1)^{j}2^k\binom{j}{k}, & \mbox{if } \lambda=2^k,1^{n-2k} \mbox{ and } k\leq j, \\ 0, & \mbox{otherwise.} \end{cases} $$ \end{corollary} For a fixed $\lambda \vdash n$, let $\gamma=f_{\lambda}$ in \eqref{eqn:main}. We next calculate the quantity $\alpha_{i}(f_{\lambda})$ in the following lemma which will be used later in the determination of the coefficient of $(-1)^rx^{n-r}$ in $d_{f_{\lambda}}(xI-\sL_{T}^q).$ \begin{lemma} \label{lem:alpha_f} Let $\lambda \vdash n$. Then for $0\leq i \leq \nhalf$, we have $$\alpha_{i}^{}(f_{\lambda})= \begin{cases} (-1)^i 2^i, & \mbox{if } \lambda=2^i,1^{n-2i}, \\ 0, & \mbox{otherwise.} \end{cases} $$ \end{lemma} \begin{proof} By Corollary~\ref{cor:remmel_forgottan} when $\lambda\neq 2^k,1^{n-2k}$, $F^{\lambda}(j)=0$ for all $j$. By \eqref{eqn:main}, $\alpha_{i}^{}(f_{\lambda})=0$ for $0 \leq i \leq \nhalf.$ We next assume $\lambda= 2^k,1^{n-2k}\vdash n$ for some $k$ with $0\leq k \leq \nhalf$. In this case, by Corollary~\ref{cor:remmel_forgottan}, we see that \begin{align*} \alpha_{i}^{}(f_{\lambda}) & = \sum_{j=0}^{i}F^{\lambda}(j)\binom{i}{j} = \sum\limits_{j=0}^{i}(-1)^{j}2^k\binom{j}{k}\binom{i}{j} \\ & = 2^k \binom{i}{k}\sum\limits_{j=k}^{i}(-1)^{j}\binom{i-k}{j-k} % = % (-1)^k 2^k \binom{i}{k}\sum\limits_{i=0}^{i}(-1)^{i}\binom{i-k}{i} \\& = \begin{cases} (-1)^i 2^i, & \mbox{if } i=k, \\ 0, & \mbox{otherwise.} \end{cases} \end{align*} All equalities above follow by simple manipulations and hence the proof is complete. \end{proof} \begin{theorem} \label{thm:main_f_lambda} Let $T$ be a tree on $n$ vertices with $q$-Laplacian matrix $\sL_{T}^q$. Then for $0\leq r \leq n$, the coefficient of $(-1)^rx^{n-r}$ in $d_{f_{\lambda}}(xI-\sL_{T}^q)$ is given by \begin{equation*} \label{eqn:coeff_f} c_{f_{\lambda},r}^{\sL_{T}^q}(q) = \begin{cases} (-1)^k2^k a_{k,r}(T,q), & \mbox{if } \lambda=2^k,1^{n-2k} \mbox{ with } k \leq \rhalf, \\ 0, & \mbox{otherwise.} \end{cases} \end{equation*} Furthermore, for all $\lambda\vdash n$ and all $q\in \RR$, moving up on $\GTS_n$ decreases $c_{f_{\lambda},r}^{\sL_{T}^q}(q)$ in absolute value. \end{theorem} \begin{proof} From Lemmas~\ref{lem:coeff_gmf_poly} and \ref{lem:alpha_f}, it is simple to see that \begin{equation*} c_{f_{\lambda},r}^{\sL_{T}^q}(q) =\sum_{i=0}^{\rhalf}\alpha_{i}(f_{\lambda})a_{i,r}^{}(T,q) = \begin{cases} (-1)^k2^k a_{k,r}^{}(T,q),& \mbox{if } \lambda=2^k,1^{n-2k} \mbox{ and } k\leq \rhalf, \\ 0, & \mbox{otherwise.} \end{cases} \end{equation*} Let $T_1$ and $T_2$ be two trees with $n$ vertices such that $T_2$ covers $T_1$ in $\GTS_n$. By Lemma~\ref{lem:diff_a_k,r(T,q)}, $\left|c_{f_{\lambda},r}^{\sL_{T_1}^q}(q)\right|-\left|c_{f_{\lambda},r}^{\sL_{T_2}^q}(q)\right|\in \RR^+[q^2]$ completing the proof. \end{proof} \vspace{2mm} Thus, for all $\lambda\vdash n$ when $ 0 \leq r \leq n$, by Theorem~\ref{thm:main}, Theorem~\ref{thm:main_f_lambda} and Corollary~\ref{cor:three}, we get that the coefficient of $x^r$ in $d_{\gamma}(xI-\sL_{T}^q)$ decreases as we move up along $\GTS_n$ in absolute value for each $\gamma\in \{m_{\lambda},s_{\lambda},p_{\lambda},h_{\lambda},e_{\lambda},f_{\lambda}\}$. Plugging in $q=1$, we get the following corollary of Theorem~\ref{thm:main}, Theorem~\ref{thm:main_f_lambda} and Corollary~\ref{cor:three}. \begin{corollary} For all $\lambda\vdash n$ and for all $\gamma\in \{m_{\lambda},s_{\lambda},p_{\lambda},h_{\lambda},e_{\lambda},f_{\lambda}\}$, moving up on $\GTS_n$ decreases the coefficient of $x^r$ in $d_{\gamma}(xI-L_T)$ in absolute value, for $0 \leq r \leq n$. Here $L_T$ is the usual combinatorial Laplacian of the tree $T$. \end{corollary} \section*{Acknowledgement} The first author would like to acknowledge SERB, Government of India for providing a national postdoctoral fellowship with the file number PDF/2018/000828. The second author acknowledges support from project MTR/2018/000453 given by SERB, Government of India. Our main theorem in this work was in its conjecture form, tested using the computer package ``SageMath''. We thank the authors for generously releasing SageMath as an open-source package. \begin{thebibliography}{10} \bibitem{bapat-lal-pati} {\sc Bapat, R.~B., Lal, A.~K., and Pati, S.} \newblock A $q$-analogue of the distance matrix of a tree. \newblock {\em Linear Algebra and its Applications 416\/} (2006), 799--814. (p.~3) \bibitem{bass} {\sc Bass, H.} \newblock The {I}hara-{S}elberg zeta function of a tree lattice. \newblock {\em International Journal of Mathematics 3\/} (1992), 717--797. (p.~3) \bibitem{chan-lam-binom-coeffs-char} {\sc Chan, O., and Lam, T.~K.} \newblock Binomial {C}oefficients and characters of the symmetric group. \newblock {\em Technical Report 693\/} (1996), National University of Singapore. (p.~3) \bibitem{hook_immanant_explained-chan_lam} {\sc Chan, O., and Lam, T.~K.} \newblock Hook immanantal inequalities for trees explained. \newblock {\em Linear Algebra and its Applications 273\/} (1998), 119--131. (p.~3) \bibitem{csikvari-poset1} {\sc Csikv{\'a}ri, P.} \newblock On a poset of trees. \newblock {\em Combinatorica 30 (2)\/} (2010), 125--137. (p.~2) \bibitem{remmel-monomial-sym-frob} {\sc E\u{g}ecio\u{g}lu, {\"O}., and Remmel, J.~B.} \newblock The monomial symmetric functions and the {F}robenius map. \newblock {\em Journal of Combinatorial Theory, Series A 54}, 2 (1990), 272--295. (pp.~6,~8) \bibitem{foata-zeilberger-bass-trams} {\sc Foata, D., and Zeilberger, D.} \newblock Combinatorial proofs of {B}ass's evaluations of the {I}hara--{S}elberg zeta function of a graph. \newblock {\em Transactions of the American Mathematical Society 351\/} (1999), 2257--2274. (p.~3) \bibitem{godsil-royle} {\sc Godsil, C., and Royle, G.} \newblock {\em Algebraic Graph Theory}. \newblock Springer Verlag, 2001. (p.~3) \bibitem{mendes-remmel-book} {\sc Mendes, A., and Remmel, J.~B.} \newblock {\em {C}ounting with {S}ymmetric {F}unctions}. \newblock Developments in Mathematics. Springer, Cham, 2015. (pp.~1,~6) \bibitem{nagar} {\sc Nagar, M.~K.} \newblock Eigenvalue monotonicity of $q$-{L}aplacians of trees along a poset. \newblock {\em Linear Algebra and its Applications 571\/} (2019), 110--131. (p.~3) \bibitem{mukesh-siva-hook} {\sc Nagar, M.~K., and Sivasubramanian, S.} \newblock Hook immanantal and {H}adamard inequalities for $q$-{L}aplacians of trees. \newblock {\em Linear Algebra and its Applications 523\/} (2017), 131--151. (pp.~3,~4) \bibitem{mukesh-siva-gts} {\sc Nagar, M.~K., and Sivasubramanian, S.} \newblock Laplacian immanantal polynomials and the {G}{T}{S} poset on trees. \newblock {\em Linear Algebra and its Applications 561\/} (2019), 1--23. (pp.~2, 4,~6) \bibitem{schur-immanant-ineqs} {\sc Schur, I.} \newblock {\"U}ber endliche {G}ruppen und {H}ermitesche {F}ormen. \newblock {\em Mathematische Zeitschrift 1\/} (1918), 184--207. (p.~3) \bibitem{EC2} {\sc Stanley, R.~P.} \newblock {\em Enumerative Combinatorics, vol.~2}. \newblock Cambridge University Press, 2001. (pp.~1, 2,~5) \end{thebibliography} \end{document} .