\documentclass[reqno]{amsart} \usepackage{hyperref} \AtBeginDocument{{\noindent\small \emph{Electronic Journal of Differential Equations}, Vol. 2015 (2015), No. 199, pp. 1--11.\newline ISSN: 1072-6691. URL: http://ejde.math.txstate.edu or http://ejde.math.unt.edu \newline ftp ejde.math.txstate.edu} \thanks{\copyright 2015 Texas State University - San Marcos.} \vspace{9mm}} \begin{document} \title[\hfilneg EJDE-2015/199\hfil Schwarzian difference equations] {Properties of Schwarzian difference equations} \author[S.-T. Lan, Z.-X. Chen \hfil EJDE-2015/199\hfilneg] {Shuang-Ting Lan, Zong-Xuan Chen} \address{Shuang-Ting Lan \newline Department of Mathematics, Guangzhou Civil Aviation College, Guangzhou 510403, China} \email{wqh200811@126.com} \address{Zong-Xuan Chen (corresponding author) \newline School of Mathematical Sciences, South China Normal University, Guangzhou 510631, China} \email{chzx@vip.sina.com} \thanks{Submitted April 22, 2015. Published August 4, 2015.} \subjclass[2010]{30D35, 34A20} \keywords{Schwarzian difference equation; rational solution; value distribution} \begin{abstract} We consider the Schwarzian type difference equation $$ \Big[\frac{\Delta^{3}f(z)}{\Delta f(z)}-\frac{3}{2} \Big(\frac{\Delta^{2}f(z)}{\Delta f(z)}\Big)^{2}\Big]^{k}=R(z), $$ where $R(z)$ is a nonconstant rational function. We study the existence of rational solutions and value distribution of transcendental meromorphic solutions with finite order of the above equation. \end{abstract} \maketitle \numberwithin{equation}{section} \newtheorem{theorem}{Theorem}[section] \newtheorem{lemma}[theorem]{Lemma} \newtheorem{remark}[theorem]{Remark} \newtheorem{corollary}[theorem]{Corollary} \newtheorem{example}[theorem]{Example} \allowdisplaybreaks \section{Introduction and statement of main results} In this article, we use the basic notions of Nevanlinna's theory \cite{ha2, ya}. In addition, $\sigma(f)$ denotes the order of growth of the meromorphic function $f(z)$; $\lambda(f)$ and $\lambda\left(\frac{1}{f}\right)$ denote the exponents of convergence of zeros and poles of $f(z)$. Let $S(r,w)$ denote any quantity satisfying $S(r,w) = o\big(T(r,w)\big)$ for all $r$ outside of a set with finite logarithmic measure. A meromorphic solution $w$ of a difference (or differential) equation is called \textit{admissible} if the characteristic function of all coefficients of the equation are $S(r,w)$. For every $n\in\mathbb{N}^{+}$, the forward differences $\Delta^{n}f(z)$ are defined in the standard way \cite{wh} by \[ \Delta f(z)=f(z+1)-f(z),\quad \Delta^{n+1} f(z)=\Delta^{n}f(z+1)-\Delta^{n}f(z). \] The Schwarzian differential equation \begin{equation}\label{eq0} \big[\frac{f'''}{f'}-\frac{3}{2}\big(\frac{f''}{f'}\big)^{2}\big]^{k} =R(z,f)=\frac{P(z,f)}{Q(z,f)} \end{equation} was studied by Ishizaki \cite{is}, and obtained some important results. Chen and Li \cite{ch2} investigated Schwarzian difference equation, and obtained the following theorem. \begin{theorem} \label{thmA} Let $f(z)$ be an admissible solution of difference equation \[%\label{eq4} \Big[\frac{\Delta^{3}f(z)}{\Delta f(z)}-\frac{3}{2} \Big(\frac{\Delta^{2}f(z)}{\Delta f(z)}\Big)^{2}\Big]^{k} =R(z,f) =\frac{P(z,f)}{Q(z,f)} \] such that $\sigma_{2}(f)<1$, where $k(\geq1)$ is an integer, $P(z,f)$ and $Q(z,f)$ are polynomials with $\deg_{f}P(z,f)=p,~ \deg_{f}Q(z,f)=q$, $d=\max\{p,q\}$. Let $\alpha_{1}, \ldots, \alpha_{s}$ be $s(\geq2)$ distinct complex constants. Then $$ \sum_{j=1}^{s}\delta(\alpha_{j}, f)\leq4-\frac{q}{2k}. $$ In particular, if $N(r,f)=S(r,f)$, then $$ \sum_{j=1}^{s}\delta(\alpha_{j}, f)\leq2-\frac{d}{2k}. $$ \end{theorem} Set $\deg_{f}P(z,f)=\deg_{f}Q(z,f)=0$ in equation \eqref{eq0}, then $R(z,f)\equiv R(z)$ is a small function with respect to $f(z)$. Liao and Ye \cite{li} studied this type of Schwarzian differential equation, and obtained the following result. \begin{theorem} \label{thmB} Let $P$ and $Q$ be polynomials with $\deg P=p, ~\deg Q=q$, and let $R(z)=\frac{P(z)}{Q(z)}$ and $k$ a positive integer. If $f(z)$ is a transcendental meromorphic solution of equation $$ \Big[\frac{f'''}{f'}-\frac{3}{2}\Big(\frac{f''}{f'}\Big)^{2}\Big]^{k}=R(z), $$ then $p-q+2k>0$ and the order $\sigma(f)=\frac{p-q+2k}{2k}$. \end{theorem} In this article, we study a Schwarzian difference equation, and obtain the following result. \begin{theorem}\label{thm1} Let $R(z)=\frac{P(z)}{Q(z)}$ be an irreducible rational function with $\deg P(z)=p$, $\deg Q(z)=q$. Consider the difference equation \begin{equation}\label{eq1} \Big[\frac{\Delta^{3}f(z)}{\Delta f(z)} -\frac{3}{2}\Big(\frac{\Delta^{2}f(z)}{\Delta f(z)}\Big)^{2}\Big]^{k}=R(z), \end{equation} where $k$ is a positive integer. Then \begin{itemize} \item[(i)] every transcendental meromorphic solution $f(z)$ of \eqref{eq1} satisfies $\sigma(f)\geq 1$; if $p-q+2k>0$, then \eqref{eq1} has no rational solutions; \item[(ii)] if $f(z)$ is a mereomorphic solution of \eqref{eq1} with finite order, terms $\frac{\Delta^{2}f(z)}{\Delta f(z)}$ and $\frac{\Delta^{3}f(z)}{\Delta f(z)}$ in \eqref{eq1} are nonconstant rational functions; \item[(iii)] every transcendental meromorphic solution $f(z)$ with finite order has at most one Borel exceptional value unless \begin{equation}\label{eq00} f(z)=b+R_0(z)e^{az}, \end{equation} where $b\in\mathbb{C}, ~a\in\mathbb{C}\setminus\{0\}$ and $R_0(z)$ is a nonzero rational function. \item[(iv)] if $p-q+2k>0, \sigma(f)<\infty$, then $\Delta f(z)$ has at most one Borel exceptional value unless \begin{equation}\label{eq01} \Delta f(z)=R_{1}(z)e^{az}, \end{equation} where $a\in\mathbb{C},~ a\neq i2k_{1}\pi$ for any $k_{1}\in\mathbb{Z}$, and $R_{1}(z)$ is a nonzero rational function. \end{itemize} \end{theorem} \begin{corollary}\label{cor1} Let $f(z)$ be a finite order meromorphic solution of \eqref{eq1}, if $p-q+2k>0$, then $f(z)$, $\Delta f(z)$, $\Delta^{2}f(z)$ and $\Delta^{3}f(z)$ cannot be rational functions, and $\frac{\Delta^{2}f(z)}{\Delta f(z)}$ and $\frac{\Delta^{3}f(z)}{\Delta f(z)}$ are nonconstant rational functions. \end{corollary} \begin{remark}\label{re0} \rm Let $f(z)$ be the function in the form \eqref{eq00}, then the Schwarzian difference is an irreducible rational function $R(z)=\frac{P(z)}{Q(z)}$ with $\deg P\leq\deg Q$. \end{remark} \begin{proof} Suppose that $f(z)$ has the form \eqref{eq00}. Since $R_0(z)$ is a rational function, we see $R_0(z)$ satisfies \begin{equation}\label{043} \frac{R_0(z+j)}{R_0(z)}\to 1,\quad z\to \infty,\; j=1,2,3. \end{equation} By \eqref{eq00}, we have \begin{gather*} \Delta f(z)=e^{az}(e^{a}R_0(z+1)-R_0(z));\\ \Delta^{2}f(z)=e^{az}(e^{2a}R_0(z+2)-2e^{a}R_0(z+1)+R_0(z));\\ \Delta^{3}f(z)=e^{az}(e^{3a}R_0(z+3)-3e^{2a}R_0(z+2)+3e^{a}R_0(z+1)-R_0(z)). \end{gather*} Combining these with \eqref{043}, we have \begin{equation} \label{001} \begin{aligned} \frac{\Delta^{3}f(z)}{\Delta f(z)} &=\frac{e^{3a}R_0(z+3)-3e^{2a}R_0(z+2)+3e^{a}R_0(z+1)-R_0(z)} {e^{a}R_0(z+1)-R_0(z)}\\ &=\frac{e^{3a}\frac{R_0(z+3)}{R_0(z)}-3e^{2a}\frac{R_0(z+2)}{R_0(z)} +3e^{a}\frac{R_0(z+1)}{R_0(z)}-1}{e^{a}\frac{R_0(z+1)}{R_0(z)}-1} \\ &\to \frac{e^{3a}-3e^{2a}+3e^{a}-1}{e^{a}-1}=(e^{a}-1)^{2},\quad z\to \infty, \end{aligned} \end{equation} and \begin{equation} \label{003} \begin{aligned} \frac{\Delta^{2}f(z)}{\Delta f(z)} &=\frac{e^{2a}R_0(z+2)-2e^{a}R_0(z+1)+R_0(z)}{e^{a}R_0(z+1)-R_0(z)}\\ &=\frac{e^{2a}\frac{R_0(z+2)}{R_0(z)}-2e^{a}\frac{R_0(z+1)}{R_0(z)}+1} {e^{a}\frac{R_0(z+1)}{R_0(z)}-1} \\ &\to \frac{e^{2a}-2e^{a}+1}{e^{a}-1}=e^{a}-1,\quad z\to \infty. \end{aligned} \end{equation} Thus, \begin{equation}\label{004} \frac{\Delta^{3}f(z)}{\Delta f(z)} -\frac{3}{2}\Big(\frac{\Delta^{2}f(z)}{\Delta f(z)}\Big)^{2} \to (e^{a}-1)^{2}-\frac{3}{2}(e^{a}-1)^{2} =-\frac{1}{2}(e^{a}-1)^{2},\quad z\to \infty. \end{equation} By \eqref{001}, \eqref{003} and $R_0(z)$ begin a rational function, we see that $$ R(z)=\Big[\frac{\Delta^{3}f(z)}{\Delta f(z)}-\frac{3}{2} \Big(\frac{\Delta^{2}f(z)}{\Delta f(z)}\Big)^{2}\Big]^{k} $$ is a rational function. Denote $R(z)=\frac{P(z)}{Q(z)}$, where $P(z)$ and $Q(z)$ are prime polynomials. By \eqref{004}, we see $$ R(z)=\frac{P(z)}{Q(z)} =\Big[\frac{\Delta^{3}f(z)}{\Delta f(z)}-\frac{3}{2} \Big(\frac{\Delta^{2}f(z)}{\Delta f(z)}\Big)^{2}\Big]^{k}\to \frac{(-1)^{k}}{2^{k}}(e^{a}-1)^{2k}, \quad z\to \infty. $$ If $e^{a}\neq1$, then $\deg P=\deg Q$; if $e^{a}=1$, then $\deg P<\deg Q$. So, $\deg P\leq \deg Q$. \end{proof} \begin{remark}\label{re1}\rm Checking the proof of Theorem \ref{thm1} (iv), we see that for $f(z)$ a function such that $\Delta f(z)$ in the form \eqref{eq01}, then the Schwarzian difference satisfies $$ \frac{\Delta^{3}f(z)}{\Delta f(z)}-\frac{3}{2} \Big(\frac{\Delta^{2}f(z)}{\Delta f(z)}\Big)^{2} =e^{2a}\frac{R_{1}(z+2)}{R_{1}(z)}-\frac{3}{2}e^{2a} \Big(\frac{R_{1}(z+1)}{R_{1}(z)}\Big)^{2} +e^{a}\frac{R_{1}(z+1)}{R_{1}(z)}-\frac{1}{2}. $$ \end{remark} Examples \ref{ex1} and \ref{ex2} below show that the condition ``$p-q+2k>0$'' in Theorem \ref{thm1} (i) cannot be omitted. \begin{example}\label{ex1} \rm Consider the Schwarzian type difference equation $$ \frac{\Delta^{3}f(z)}{\Delta f(z)} -\frac{3}{2}\Big(\frac{\Delta^{2}f(z)}{\Delta f(z)}\Big)^{2} =-\frac{6}{(2z+1)^{2}}, $$ where $k=1$, $p=0$, $q=2$, and $p-q+2k=0$. This equation has a rational solution $f_{1}(z)=z^{2}$, and a transcendental meromorphic solution $f_{2}(z)=e^{i2\pi z}+z^{2}$. \end{example} \begin{example}\label{ex2} \rm Consider the Schwarzian type difference equation $$ \frac{\Delta^{3}f(z)}{\Delta f(z)} -\frac{3}{2}\Big(\frac{\Delta^{2}f(z)}{\Delta f(z)}\Big)^{2} =\frac{-6}{(z+3)(z+2)^{2}}, $$ where $k=1$, $p=0$, $q=3$, and $p-q+2k=-1<0$. This equation has a rational solution $f_{1}(z)=\frac{1}{z}$, and a transcendental meromorphic solution $f_{2}(z)=e^{i2\pi z}+\frac{1}{z}$. \end{example} \begin{example}\label{ex3} \rm The function $f(z)=ze^{(\log3)z}$ satisfies Schwarzian type difference equation $$ \frac{\Delta^{3}f(z)}{\Delta f(z)} -\frac{3}{2}\Big(\frac{\Delta^{2}f(z)}{\Delta f(z)}\Big)^{2} =\frac{-8z^{2}-48z-108}{(2z+3)^{2}}. $$ We see $\sigma(f)=1$ and $f(z)$ has finitely many zeros and poles. It shows the result of Theorem \ref{thm1} (iii) is precise. \end{example} \section{Preliminaries} \begin{lemma}[\cite{ch1}] \label{le11} Let $f(z)$ be a meromorphic function of finite order $\sigma$ and let $\eta$ be a nonzero complex constant. Then for each $\varepsilon(0<\varepsilon<1)$, we have \[ m\Big(r, \frac{f(z+\eta)}{f(z)}\Big)+m\Big(r, \frac{f(z)}{f(z+\eta)}\Big) =O(r^{\sigma-1+\varepsilon}). \] \end{lemma} \begin{lemma}[\cite{ch1}] \label{le1} Let $f(z)$ be a meromorphic function with order $\sigma=\sigma(f), \sigma<\infty$, and let $\eta$ be a fixed nonzero complex number, then for each $\varepsilon>0$, \[ T(r, f(z+\eta))=T(r, f(z))+O\big(r^{\sigma-1+\varepsilon}\big)+O(\log r). \] \end{lemma} \begin{lemma}[{\cite[Theorem 1.8.1]{ch3}, \cite{la2}}] \label{lecha} Let $c\in\mathbb{C}\setminus\{0\}$ and $f(z)$ be a finite order meromorphic function with two finite Borel exceptional values $a$ and $b$. Then for every $n\in\mathbb{N}^{+}$, $$ T(r,\Delta^{n}f)=(n+1)T(r,f)+S(r,f) $$ unless $f(z)$ and $c$ satisfy \begin{gather*} f(z)=b+\frac{b-a}{pe^{dz}-1}, \quad p, d\in\mathbb{C}\setminus\{0\},\\ mdc=i2k_{1}\pi\quad \text{for some $k_{1}\in\mathbb{Z}$ and } m\in\{1, 2, \ldots, n\}. \end{gather*} \end{lemma} \begin{remark}\label{re2} \rm Checking the proof of Lemma \ref{lecha}, we point out that when $c\in\mathbb{C}\setminus\{0\}$ and $f(z)$ is a finite order meromorphic function with two finite Borel exceptional values, for every $n\in\mathbb{N}^{+}$, if $c,2c,\ldots, nc$ are not periods of $f(z)$, then $$ T(r,\Delta^{n}f)=(n+1)T(r,f)+S(r,f). $$ \end{remark} \begin{lemma}[\cite{be}] \label{le7} Let $f(z)$ be a function transcendental and meromorphic in the plane which satisfies $$ \liminf_{r\to \infty}\frac{T(r,f)}{r}=0. $$ Then $\Delta f$ and $\Delta f/f$ are both transcendental. \end{lemma} \begin{lemma}\label{le4} Suppose that $f(z)=H(z)e^{az}$, where $a\neq0$ is a constant, $H(z)$ is a transcendental meromorphic function with $\sigma(H)<1$. Then $\frac{\Delta f(z)}{f(z)}$ is transcendental. \end{lemma} \begin{proof} Substituting $f(z)=H(z)e^{az}$ into $\frac{\Delta f(z)}{f(z)}$, we see that \begin{equation} \label{030} \begin{aligned} \frac{\Delta f(z)}{f(z)} &=\frac{f(z+1)-f(z)}{f(z)} =\frac{H(z+1)e^{a(z+1)}-H(z)e^{az}}{H(z)e^{az}} \\ &=e^{a}\frac{H(z+1)}{H(z)}-1 =e^{a}\Big(\frac{H(z+1)}{H(z)}-1\Big)+e^{a}-1\\ &=e^{a}\frac{\Delta H(z)}{H(z)}+e^{a}-1. \end{aligned} \end{equation} From the fact $\sigma(H)<1$, we see that $$ \limsup_{r\to \infty}\frac{\log T(r,H)}{\log r}=\sigma(H)<1. $$ Then for large enough $r$, choose $\varepsilon=\frac{1-\sigma(H)}{2}>0$, we have $$ \log T(r,H)<(\sigma(H)+\varepsilon)\log r; $$ that is, $$ T(r,H)0$. By \eqref{010} and \eqref{011}, when $z$ is large enough, $g(z)$ can be written as \begin{equation}\label{012} g(z)=c_0z^{l}(1+o(1)). \end{equation} Hence, \begin{equation}\label{013} \Delta g(z)=lc_0z^{l-1}(1+o(1)), \quad \Delta^{2}g(z)=l(l-1)c_0z^{l-2}(1+o(1)). \end{equation} Substituting \eqref{014z}, \eqref{012}, \eqref{013} in \eqref{014}, we obtain $$ c_0z^{l}l(l-1)c_0z^{l-2}(1+o(1))-\frac{3}{2}(lc_0z^{l-1})^{2}(1+o(1)) =Bz^{\frac{p-q}{k}}c_0^{2}z^{2l}(1+o(1)); $$ that is, $$ -\big(\frac{l}{2}+1\big)lc_0^{2}z^{2l-2}(1+o(1)) =Bz^{\frac{p-q}{k}}c_0^{2}z^{2l}(1+o(1)), $$ from which it follows $$ 2l-2=\frac{p-q}{k}+2l. $$ So, $p-q+2k=0$. \smallskip \noindent\textbf{Case 2.} $l=0$, $c_0\neq0$. By \eqref{010} and \eqref{011}, when $z$ is large enough, $g(z)$ can be written as \begin{equation}\label{015} g(z)=c_0+\frac{m(z)}{n(z)}=c_0+o(1). \end{equation} By calculation and $m\frac{p-q}{k}+(2m-2n), $$ thus, $p-q+2k<0$. By the above Cases 1--3, we see if \eqref{eq1} has a rational solution $f(z)$, then $p-q+2k\leq0$. (ii) By Lemma \ref{le6}, we see that Theorem \ref{thm1} (ii) holds. (iii) Set $G(z)=\frac{\Delta^{2}f(z)}{\Delta f(z)}$. Lemma \ref{le6} shows $G(z)$ is a nonconstant rational function. Then \begin{equation}\label{019} \Delta^{2}f(z)=G(z)\Delta f(z), \end{equation} By \eqref{eq1}, we easily see $\Delta f(z)\not\equiv0$, that is $f(z+1)\not\equiv f(z)$. Assert that $f(z+2)\not\equiv f(z)$. Otherwise, $$ \Delta^{2}f(z)=f(z+2)-2f(z+1)+f(z)=2f(z)-2f(z+1)=-2\Delta f(z). $$ Together with \eqref{019}, $$ G(z)=\frac{\Delta^{2}f(z)}{\Delta f(z)}\equiv-2, $$ which contradicts with the fact $G(z)$ is a nonconstant rational function. If $f(z)$ has two finite Borel exceptional values, by $f(z+2)\not\equiv f(z)$, $f(z+1)\not\equiv f(z)$ and Remark \ref{re2}, we have $$ T(r,\Delta^{2}f)=3T(r, f)+S(r,f),\quad T(r,\Delta f)=2T(r, f)+S(r,f). $$ On the other hand, \eqref{019} shows that $$ T(r,\Delta^{2}f)=T(r,\Delta f)+O(\log r). $$ The last two equalities follows $T(r,f)=S(r,f).$ It is a contradiction. So, $f(z)$ cannot have two finite Borel exceptional values. Suppose that $f(z)$ has two Borel exceptional values $b\in\mathbb{C}$ and $\infty$. By Hadamard's factorization theory, $f(z)$ takes the form \begin{equation}\label{040} f(z)=b+R_0(z)e^{h(z)}, \end{equation} where $R_0(z)$ is a meromorphic function, and $h(z)$ is a polynomial such that $$ \sigma(R_0)=\max\big\{\lambda(f-b), \lambda\big(\frac{1}{f}\big)\big\}<\deg h. $$ Thus, \begin{equation}\label{041} \Delta f(z)=\left(R_0(z+1)e^{h(z+1)-h(z)}-R_0(z)\right)e^{h(z)}=R_{1}(z)e^{h(z)}, \end{equation} where $R_{1}(z)=R_0(z+1)e^{h(z+1)-h(z)}-R_0(z)$. Obviously, \begin{equation}\label{042} \sigma(R_{1})=\sigma\Big(R_0(z+1)e^{h(z+1)-h(z)}-R_0(z)\Big) \leq\max\{\sigma(R_0), \deg h-1\}<\deg h. \end{equation} From \eqref{041} and \eqref{042}, we see that $\sigma(\Delta f)=\sigma(f)$, and $\Delta f(z)$ has two Borel exceptional values $0$ and $\infty$. Substituting $\Delta f(z)=R_{1}(z)e^{h(z)}$ into \eqref{019}, we have \begin{equation}\label{042z} R_{1}(z+1)e^{h(z+1)-h(z)}=R_{1}(z)(G(z)+1). \end{equation} If $\deg h\geq2$, then $\sigma(e^{h(z+1)-h(z)})=\deg h-1\geq1$. By \eqref{042z} and Lemma \ref{le11}, for any given $\varepsilon>0$, we have \begin{align*} m(r,e^{h(z+1)-h(z)}) &\leq m\Big(r,\frac{R_{1}(z)}{R_{1}(z+1)}\Big)+m(r, G(z)+1)\\ &=O(r^{\sigma(R_{1})-1+\varepsilon})+O(\log r), \end{align*} which yields $\deg h-1\leq \sigma(R_{1})-1+\varepsilon$. Letting $\varepsilon\to 0$, we have $\deg h\leq\sigma(R_{1})$, which contradicts with \eqref{042}. Hence, if $\deg h\geq2$, then $f(z)$ has at most one Borel exceptional value. If $\deg h=1$, then $F(z)=\Delta f(z)=R_{1}(z)e^{az}$, where $a\in\mathbb{C}\setminus\{0\}$. If $R_{1}(z)$ is transcendental with $\sigma(R_{1})<1$, by Lemma \ref{le4}, we see $G(z)=\frac{\Delta^{2}f(z)}{\Delta f(z)}=\frac{\Delta F(z)}{F(z)}$ is also transcendental. This contradicts with the fact $G(z)$ is a rational function. Therefore, $R_{1}(z)$ is a rational function. Combining this with \eqref{040} and \eqref{041}, we have \begin{equation}\label{043z} f(z)=b+R_0(z)e^{az} \end{equation} and $$ R_{1}(z)=e^{a}R_0(z+1)-R_0(z), $$ where $\sigma(R_0)<1$. If $R_0(z)$ is transcendental, by Lemma \ref{le10}, we see $e^{a}R_0(z+1)-R_0(z)$ is transcendental, which contradicts with $R_{1}(z)=e^{a}R_0(z+1)-R_0(z)$ is a rational function. Hence, $R_0(z)$ is a rational function. (iv) Suppose that $f(z)$ is a meromorphic solution of equation \eqref{eq1}, then $g(z)=\Delta f(z)$ is a meromorphic solution of equation \eqref{eq3}. Checking the proof of (i), we see if $g(z)$ is a rational solution of \eqref{eq3}, then $p-q+2k\leq0$. Since $p-q+2k>0$, we know $\Delta f(z)$ is transcendental. \eqref{019} still hold. By \eqref{019}, set $$ P(z,\Delta f):=\Delta^{2}f(z)-G(z)\Delta f(z)=0. $$ Since $G(z)$ is a nonconstant rational function, then for any given $a\in\mathbb{C}\setminus\{0\}$, we have $P(z,a)=-aG(z)\not\equiv0$. Together with Lemma \ref{le3} , we have $m\big(r,\frac{1}{\Delta f-a}\big)=S(r,\Delta f)$. Thus, $\delta(a, \Delta f)=0$. By this and the proof of (iii), we see taht $\Delta f(z)$ has at most one Borel exceptional value $0$ or $\infty$ unless \begin{equation}\label{047} \Delta f(z)=R_{1}(z)e^{az} \end{equation} where $a\in\mathbb{C}\setminus\{0\}, R_{1}(z)$ is a nonzero rational function. Now we prove that $a\neq i2k_{1}\pi$ for any $k_{1}\in\mathbb{Z}$. We see $R_{1}(z)$ satisfies \begin{equation}\label{044} \frac{R_{1}(z+2)}{R_{1}(z)}\to 1, \quad \frac{R_{1}(z+1)}{R_{1}(z)}\to 1, \quad z\to \infty. \end{equation} By \eqref{047}, we have \begin{equation}\label{048} \begin{gathered} \Delta^{2}f(z)=\Delta(\Delta f(z))=e^{az}(e^{a}R_{1}(z+1)-R_{1}(z)),\\ \Delta^{3}f(z)=\Delta^{2}(\Delta f(z)) =e^{az}(e^{2a}R_{1}(z+2)-2e^{a}R_{1}(z+1)+R_{1}(z)). \end{gathered} \end{equation} From \eqref{047}--\eqref{048}, we deduce that \begin{align*}% \label{045} &\frac{\Delta^{3}f(z)}{\Delta f(z)} -\frac{3}{2}\Big(\frac{\Delta^{2}f(z)}{\Delta f(z)}\Big)^{2}\\ &=e^{2a}\frac{R_{1}(z+2)}{R_{1}(z)} -\frac{3}{2}e^{2a}\Big(\frac{R_{1}(z+1)}{R_{1}(z)}\Big)^{2} +e^{a}\frac{R_{1}(z+1)}{R_{1}(z)}-\frac{1}{2} \\ &\to e^{2a}-\frac{3}{2}e^{2a}+e^{a}-\frac{1}{2} =-\frac{1}{2}(e^{a}-1)^{2},\quad z\to \infty. \end{align*} Combining this with \eqref{eq1}, we have $$ \Big[\frac{\Delta^{3}f(z)}{\Delta f(z)}-\frac{3}{2} \Big(\frac{\Delta^{2}f(z)}{\Delta f(z)}\Big)^{2}\Big]^{k} =R(z)\to \frac{(-1)^{k}}{2^{k}}(e^{a}-1)^{2k}, \quad z\to \infty. $$ If $e^{a}=1$, by \eqref{048}, we have $$ \Delta^{2}f(z)=e^{az}\Delta R_{1}(z),\quad \Delta^{3}f(z)=e^{az}\Delta^{2}R_{1}(z). $$ Combining this with \eqref{eq1} and \eqref{047}, we obtain \[ \Big[\frac{\Delta^{3}f(z)}{\Delta f(z)} -\frac{3}{2}\Big(\frac{\Delta^{2}f(z)}{\Delta f(z)}\Big)^{2}\Big]^{k} = \Big[\frac{\Delta^{2}R_{1}(z)}{R_{1}(z)} -\frac{3}{2}\Big(\frac{\Delta R_{1}(z)}{R_{1}(z)}\Big)^{2}\Big]^{k}=R(z). \] Hence, $R_{1}(z)$ is a rational solution of the equation \begin{equation}\label{046} \Big[\frac{\Delta^{2}g(z)}{g(z)}-\frac{3}{2}\Big(\frac{\Delta g(z)}{g(z)}\Big)^{2} \Big]^{k}=R(z). \end{equation} By the conclusion of (i), we see if $p-q+2k>0$, equation \eqref{046} has no rational solutions. It is a contradiction. Thus, $e^{a}\neq1$. So, $a\neq i2k_{1}\pi$ for any $k_{1}\in\mathbb{Z}$. \end{proof} \subsection*{Acknowledgements} The project was supported by the National Natural Science Foundation of China (11171119). \begin{thebibliography}{00} \bibitem{be} W. Bergweiler, J. K. Langley; Zeros of differences of meromorphic functions. \emph{Math. Proc. Cambridge Philos. Soc.} {\bf 142} (2007), 133-147. \bibitem{ch1} Y. M. Chiang, S. J. Feng; On the Nevanlinna characteristic of $f(z+\eta)$ and difference equations in the complex plane, \emph{Ramanujan J.}, {\bf 16} (2008), 105--129. \bibitem{ch2} B. Q. Chen, S. Li; Admissible solutions of the Schwarzian type difference equation, \emph{Abstr. Appl. Anal.}, {\bf 2014} (2014), Article ID 306360. \bibitem{ch3} Z. X. Chen; \emph{Complex Differences and Difference Equations}, Mathematics Monograph Series 29, Science Press, Beijing, 2014. \bibitem{ha1} R. G. Halburd, R. J. Korhonen; Difference analogue of the Lemma on the Logarithmic Derivative with applications to difference equations, \emph{J. Math. Anal. Appl.}, {\bf 314} (2006), 477--487. \bibitem{ha2} W. K. Hayman; \emph{ Meromorphic Functions.} Clarendon Press, Oxford, 1964. \bibitem{is} K. Ishizaki; Admissible solutions of the Schwarzian differential equation, AustralianMathematical SocietyA: \emph{Pure Mathematics and Statistics,} {\bf 50}, (2) (1991), pp. 258-278. \bibitem{la1} I. Laine, C. C. Yang; Clunie theorems for difference and q-difference polynomials, \emph{J. Lond. Math. Soc.} {\bf 76} (2007), 556--566. \bibitem{la2} S. T. Lan, Z. X. Chen; Relationships between characteristic functions of $\Delta^{n}f$ and $f$, \emph{J. Math. Anal. Appl.,} {\bf 412}(2014), 922--942. \bibitem{li} L. W. Liao, Z. Ye; On the growth of meromorphic solutions of the Schwarzian differential equations, \emph{J. Math. Anal. Appl.,} {\bf 309} (2005), 91--102. \bibitem{wh} J. M. Whittaker; \emph{Interpolatory Function Theory}, Cambridge Tracts in Math. and Math. Phys., vol. 33, Cambridge University Press, 1935. \bibitem{ya} L. Yang; \emph{Value distribution theory and its new research} (in Chinese). Beijing: Science Press. 1982. \end{thebibliography} \end{document}