Hostname: page-component-7bb8b95d7b-dtkg6 Total loading time: 0 Render date: 2024-09-27T03:06:30.164Z Has data issue: false hasContentIssue false

NOTES ON FERMAT-TYPE DIFFERENCE EQUATIONS

Published online by Cambridge University Press:  03 June 2024

ILPO LAINE
Affiliation:
Department of Physics and Mathematics, University of Eastern Finland, P.O. Box 111, FI-80101, Joensuu, Finland e-mail: ilpo.laine@uef.fi
ZINELAABIDINE LATREUCH*
Affiliation:
National Higher School of Mathematics, Scientific and Technology Hub of Sidi Abdellah, P.O. Box 75, Algiers 16093, Algeria
Rights & Permissions [Opens in a new window]

Abstract

We consider the existence problem of meromorphic solutions of the Fermat-type difference equation

$$ \begin{align*} f(z)^p+f(z+c)^q=h(z), \end{align*} $$

where $p,q$ are positive integers, and h has few zeros and poles in the sense that $N(r,h) + N(r,1/h) = S(r,h)$. As a particular case, we consider $h=e^g$, where g is an entire function. Additionally, we briefly discuss the case where h is small with respect to f in the standard sense $T(r,h)=S(r,f)$.

Type
Research Article
Copyright
© The Author(s), 2024. Published by Cambridge University Press on behalf of Australian Mathematical Publishing Association Inc.

1. Introduction and motivation

Throughout this note, we assume familiarity with standard notation and key results from Nevanlinna theory [Reference Hayman7]. In particular, a meromorphic function $\varphi $ is termed a small function of f if the Nevanlinna characteristic $T(r, \varphi )$ satisfies $T(r, \varphi ) = S(r, f)$ . Here, $S(r, f)$ denotes any quantity such that $S(r, f) = o(T(r, f))$ as $r \rightarrow \infty $ , potentially outside an exceptional set of finite linear or logarithmic measure. For brevity, we denote the set of all meromorphic functions that are small with respect to f by $\mathcal {S}(f)$ . The order and hyper-order of a meromorphic function f are defined respectively by

$$ \begin{align*} \rho(f):=\limsup _{r \rightarrow \infty} \frac{\log T(r, f)}{\log r} \quad \text{and}\quad \rho_2(f):=\limsup _{r \rightarrow \infty} \frac{\log \log T(r, f)}{\log r}. \end{align*} $$

In 1970, Yang [Reference Yang16] applied Nevanlinna’s value distribution theory to investigate the Fermat functional equation

(1.1) $$ \begin{align} a(z)f(z)^p+b(z)g(z)^q=1, \end{align} $$

where $p, q \geq 3$ are integers, f and g denote nonconstant meromorphic functions, and $a, b$ are meromorphic functions that are respectively small compared with f and g. Yang [Reference Yang16, Theorem 1] showed that (1.1) cannot hold unless $p=q=3$ . However, if f and g are entire, then (1.1) cannot hold even if $p=q=3$ . Indeed, upon careful inspection of the proof of [Reference Yang16, Theorem 1], one can express Yang’s result in terms of

$$ \begin{align*} \Theta(a, f):=1-\limsup _{r \rightarrow \infty} \frac{\overline{N}(r, {1}/{(f-a)})}{T(r, f)} \end{align*} $$

as follows.

Theorem 1.1. Let p and q be positive integers satisfying ${1}/{p} + {1}/{q} < {2}/{3}$ . Then, (1.1) has no nonconstant meromorphic solutions f and g. Moreover, if ${1}/{p} + {1}/{q} < 1$ , there exists no nonconstant meromorphic solution of (1.1) such that $\Theta (\infty , f) = \Theta (\infty , g) = 1$ .

Motivated by this and the development of difference analogues of Nevanlinna’s theory (see, for example, [Reference Chen3, Reference Halburd, Korhonen and Tohge6, Reference Liu, Laine and Yang12]), there are many explorations of the Fermat-type difference equation

(1.2) $$ \begin{align} f(z)^p+f(z+c)^q=h(z), \end{align} $$

where c is a nonzero constant and h is a given meromorphic function. For instance, it has been shown in [Reference Korhonen and Zhang8, Reference Lü and Han14] that if h is a nonzero constant and $p=q=3$ , then (1.2) does not admit nonconstant meromorphic solutions with $\rho _{2}(f)<1$ . Later, Lü and Guo [Reference Lü and Guo13] extended this conclusion to the case when $h = e^{\alpha z+\beta }$ , under the conditions $p \geq 3$ and $q \geq 2$ or $p \geq 2$ and $q \geq 3$ , excluding the exceptional case of trivial solutions, which applies when $p=q$ and

$$ \begin{align*} f(z)=\mu e^{(\alpha z+\beta) / p} \quad \text {with}\quad \mu^p(1+e^{\alpha c})=1. \end{align*} $$

For the case $h = e^{g}$ , where g is a nonconstant polynomial, Bi and Lü [Reference Bi and Lü2] employed properties of elliptic functions to establish the nonexistence of nontrivial meromorphic solutions with $\rho _{2}(f)<1$ for (1.2) when $p=q=3$ . Additionally, they proved the same result [Reference Bi and Lü2, Theorem 2] when $h\in \mathcal {S}(f)$ and has $0$ and $\infty $ as Borel exceptional values. More recently, Guo and Liu [Reference Guo and Liu5] investigated the case $h = e^{g}$ in (1.2), with g being a nonconstant entire function.

Motivated by these results, we explore the existence problem of meromorphic solutions to (1.2) when h belongs to some specific classes. In Section 2, we investigate the case where h has few zeros and poles, in the sense that

(1.3) $$ \begin{align} N(r,h)+N\bigg( r,\frac{1}{h}\bigg) =S(r,h). \end{align} $$

We will refer to the class of functions that satisfy (1.3) as $\mathcal {A}$ . In particular, we will consider the case $h=e^g$ , where g is a nonconstant entire function. In Section 3, we discuss the case when h is small with respect to the solution f in the standard sense of $T(r,h)=S(r,f)$ . Indeed, we show that a large class of meromorphic functions cannot satisfy the equation

(1.4) $$ \begin{align} f(z)^{p}+f(z+c)^{p}=h(z), \end{align} $$

where $p\geq 3$ and h has two Borel exceptional values.

2. The case $h\in \mathcal {A}$

In this section, we consider (1.2) where p and q are distinct positive integers and h belongs to the class $\mathcal {A}$ . Note that all functions in $\mathcal {A}$ must be transcendental meromorphic functions. Then, we will examine the case $h=e^P$ , with P a nonconstant entire function.

Before we proceed to state our main results, we require the following elementary lemma.

Lemma 2.1. Let $p,q$ be distinct positive integers and let h be a meromorphic function. If f is a meromorphic function with $\rho _{2}(f)<1$ that satisfies (1.2), then

$$ \begin{align*} N(r,f)\asymp N(r,h) \quad \text{and} \quad T(r,f)\asymp T(r,h), \end{align*} $$

outside an exceptional set of finite logarithmic measure. In particular, $\rho _{2}(h)=\rho _{2}(f)$ .

Proof. Without loss of generality, we assume that $p>q$ . From [Reference Liu, Laine and Yang12, Lemma 1.2.10],

$$ \begin{align*} (p-q-o(1))T(r,f)\leq T(r,h)\leq (p+q+o(1))T(r,f) \end{align*} $$

and

$$ \begin{align*} (p-q-o(1))N(r,f)\leq N(r,h)\leq (p+q+o(1))N(r,f), \end{align*} $$

outside an exceptional set of finite logarithmic measure. By a standard argument, see [Reference Gundersen4, Lemma 5], we may remove the exceptional set to obtain $\rho _{2}(h)=\rho _{2}(f)$ .

Now, we are ready to state our first result.

Theorem 2.2. Let $h \in \mathcal {A}$ and let $p>q$ be positive integers. Then, (1.2) has no meromorphic solutions with $\rho _{2}(f) < 1$ except when $p=2$ and $q=1$ . In this case, either:

  1. (1) $f(z)=e^{\alpha z+\beta }-1$ , where $\alpha , \beta \in \mathbb {C}$ satisfy $e^{\alpha c}=2$ ; or

  2. (2) $T(r,f)= \overline {N}( r,1/f') +S(r,f).$

Remark 2.3. (a) According to [Reference Guo and Liu5, Remark 1.7], we observe that conclusion (1) of Theorem 2.2 might hold. In fact, an example such as $f(z) = e^{\alpha z} - 1$ solves the equation

$$ \begin{align*}f(z)^{2}+f(z+c)=e^{2\alpha z},\end{align*} $$

where $\alpha $ is a nonzero constant satisfying $e^{\alpha c} = 2$ .

(b) Theorem 2.2 includes the result given in [Reference Guo and Liu5, Theorem 1.6(iii)]. Indeed, it was shown there that the equation

$$ \begin{align*} f(z)^2+f(z+c)=e^{g(z)}, \end{align*} $$

where g is a nonconstant entire function, has no meromorphic solutions with $\rho _2(f) < 1$ satisfying $N(r,1/f') = S(r,f)$ except functions f of the type

(2.1) $$ \begin{align} f(z)=\bigg(\int \frac{a}{2} e^{-g(z) / 2} \,dz+b\bigg) e^{g(z) / 2},\quad a,b \in\mathbb{C}. \end{align} $$

However, according to Theorem 2.2(2), $N(r,1/f') = S(r,f)$ implies $T(r,f)=S(r,f)$ , which is impossible. Hence, (2.1) reduces to Theorem 2.2(1).

(c) To find a concrete example for the conclusion of Theorem 2.2(2) remains open at present. Note that if conclusion (2) holds, then for any $a\in \mathbb {C}$ , we have $\delta (a,f)=0$ . This can be deduced from the estimate

$$ \begin{align*} m\bigg( r,\frac{1}{f-a}\bigg) &\leq m\bigg( r,\frac{1}{f'}\bigg) +S(r,f)\nonumber\\ &=T(r,f')-N\bigg( r,\frac{1}{f'}\bigg)+S(r,f) \nonumber\\ &\leq T(r,f)-N\bigg( r,\frac{1}{f'}\bigg)+S(r,f)=S(r,f).\nonumber \end{align*} $$

Proof of Theorem 2.2.

Clearly, from Lemma 2.1, f must be a transcendental meromorphic function with $N(r,f)=S(r,f)$ . Differentiating (1.2) gives

(2.2) $$ \begin{align} f(z)^{p-1}\bigg( pf'(z)-\frac{h'(z)}{h(z)}f(z)\bigg) =- f(z+c)^{q-1}\bigg( qf'(z+c)-\frac{h'(z)}{h(z)}f(z+c)\bigg). \end{align} $$

Denote

(2.3) $$ \begin{align} \varphi(z):=pf'(z)-\frac{h'(z)}{h(z)}f(z). \end{align} $$

Note that

$$ \begin{align*} T\bigg( r,\frac{h'}{h}\bigg) =\overline{N}(r,h)+\overline{N}\bigg( r,\frac{1}{h}\bigg)+S(r,h)=S(r,h). \end{align*} $$

This, along with $N(r,f)=S(r,f)$ , leads to $N(r,\varphi )=S(r,f)$ .

Since $p\geq q+1$ , we may recall the delay-difference variant of the Clunie lemma (see [Reference Laine and Yang10] and [Reference Liu, Laine and Yang12, page 20]) to obtain $m(r,\varphi )=S(r,f)$ and hence $T(r,\varphi )=S(r,f)$ .

Suppose first that $\varphi \equiv 0$ . Then, by simple integration,

(2.4) $$ \begin{align} f(z)^p=Ah(z) \end{align} $$

for some constant A. However, from (2.2),

$$ \begin{align*} qf'(z+c)-\frac{h'(z)}{h(z)}f(z+c)\equiv 0, \end{align*} $$

which in turn implies that

(2.5) $$ \begin{align} f(z+c)^q=Bh(z), \end{align} $$

where B is a constant. Recalling again [Reference Liu, Laine and Yang12, Lemma 1.2.10], and combining (2.4) and (2.5),

$$ \begin{align*} (p-q)T(r,f)=S(r,f), \end{align*} $$

which is a contradiction. Thus, we may now assume that $\varphi \not \equiv 0$ and (2.3) can be rewritten as

(2.6) $$ \begin{align} \frac{1}{f(z)}=\frac{1}{\varphi(z)}\bigg( p\frac{f'(z)}{f(z)}-\frac{h'(z)}{h(z)}\bigg). \end{align} $$

This implies

$$ \begin{align*} T(r,f)=N\bigg(r,\frac{1}{f}\bigg) +S(r,f). \end{align*} $$

Note that if $z_0$ is a multiple zero of f that is not a zero or pole of h, then $z_0$ is a zero of $\varphi $ . Hence, the contribution of $N_{2)}(r,1/f)$ is $S(r,f)$ , and so

(2.7) $$ \begin{align} T(r,f)=N_{1)}\bigg(r,\frac{1}{f}\bigg) +S(r,f). \end{align} $$

If $p\geq q+2$ , then we may write (2.2) as

$$ \begin{align*} f(z)^{p-2}(f(z)\varphi(z))=- f(z+c)^{q-1}\bigg( qf'(z+c)-\frac{h'(z)}{h(z)}f(z+c)\bigg). \end{align*} $$

Applying again the delay-difference Clunie lemma to conclude that

$$ \begin{align*}m(r,f\varphi )=T(r,f\varphi )=S(r,f),\end{align*} $$

we obtain

$$ \begin{align*}T(r,f)=m\bigg( r,\frac{f\varphi}{\varphi} \bigg) \leq m(r,f\varphi )+T(r,\varphi )+O(1)=S(r,f),\end{align*} $$

which is a contradiction.

So, we must have $p=q+1$ . Since $h\in \mathcal {A}$ , we can write $h=\pi e^{g}$ , where g is an entire function and $\pi $ is a small function of h. Consequently, $T(r,\pi )=S(r,f)$ . Therefore, (1.2) can be rewritten as

(2.8) $$ \begin{align} F(z)^p+G(z) ^{p-1}=\pi(z), \end{align} $$

where $F:=f/e^{g/p}$ and $G:=f_c/e^{g/(p-1)},$ where $f_c:=f(z+c)$ . To apply Theorem 1.1 to (2.8), we have to show that $\pi $ is a small function of F and G. Indeed, from (2.7), we have $ N(r,1/f) =~(1-~o(1))~T(r,f) $ outside of an exceptional set E of finite linear measure. Therefore, provided $r\not \in E$ ,

$$ \begin{align*} \frac{T(r,\pi)}{T(r,f/\exp (g/p))}&=\frac{T(r,\pi)}{T(r,\exp (g/p)/f)+O(1)} =(1+o(1))\frac{T(r,\pi)}{T(r,\exp (g/p)/f)}\\ &\leq (1+o(1))\frac{T(r,\pi)}{N(r,\exp (g/p)/f)} =(1+o(1))\frac{T(r,\pi)}{N(r,1/f)}\\ &=\frac{1+o(1)}{1-o(1)}\frac{T(r,\pi)}{T(r,f)}=o(1). \end{align*} $$

Recalling that $T(r,f_c)=T(r,f)+S(r,f)$ , we may use a similar reasoning to conclude that $\pi $ is a small function of $f_c/e^{g/(p-1)}$ as well.

In the remainder of the proof, we discuss cases according to the values of p.

Case 1. If $p\geq 4$ , then according to Theorem 1.1, (2.8) admits only constant solutions. Therefore,

(2.9) $$ \begin{align} f(z)=\alpha e^{g(z)/p}, \quad f(z+c)=\beta e^{g(z)/(p-1)}, \end{align} $$

where $\alpha $ , $\beta $ are constants that satisfy $\alpha ^p+\beta ^{p-1}=\pi $ . Now,

$$ \begin{align*} f(z+c)=\alpha e^{g(z+c)/p}=\beta e^{g(z)/(p-1)}, \end{align*} $$

which means that $(p-1)g_c-pg$ is a constant. If g is a polynomial, then it must be a constant, which contradicts the fact that $h=\pi e^g$ is transcendental. If g is not a polynomial, then it satisfies the equation

$$ \begin{align*} \frac{g'(z+c)}{g'(z)}=\frac{p}{p-1}, \end{align*} $$

and by using [Reference Bergweiler and Langley1, Lemma 3.3], we deduce that $\rho (g)\geq 1$ . Therefore, $\rho _{2}(f)=\rho _{2}(h)\geq 1,$ which is again a contradiction.

Case 2. If $p=3$ , then by recalling that $N(r,F)=N(r,G)=N(r,f)=S(r,f)$ , we can infer that

$$ \begin{align*} \Theta(\infty,F)=\Theta(\infty,G)=\Theta(\infty,f)=1. \end{align*} $$

Hence, by applying Theorem 1.1 again, we conclude that (1.2) has solutions only in the form of (2.9) with $\alpha ^3+\beta ^2=\pi $ . Similar reasoning as in the case $p\geq 4$ leads to a contradictory conclusion $\rho _{2}(f)\geq 1$ .

Case 3. Assume now that $p=2$ (implying that we also have $q=1$ ). Recall that the auxiliary function $\varphi \not \equiv 0$ takes the form

(2.10) $$ \begin{align} \varphi(z):=2f'(z)-\frac{h'(z)}{h(z)}f(z). \end{align} $$

Taking the first derivative of $\varphi $ yields

$$ \begin{align*} \varphi'(z)=2f"(z)-\frac{h'(z)}{h(z)}f'(z)-\bigg( \frac{h'(z)}{h(z)}\bigg) 'f(z)=\frac{\varphi'(z)}{\varphi(z)}\bigg(2f'(z)-\frac{h'(z)}{h(z)}f(z) \bigg), \end{align*} $$

which can be expressed as

$$ \begin{align*} A(z)f'(z)+B(z) f(z)=0, \end{align*} $$

where

$$ \begin{align*} A(z):=2\frac{f"(z)}{f'(z)}-\frac{h'(z)}{h(z)}-2\frac{\varphi'(z)}{\varphi(z)} \quad \text{and}\quad B(z):=\frac{\varphi'(z)}{\varphi(z)}\frac{h'(z)}{h(z)}-\bigg( \frac{h'(z)}{h(z)}\bigg)'. \end{align*} $$

Since f mainly has simple zeros, assume that $z_0$ is a simple zero of f that is neither a zero nor a pole of h. Consequently, $z_0$ is not a zero of $\varphi $ and so not a pole of B. Thus, $z_0$ is a zero of A. Now, define the function

$$ \begin{align*} \psi(z):=\frac{A(z)}{f(z)}. \end{align*} $$

Based on the preceding discussion, we can deduce that $N(r, \psi ) = \overline {N}(r,1/f') + S(r, f)$ . Combining this with the fact that $m(r,1/f) = S(r,f)$ (see (2.6)), we conclude that

$$ \begin{align*}T(r,\psi)=\overline{N}\bigg( r,\frac{1}{f'}\bigg) +S(r,f).\end{align*} $$

Note that, according to [Reference Yang and Yi17, Theorem 1.24],

$$ \begin{align*} N\bigg(r, \frac{1}{f'}\bigg) \leq N\bigg(r, \frac{1}{f}\bigg)+ \overline{N}(r, f)+S(r, f)= N_{1)}\bigg( r,\frac{1}{f}\bigg) +S(r, f). \end{align*} $$

Assume first that $\psi \not \equiv 0$ . Then,

$$ \begin{align*} T(r,f)=m(r,f)+S(r,f)&=m\bigg( r,\frac{A}{\psi}\bigg)+S(r,f)\nonumber\\ &\leq T\bigg( r,\frac{1}{\psi}\bigg) -N\bigg( r,\frac{1}{\psi}\bigg) +S(r,f)\nonumber\\ &=\overline{N}\bigg( r,\frac{1}{f'}\bigg) -N\bigg( r,\frac{1}{\psi}\bigg) +S(r,f)\nonumber\\ &\leq N_{1)}\bigg( r,\frac{1}{f}\bigg) -N\bigg( r,\frac{1}{\psi}\bigg) +S(r,f).\nonumber \end{align*} $$

Combining this with (2.7) yields $N(r,1/\psi )=S(r,f)$ , that is,

$$ \begin{align*} T(r,f)= \overline{N}\bigg( r,\frac{1}{f'}\bigg) +S(r,f). \end{align*} $$

Consider now the case $\psi \equiv 0$ . This implies that both $A \equiv 0$ and $B \equiv 0$ . To proceed, we distinguish two possible sub-cases.

Subcase (i). If $\varphi $ is not constant, then straightforward integration of the equations $A \equiv 0$ and $B \equiv 0$ yields

(2.11) $$ \begin{align} (f'(z))^2=C_1\varphi(z)^2h(z) \end{align} $$

and

(2.12) $$ \begin{align} \varphi(z)=C_2\frac{h'(z)}{h(z)}, \end{align} $$

where $C_1$ and $C_2$ are nonzero constants. Substituting (2.12) into (2.10) and integrating the result yields

(2.13) $$ \begin{align} (f(z)+C_2)^2=C_3h(z), \end{align} $$

where $C_3$ is a nonzero constant. Combining (2.11), (2.12) and (2.13) results in

$$ \begin{align*} \frac{ f'(z)}{f(z)+C_2}=C_4\frac{h'(z)}{h(z)}, \end{align*} $$

where $C_4$ is a nonzero constant. By simple integration,

$$ \begin{align*} f(z)+C_2=C_5h(z). \end{align*} $$

This, together with (2.13) results in a contradiction.

Now assume that $\varphi $ is a nonzero constant. Keeping in mind that $A\equiv 0$ and $B\equiv 0$ , we find $h'/h$ is a constant, say $h'/h\equiv 2\alpha $ . This results in $h\equiv C_{h}e^{2\alpha z},$ with $C_{h}$ being a constant. Since $2{f"}/{f'}\equiv {h'}/{h}$ , by elementary integration,

$$ \begin{align*}(f'(z))^{2}=C_{6}h(z)=C_{6}C_{h}e^{2\alpha z}.\end{align*} $$

This means that $f'\equiv C_{7}e^{\alpha z}$ and so, by integration, f must be of the form $e^{\alpha z+\beta }+\gamma ,$ with constants $\beta , \gamma $ . Substituting this into (1.2),

$$ \begin{align*}e^{2\alpha z+2\beta}+2\gamma e^{\alpha z+\beta}+\gamma^{2}+e^{\alpha c}e^{\alpha z+\beta}+\gamma =C_{h}e^{2\alpha z}.\end{align*} $$

Clearly, the constant term $\gamma ^{2}+\gamma $ must vanish; thus, $\gamma = 0$ or $\gamma = -1$ . However, $\gamma \neq 0$ , as it would contradict (2.7). Therefore, the only possibility is $\gamma = -1$ , leading to

$$ \begin{align*}(e^{2\beta}-C_{h})e^{2\alpha z}+(e^{\alpha c}-2)e^{\alpha z+\beta}=0.\end{align*} $$

This may happen if $e^{\alpha c}=2$ and $C_{h}=e^{2\beta }$ , thereby completing the proof.

We close this section by considering the equation

(2.14) $$ \begin{align} f(z)^p+f(z+c)^q=e^{g(z)}, \end{align} $$

where g is a nonconstant entire function and $p, q$ are positive integers that are not necessarily distinct. The following corollary improves [Reference Guo and Liu5, Theorem 1.6(iv)].

Corollary 2.4. Let g be a nonconstant entire function and let $p,q\geq 3$ be integers. Then:

  1. (1) if $p\neq q$ , (2.14) has no meromorphic solutions with $\rho _{2}(f) < 1$ ;

  2. (2) if $p=q\geq 3$ , (2.14) has no nontrivial meromorphic solutions with $\rho _{2}(f) < 1$ .

Proof. Assertion (1) immediately follows from Theorem 2.2.

As for assertion (2), if $p=q\geq 4$ , we may write (2.14) as

$$ \begin{align*} \bigg(\frac{f(z)}{e^{g(z)/p}}\bigg)^p + \bigg(\frac{f(z+c)}{e^{g(z)/p}}\bigg)^p = 1. \end{align*} $$

By applying Theorem 1.1, we obtain $ f(z) = \alpha e^{g(z)/p}$ and $f(z+c) = \beta e^{g(z)/p}$ , where ${\alpha ^p + \beta ^p = 1}$ . A reasoning similar to Case 1 in the proof of Theorem 2.2 leads to the conclusion that f takes the form

$$ \begin{align*}f(z)=\alpha \exp\bigg( \frac{az+b}{p}\bigg),\end{align*} $$

where $\alpha $ is a nonzero constant satisfying the condition $\alpha ^p(1 + e^{ac}) = 1$ . This representation corresponds to a trivial solution.

Finally, we assume that $p=q=3$ . Then (2.14) can be written as follows:

$$ \begin{align*} F(z)^3+\beta(z)F(z+c)^3 =1, \end{align*} $$

where

$$ \begin{align*}F(z)=\frac{f(z)}{e^{g(z) / 3}} \quad \text{and}\quad \beta(z)=e^{g(z+c)-g(z)}.\end{align*} $$

By making use of [Reference Liu, Laine and Yang12, Lemma 1.2.10],

$$ \begin{align*} T(r,e^g)\leq (3+o(1))T(r,f) \end{align*} $$

possibly outside an exceptional set of finite logarithmic measure $E_1$ and, moreover, $\rho _{2}(e^g)=\delta <1$ . Note that [Reference Halburd, Korhonen and Tohge6, Theorem 5.1] yields

$$ \begin{align*} T(r,\beta)=m\bigg( r,\frac{e^{g(z+c)}}{e^{g(z)}}\bigg) =o\bigg( \frac{T(r,e^g)}{r^{1-\delta-\varepsilon}}\bigg) \end{align*} $$

for all r outside an exceptional set of finite logarithmic measure $E_2$ . Combining these inequalities yields, for all $r\not \in E_1 \cup E_2$ ,

$$ \begin{align*} \frac{T(r,\beta)}{T(r,f)} \leq (3+o(1))\frac{T(r,\beta)}{T(r,e^g)}=o\bigg( \frac{1}{r^{1-\delta-\varepsilon}}\bigg), \end{align*} $$

implying $T(r,\beta )=S(r,f)$ . Consequently, following the same proof as in Case (iv) of [Reference Guo and Liu5, Theorem 1.6], we conclude that $\beta $ is a constant and $f = Ae^{g/3}$ with $A^3 = 1$ , constituting a trivial solution.

3. The case $h\in \mathcal {S}(f)$

In this section, we discuss the existence of solutions to (1.2), where h is a small meromorphic function in the standard sense of $T(r,h)=S(r,f)$ . For clarity, we first present the following proposition which discusses the existence of nonconstant meromorphic solutions to (1.2) when p and q are distinct integers.

Proposition 3.1. Let f be a meromorphic function and let $h\in \mathcal {S}(f)$ .

  1. (1) If $p> q \geq 3$ , then (1.2) has no nonconstant meromorphic solutions.

  2. (2) If $p> q \geq 2$ , then (1.2) has no nonconstant entire solutions.

  3. (3) If $p> q \geq 1$ and $\rho _{2}(f) < 1$ , then (1.2) has no nonconstant meromorphic solutions.

Proof. The cases (1) and (2) can be deduced easily from Theorem 1.1. Suppose now that $p>q\geq 1$ and f is a solution to (1.2) with $\rho _{2}(f)<1$ . We may write (1.2) in the form

$$ \begin{align*}f(z)^{p}+\bigg(\frac{f(z+c)}{f(z)}\bigg)^{q}f(z)^{q}=h (z). \end{align*} $$

By [Reference Halburd, Korhonen and Tohge6, Theorem 5.1] (see also [Reference Liu, Laine and Yang12, Lemma 1.2.8]), we conclude that

$$ \begin{align*}p m(r,f)\leq q m(r,f)+S(r,f),\end{align*} $$

and so $m(r,f)=S(r,f)$ . However, we can apply [Reference Halburd, Korhonen and Tohge6, Lemma 8.3] (see also [Reference Liu, Laine and Yang12, Lemma 1.2.10]) to (1.2). This yields

$$ \begin{align*}pN(r,f)\leq qN(r,f)+S(r,f),\end{align*} $$

thus implying $N(r,f) = S(r,f)$ . Consequently, $T(r,f) = S(r,f)$ , which is a contradiction.

Next, we consider the case where $p=q\geq 3$ in (1.2). In fact, referring to Theorem 1.1, it becomes evident that no nonconstant meromorphic solutions exist to (1.2) when $p=q>3$ . By combining this result with [Reference Bi and Lü2, Theorem 2], we can immediately deduce the following corollary.

Corollary 3.2. Let f be a meromorphic function of hyper-order $\rho _{2}(f)<1$ , and let $h\in \mathcal {S}(f)$ with two Borel exceptional values $0$ and $\infty .$ Then f does not satisfy (1.4) provided $p\geq 3$ .

Motivated by this, it is natural to ask whether this phenomenon remains valid when considering $h \in \mathcal {S}(f)$ with two Borel exceptional values $d \neq 0$ and $\infty $ .

Unfortunately, we could not give a positive answer to this question. However, the following proposition illustrates that such an assumption prevents the existence of a large class of meromorphic functions. In particular, it shows that any meromorphic solution to (1.4) must be periodic and cannot possess any Borel exceptional value.

Proposition 3.3. Let f be a meromorphic function of hyper-order $\rho _{2}(f)<1$ , and let $h\in \mathcal {S}(f)$ with two Borel exceptional values $d\in \mathbb {C}$ and $\infty $ . If f satisfies (1.4), then $p=3$ and the following assertions hold:

  1. (1) $h=e^{az+b}+d$ , where $a,b$ are constants with $e^{ac}=1$ and $d\neq 0$ ;

  2. (2) f is periodic and satisfies the equation

    $$ \begin{align*} f(z)^3=\pi(z)\exp\bigg( \frac{\log(-1)}{c}z\bigg) +\frac{1}{2}( e^{az+b}+d) , \end{align*} $$
    where $\pi $ is a c-periodic meromorphic function such that
    $$ \begin{align*}\lim\limits_{\substack{r\to \infty\\ r\not\in E}}\frac{T(r,\pi)}{r}=\infty,\end{align*} $$
    where E is an exceptional set of finite linear measure;
  3. (3) $T(r,f)=\overline {N}(r,f)+S(r,f)$ ;

  4. (4) $T(r,f)=\overline {N}(r,1/f)+S(r,f)$ . In addition, for any $\alpha \in \mathcal {S}(f)$ that does not satisfy (1.4),

    $$ \begin{align*} T(r,f)=N\bigg( r,\frac{1}{f-\alpha}\bigg) +S(r,f). \end{align*} $$

Proof of Proposition 3.3.

Assuming the existence of a meromorphic function f that satisfies (1.4), we infer from Theorem 1.1 and [Reference Bi and Lü2, Theorem 2] that $p=3$ and $d\neq 0$ .

Following the reasoning in the beginning of the proof of [Reference Bi and Lü2, Theorem 2] (see [Reference Bi and Lü2, page 6]), we conclude that h is periodic with period c and, consequently, $f^3$ is periodic with period $2c$ . Now, recall that h has two Borel exceptional values $d\neq 0$ and $\infty $ . Then we have the representation

$$ \begin{align*}h(z)=\pi(z)e^{P(z)}+d,\end{align*} $$

where P is an entire function, $\pi $ a meromorphic function of finite order and

$$ \begin{align*}\rho_{2}(h)=\rho_2(e^P)=\rho(P)<1.\end{align*} $$

Making use of the periodicity of h and [Reference Zemirni, Laine and Latreuch18, Proposition 3.1], we see that P must be a polynomial. If $\deg P\geq 2$ , then by [Reference Zemirni, Laine and Latreuch18, Proposition 3.1], we have $\rho (\pi )\geq \deg P$ . However,

$$ \begin{align*} \rho(\pi)=\lambda(h-d)<\rho(h)=\deg P, \end{align*} $$

which is a contradiction. Hence, $\deg P=1$ and $\pi $ has an order less than one. Using [Reference Zemirni, Laine and Latreuch18, Remark 3.1], we observe that $\pi $ is a constant. Therefore, $h\equiv e^{az+b}+d,$ where $a,b$ are constants with $e^{ac}=1$ . This proves assertion (1).

Now, noting that $\tfrac 12h$ is a particular solution to the difference equation

(3.1) $$ \begin{align} g(z)+g(z+c)=h(z) \end{align} $$

and making use of the theory of difference equations in the complex domain [Reference Meschkowski15, Ch. 7], we deduce that the general solution to (3.1) takes the form

$$ \begin{align*} g(z)=\pi(z)\exp\bigg( \frac{\log(-1)}{c}z\bigg) +\frac{1}{2}h(z), \end{align*} $$

where $\pi $ is a c-periodic meromorphic function. Therefore, assertion (2) follows by considering $g=f^3$ and $h\equiv e^{az+b}+d$ .

Next, since $h\in \mathcal {S}(f)$ , according to [Reference Li, Sabadini and Struppa11, Lemma 1], we derive assertion (3). Additionally, by employing Nevanlinna’s second main theorem,

$$ \begin{align*} T(r,f^3)&\leq \overline{N}(r,f^3)+\overline{N}\bigg( r,\frac{1}{f^3}\bigg) + \overline{N}\bigg( r,\frac{1}{f^3-h}\bigg) +S(r,f)\nonumber\\ &\leq 2T(r,f)+\overline{N}\bigg( r,\frac{1}{f_c^3}\bigg)+S(r,f).\nonumber \end{align*} $$

From this and [Reference Liu, Laine and Yang12, Lemma 1.2.10], we conclude the first part of assertion (4). The last part of assertion (4) is a direct consequence of [Reference Laine and Latreuch9, Lemma 2.2].

Acknowledgement

The authors are grateful to the referees for their valuable comments and suggestions.

References

Bergweiler, W. and Langley, J. K., ‘Zeros of differences of meromorphic functions’, Math. Proc. Cambridge Philos. Soc. 142(1) (2007), 133147.CrossRefGoogle Scholar
Bi, W. and , F., ‘On meromorphic solutions of the Fermat-type functional equations $f{(z)}^3+f{\left(z+c\right)}^3={e}^P$ ’, Anal. Math. Phys. 13 (2023), Article no. 24.CrossRefGoogle Scholar
Chen, Z. X., Complex Differences and Difference Equations (Science Press, Beijing, 2013).Google Scholar
Gundersen, G., ‘Finite order solutions of second order linear differential equations’, Trans. Amer. Math. Soc. 305 (1988), 415429.CrossRefGoogle Scholar
Guo, Y. and Liu, K., ‘Meromophic solutions of Fermat type differential and difference equations of certain types’, Ann. Polon. Math. 131 (2023), 119.CrossRefGoogle Scholar
Halburd, R. G., Korhonen, R. and Tohge, K., ‘Holomorphic curves with shift-invariant hyper-plane preimages’, Trans. Amer. Math. Soc. 366 (2014), 42674298.CrossRefGoogle Scholar
Hayman, W. K., Meromorphic Functions (Clarendon Press, Oxford, 1964).Google Scholar
Korhonen, R. and Zhang, Y. Y., ‘Existence of meromorphic solutions of first-order difference equations’, Constr. Approx. 51 (2020), 465504.CrossRefGoogle Scholar
Laine, I. and Latreuch, Z., ‘Remarks on meromorphic solutions of some delay-differential equations’, Anal. Math. 48 (2022), 10811104.CrossRefGoogle Scholar
Laine, I. and Yang, C. C., ‘Clunie theorems for difference and $q$ -difference polynomials’, J. Lond. Math. Soc. (2) 76 (2007), 556566.CrossRefGoogle Scholar
Li, B. Q., ‘On Fermat-type functional and partial differential equations’, in: The Mathematical Legacy of Leon Ehrenpreis, Springer Proceedings in Mathematics, 16 (eds. Sabadini, I. and Struppa, D. C.) (Springer, Milan, 2012), 209222.CrossRefGoogle Scholar
Liu, K., Laine, I. and Yang, L. Z., Complex Delay-Differential Equations (Walter de Gruyter, Berlin–Boston, 2021).CrossRefGoogle Scholar
, F. and Guo, H. X., ‘On meromorphic solutions of the Fermat-type functional equation $f{(z)}^n+f{\left(z+c\right)}^m={e}^{\alpha z+\beta }$ ’, Mediterr. J. Math. 19 (2022), Article no. 118.CrossRefGoogle Scholar
, F. and Han, Q., ‘On the Fermat-type equation $f{(z)}^3+f{\left(z+c\right)}^3=1$ ’, Aequationes Math. 91 (2017), 129136.CrossRefGoogle Scholar
Meschkowski, H., Differenzengleichungen (Vandenhoeck and Ruprecht, Göttingen, 1959).Google Scholar
Yang, C. C., ‘A generalization of a theorem of P. Montel on entire functions’, Proc. Amer. Math. Soc. 26 (1970), 332334.CrossRefGoogle Scholar
Yang, C. C. and Yi, H. X., Uniqueness Theory of Meromorphic Functions (Kluwer Academic Publishers, London, 2003).CrossRefGoogle Scholar
Zemirni, M. A., Laine, I. and Latreuch, Z., ‘New findings on the periodicity of entire functions and their differential polynomials’, Mediterr. J. Math. 20 (2023), Article no. 136.CrossRefGoogle Scholar