검색
Article Search

JMB Journal of Microbiolog and Biotechnology

OPEN ACCESS eISSN 0454-8124
pISSN 1225-6951
QR Code

Article

Kyungpook Mathematical Journal 2021; 61(3): 495-512

Published online September 30, 2021

Copyright © Kyungpook Mathematical Journal.

Value Distribution of L-functions and a Question of Chung-Chun Yang

Xiao-Min Li*, Qian-Qian Yuan, Hong-Xun Yi

Department of Mathematics, Ocean University of China, Qingdao, Shandong 266100, P. R. China
e-mail : lixiaomin@ouc.edu.cn and yuanqianqian92@163.com

Department of Mathematics, Shandong University, Jinan, Shandong 250199, P. R. China
e-mail : hxyi@sdu.edu.cn

Received: February 22, 2017; Revised: May 2, 2021; Accepted: May 18, 2021

We study the value distribution theory of L-functions and completely resolve a question from Yang [10]. This question is related to L-functions sharing three finite values with meromorphic functions. The main result in this paper extends corresponding results from Li [10].

Keywords: Nevanlinna theory, Meromorphic functions, L-functions, Shared Values, Uniqueness theorems.

Throughout this paper, by meromorphic functions we will always mean meromorphic functions in the complex plane. We assume that the reader is familiar with the basic notions and results in the Nevanlinna theory, which can be found, for example, in [4, 9, 18, 19]. It will be convenient to let E(0,+) denote any set of positive real numbers of finite linear measure, not necessarily the same at each occurrence. For a nonconstant meromorphic function h, we denote by T(r,h) the Nevanlinna characteristic function of h and by S(r,h) any quantity satisfying S(r,h)=o(T(r,h)), as rE and r runs to infinity. Let k be a positive integer, and let a be a complex value in the extended complex plane. Next we denote by N(k(r,1/(ha)) the counting function of those a-points of the nonconstant meromorphic function h in |z|<r, where each point in N(k(r,1/(ha)) is counted according to its multiplicity, and each point in N(k(r,1/(ha)) is of multiplicity k. Here N(k(r,1/(h)) means N(k(r,h).

Let f and g be two nonconstant meromorphic functions, and let a be a value in the extended plane. We say that f and g share the value a CM, provided that f and g have the same a-points in the complex plane, and each common a-point of f and g has the same multiplicities related to f and g. We say that f and g share the value a IM, provided that f and g have the same a-points in the complex plane (cf.[18]). In terms of sharing values, two nonconstant meromorphic functions in the complex plane must be identically equal if they share five values IM, and one must be a Möbius transformation of the other one if they share four values CM; the numbers "five" and "four" are the best possible, as shown by Nevanlinna (cf. [15, 18]). L-functions, with the Riemann zeta function as a prototype, are important objects in number theory, and value distribution of L-functions has been studied extensively. See, for example, Ki[7], Li[10, 11, 12] Hu-Li[6] and Steuding [17].

This paper concerns the question of how an L-function is uniquely determined in terms of the pre-images of complex values in the extended complex plane, or sharing values. We refer the reader to the monograph [17] for a detailed discussion on the topic and related works. Throughout the paper, an L-function always means an L-function L in the Selberg class, which includes the Riemann zeta function ζ(s)= n=1ns and essentially those Dirichlet series where one might expect a Riemann hypothesis. Such an L-function is defined to be a Dirichlet series L(s)= n=1a(n)ns satisfying the following axioms (cf.[16, 17]):

(i) Ramanujan hypothesis. a(n)nε for every ε>0.

(ii) Analytic continuation. There is a nonnegative integer k such that (s1)kL(s) is an entire function of finite order.

(iii) Functional equation. L satisfies a functional equation of type

ΛL(s)=ωΛL(1s¯)¯,

where

ΛL(s)=L(s)Qs j=1KΓ(λjs+νj)

with positive real numbers Q, λj and complex numbers

νj, ω with Reνj0 and |ω|=1.

(iv) Euler product hypothesis. L(s)=pexp k=1b( pk) pks with suitable coefficients b(pk) satisfying b(pk)pkθ for some θ<1/2, where the product is taken over all prime numbers p.

We first recall the following result due to Steuding [17], which actually holds without the Euler product hypothesis:

Theorem A. ([17, p.152]) If two L-functions L1 and L2 with a(1)=1 share a complex value c CM, then L1=L2.

Remark 1.1. In 2016, Hu-Li [6] pointed out that Theorem A is false when c=1. A counter example was given by Hu-Li [6, Remark 4] as follows: Let L1(s)=1+24s and L2(s)=1+39s. Then L1 and L2 trivially satisfy axioms (i) and (ii). Also, one can check that L1 satisfies the functional equation

2sL(s)=21sL(1s¯)¯,

and L2 satisfies the functional equation

3sL(s)=31sL(1s¯)¯.

Thus, L1 and L2 also satisfy axiom (iii). It is clear that L1 - 1 and L2 - 1 do not have any zeros and thus satisfy the assumptions of Theorem A with c = 1, but L1L2.

Theorem A implies that two L-functions with a(1)=1 must be identically equal if they have the same zeros with counting multiplicities. Two L-functions with "enough" common zeros without counting multiplicities are expected to be dependent in a certain sense (cf.[1]). Since L-functions are analytically continued as meromorphic functions in the complex plane, in order to study how an L-function is uniquely determined by pre-images of complex values in the extended plane, one should examine the situation involving an arbitrary L-function and an arbitrary meromorphic function. The first observation on this uniqueness question is that the above theorem no longer holds for an L-function and a meromorphic function. For instance, the function ζ and ζeg, where g is any entire function, share 0 CM, but they are not identically equal. It is natural to consider two sharing values, i.e., whether two sharing values with counting multiplicities would force an L-function and a meromorphic function to be identically equal. This turns out not to be the case either. For instance, consider the function f=2ζζ+1. It is then clear that ζ and f share 0, 1 CM, but they are not identically equal. Observe, however, when considering L-functions, these functions have only one possible pole at s = 1, which is implicit in the conditions of the above theorem. Thus, this leads us to consider the natural objects of those meromorphic functions with finitely many poles (cf.[10]). The following uniqueness theorem was then established by Li [10] in 2009:

Theorem B.([10]) Let f be a meromorphic function in the complex plane such that f has finitely many poles, and let a and b be two distinct finite values. If f and a nonconstant L-function L share a CM and b IM, then f=L.

Remark 1.2. The number "two" in Theorem B is the best possible, as shown by the above example with L=ζ and f=ζeg.

By Theorem B we can get the following result:

Corollary A.([10]) Let f be a meromorphic function in the complex plane, and let a, b, c be three distinct values in the extended complex plane such that a and b= or c=. If f shares a CM and b, c IM with a nonconstant L-function L, then f=L.

In a communication to Professor Li, Yang asked the following question:

Question A.([10]) If f is a meromorphic function in that shares three distinct values a, b CM and c IM with the Riemann zeta function ζ, where c{a,b,0,}, is f equal to ζ ?

Remark 1.3. By taking L=ζ in Corollary A, we can find that the conclusion of Corollary A holds, which gives a positive answer to Question A provided that any one of a, b, c is ∞ in Question A.

Next we consider the first, the second and the fourth Painlevé equations given respectively by

(PI)ω=z+6ω2, (PII)ω=2ω3+zω+αwithα, (PIV)2ωω=(ω)2+3ω4+8zω3+4(z2α)ω2+βwithα,β.

In 2007, Lin-Tohge [13] obtained some results similar to Theorem B. Indeed, Lin-Tohge [13] studied some shared-value properties of the first, the second and the fourth Painlevé transcendents by applying their distinctive value distribution, and proved the following results:

Theorem C.([13, Theorem 1]) Let ω be an arbitrary nonconstant solution of one of the equations (PI), (PII) and (PIV), and let f be a nonconstant meromorphic function that shares four distinct values a1, a2, a3, a4 IM with ω, where a1, a2, a3, a4 are four distinct values in the extended complex plane, then f=ω.

Theorem D.([13, Theorem 2]) Let ω be an arbitrary solution of (PI) and f be a meromorphic function. Assume that f and ω share two distinct values a1 and a2 CM, where a1 and a2 are two distinct values in the extended complex plane, then we have

limrN0r,1fωT(r,ω)=0orlimrN0r,1fωT(r,ω)=.

Theorem F.([13, Theorem 3]) Let ω be an arbitrary solution of (PI). Then there does not exist a pair of two finite values a, b such that Eω{a}E ω{b}. Here Eω({a}) denotes the set of a-points of ω in the complex plane, where each a-point of ω with multiplicity m is counted m times in the set Eω({a}). While Eω({b}) denotes the set of b-points of ω' in the complex plane, where each b-point of ω' with multiplicity m is counted m times in the set Eω({b}).

Regarding Theorem B, one may ask, what can be said about the conclusion of Theorem B if we remove the assumption "f has finitely many poles in the complex plane" in Theorem B. In this direction, we will prove the following result that is an extension to Theorem B:

Theorem 1.1. Let f be a meromorphic function in the complex plane, and let a, b, and c be three distinct finite values. If f and a nonconstant L-function L share a, b CM and c IM, then f=L.}

By Theorem 1.1 we get the following result:

Corollary 1.2. If f is a meromorphic function in that shares three distinct values a, b CM and c IM with the Riemann zeta function ζ, where c{a,b,}, then f=ζ.

As a special case of Corollary 1.2, we give the following result which completely resolves Question A:

Corollary 1.3. If f is a meromorphic function in that shares three distinct values a, b CM and c IM with the Riemann zeta function ζ, where c{a,b,0,}, then f=ζ.

In the same manner as in the proof of Theorem 1.1, we can get the following result by Lemma 2.10 in Section 2 of the present paper:

Theorem 1.4. Let f be a meromorphic function in the complex plane, let L be a nonconstant L-function, and let a, b, and c be three distinct finite values in the complex plane. Suppose that f(k)} and L(k)} share a, b CM and c IM, where k≥ 1 is a positive integer. Then f(k)=L(k).

Throughout this paper, we will apply Nevanlinna theory to prove the main result in this paper.

In this section, we will give some important lemmas to prove the main result of the present paper. For convenience in stating the following first result from Gundersen [3], we shall use the following notation: we shall let (f, H) denote a pair that consists of a transcendental meromorphic function f and a finite set

H={(k1,j1),(k2,j2),,(kq,jq)}

of distinct pairs of integers that satisfy ki>ji0 for 1iq.

Lemma 2.1.([3, Corollary 2]) Let (f, H) be a given pair where f has finite order ρ, and let ε>0 be a given constant. Then there exists a set E(1,) that has finite logarithmic measure, such that for all s satisfying |s|E[0,1] and for all (k,j)H, we have

f(k)(s)f(j)(s)|s|(kj)(ρ1+ε).

The following result is due to Mokhon-ko [14]:

Lemma 2.2.(Valiron-Mokhon-ko lemma, [14]) Let f be a nonconstant meromorphic function, and let F= k=0pakfk/ j=0qbjfj be an irreducible rational function in f with constant coefficients {ak} and {bj}, where ap≠ 0 and bq0. Then T(r,F)=dT(r,f)+O(1), where d=max{p,q}.

We also need the following result due to Lahiri-Sarkar [8]:

Lemma 2.3.([8, Lemma 6]) Let F and G be two distinct nonconstant meromorphic functions that share 0, 1, ∞ IM. If F is a Möbius transformation of G, then F and G satisfy one of the following six relations: (i) FG= 1; (ii) (F-1)(G-1)= 1; (iii) F+G= 1; (iv) F= cG; (v) F1=c(G1); (vi) ((c1)F+1)((c1)Gc)=c. Here c0,1 is a complex number.

The following result is from Gundersen [2]:

Lemma 2.4.([2, Theorem 3]) Suppose that f and g are two nonconstant meromorphic functions that share 0, 1, ∞ IM. Then

13+o(1)T(r,g)<T(r,f)<(3+o(1))T(r,g),

as r and rE, where E(0,+) is a subset of finite linear measure.

Lemma 2.5.([20, proof of Lemma 4]) Let f and g be two distinct nonconstant meromorphic functions that share 0, 1 CM and ∞ IM. If N¯(r,f)=S(r,f), then N(2r,1f+N(2r,1f1=S(r,f).

Lemma 2.6.([21, Lemma 6]) Let f1 and f2 be two nonconstant meromorphic functions such that

N¯(r,fj)+N¯r,1fj=S(r)

for 1j2. Then either N¯0(r,1;f1,f2)=S(r) or there exist two integers p and q, (|p|+|q|>0) such that f1pf2q=1, where N¯0(r,1;f1,f2) denotes the reduced counting function of the common 1-points of f1 and f2, T(r)=T(r,f1)+T(r,f2) and S(r)=o(T(r)), as r " ∞ and rE, E(0,+) is a subset of r of finite linear measure.

For introducing the following result, we first give the following notation (cf.[20]): Let F and G be two distinct nonconstant meromorphic functions sharing 0, 1 and ∞ IM. Next we use N0(r) to denote the counting function of those zeros of f-g that are not zeros of F, F-1 and 1/F, where each point in N0(r) is counted according to its multiplicity. We denote by N¯0(r) the reduced form of N0(r).

The following lemma is essentially due to Zhang [21]:

Lemma 2.7.([21, proof of Theorem 1 and Theorem 2]) Let F and G be two distinct nonconstant meromorphic functions sharing 0, 1 and ∞ CM, and let N0(r)S(r,f). If F is a Möbius transformation of G, then

N0(r)=T(r,F)+S(r,F).

If F is not any Möbius transformation of G, then

N0(r)12T(r,F)+S(r,F),

and F and G assume one of the following relations:

(i) F=e(k+1)γ1esγ1,G=e(k+1)γ1esγ1;

(ii) F=esγ1e(k+1)γ1,G=esγ1e(k+1)γ1;

(iii) F=esγ1e(k+1s)γ1,G=esγ1e(k+1s)γ1.

Here γ is a nonconstant entire function, s and k ≥ 2 are positive integers such that s and k+1 are relatively prime and 1sk.

Lemma 2.8. ([22]) Let s(>0) and t are relatively prime integers, and let c be a finite complex number such that cs=1, then there exists one and only one common zero of ωs1 and ωtc.

Finally we prove the following result which plays an important role in proving the main results of this paper:

Lemma 2.9. Let F and G be two distinct nonconstant meromorphic functions sharing 0, 1 CM and ∞ IM. Suppose that F is not a Möbius transformation of G. If N¯(r,F)=S(r,F), then

(i) N0r,1F= N¯0r,1F+S(r,F),N¯r,1F= N¯0r,1F+S(r,F), the same identities hold for G.

(ii) T(r,F)=N¯r,1G+N0(r)+S(r,F),T(r,G)=N¯r,1F+N0(r)+S(r,F),N0(r)= N ¯ 0(r)+S(r,F).

Here N0(r,1F)( N¯0(r,1F)) denotes the counting function corresponding to the zeros of F' that are not zeros of F and F-1, (ignoring multiplicities) and N0(r)( N¯0(r)) is the counting function of the zeros of F-G that are not zeros of G, G-1 and 1G, (ignoring multiplicities ).

Proof. First of all, we set

F1G1=h1, FG=h2

and

h0=h1h2.

By Lemma 2.4 we have

S(r,F)=S(r,G).

By (2.4), Lemma 2.5 and the assumption of Lemma 2.9 we have

N(2r,1F+N(2r,1F1+N¯(r,F)=S(r,F)

and

N(2r,1G+N(2r,1G1+N¯(r,G)=S(r,F).

By (2.1)-(2.3), (2.5), (2.6) and the assumption that F and G share 0, 1 CM and ∞ IM we get

N¯r,hj+N¯r,1hj=S(r,F)with0j2.

By (2.1)-(2.3) and the assumption that F is not a Möbius transformation .of G we can see that none of h1, h2 and h0 are constants. Therefore h11, h21 and h01. This together with (2.1)-(2.3) gives

F=h11h01

and

G=h111h011.

Set

h=h1h1h0h0=h1h1h1h1h2h2.

Then from (2.1), (2.2), (2.3) and (2.10) we can deduce

T(r,h)=S(r,F).

If

h1h1(h1)h=0,

then

h1=c1(h1),

where c10 is a finite complex number. By (2.11) and (2.12) we deduce

T(r,h1)=S(r,F).

Again from (2.10) and (2.12) we have

h0h0=c1h1 h1h1+c1=(c1 h11+1)c1h11+1.

By integrating two sides of (2.14) we can get

h0=c2c1h11+1,

where c20 is a finite complex number. By (2.13) and (2.15) we have

T(r,h0)=T(r,h1)+O(1)=S(r,F).

By (2.8), (2.13) and (2.16) we can get T(r,F)=S(r, F), this is impossible. Thus

h1h1(h1)h0,

which together with (2.8) gives

Fh=h1h0h+h1h01.

Set

H=(Fh)(h01)=h1h0h+h1.

By (2.10) and (2.18) we get

HHh1h1= ( h1 h0h+h1) h1 h1(h1h0h+h1)(Fh)(h01)=h1 h1(h1)h Fh,

and so we have

1Fh=HHh1h1h1h1(h1)h.

By (2.3), (2.11) and (2.19) we deduce

mr,1Fh=S(r,F)

and

N(2(r,1Fh)=S(r,F).

By (2.1), (2.3) and (2.9) we have

FGG1=h11andG=h11h1h2.

Thus

G(FG)G(G1)=( h 2h2 h 1h1)h1+ h 1h1h0 h 2h2h01.

On the other hand, by (2.10) and (2.19) we have

(Fh)(h2h2h1h1)=(h2 h2h1 h1)h1+h1 h1h0h2 h2h01.

By (2.23) and (2.24) we have

h0h0(Fh)=G(FG)G(G1).

By (2.3), (2.5), (2.21), (2.22) and (2.25) we easily deduce

Nr,1Fh=N0(r)+N0r,1G+S(r,F), N0(r)= N¯0(r)+S(r,F)

and

N0r,1G= N¯0r,1G+S(r,F).

By (2.28) and Lemma 2.5 we deduce

N0r,1G=N¯r,1G+S(r,F).

By (2.11), (2.20) and (2.26) we deduce

T(r,F)=N0(r)+N0r,1G+S(r,F).

In the same manner as above we get

N0r,1F= N¯0r,1F+S(r,F),N0r,1F=N¯r,1F+S(r,F)

and

T(r,G)=N0(r)+N0r,1F+S(r,F).

By (2.28), (2.29) and (2.31) we get the conclusion (i) of Lemma 2.9. By (2.27), (2.30) and (2.32) we get (ii) of Lemma 2.9. This completely proves Lemma 2.9.

Lemma 2.10.([5]) Let f be a transcendental meromorphic function in C. Then, for each K > 1, there exists a set M(K) of the lower logarithmic density at most d(K)=1(2eK11)1>0, that is

logdens¯M(K)=liminfr1logrM(K)[1,r] dtt d(K),

such that, for every positive integer k,

limsuprM(K)rT(r,f)T(r,f(k))3eK.

First of all, we denote by d the degree of L. Then d=2 j=1Kλj>0 (cf.[17, p.113]), where K and λj are respectively the positive integer and the positive real number in the functional equation of the axiom (iii) of the definition of L-function. Therefore, by Steuding [17, p.150] we have

T(r,L)=dπrlogr+O(r),

which together with the definition of the order of a meromorphic function implies that

ρ(L)=1.

By noting that L has only one possible pole at s=1, we have

N(r,L)logr+O(1),asr.

On the other hand, by the assumption that f and L share a, b CM and c IM, we have by the second fundamental theorem that

T(r,f)N¯r,1fa+N¯r,1fb+N¯r,1fc+O(logr+logT(r,f))=N¯r,1La+N¯r,1Lb+N¯r,1Lc+O(logr+logT(r,f))3T(r,L)+O(logr+logT(r,f)),

i.e.,

T(r,f)3T(r,L)+O(logr+logT(r,f)),

as r possibly outside of an exceptional set of finite linear measure. Similarly

T(r,L)3T(r,f)+O(logr+logT(r,L)),

as r possibly outside of an exceptional set of finite linear measure.

By (3.4), (3.5), the definition of the order of a meromorphic function and the standard reasoning of removing an exceptional set we deduce

ρ(f)=ρ(L)=1.

Now we set

ψ=f(fa)(fb)L(La)(Lb)=1abffaffbLLa+LLb.

By the assumption that f and L share a, b CM, we can deduce from (3.7) that ψ is an entire function. Therefore, by (3.6) and Lemma 2.1 we have

|ψ(z)|O(|z|ε)

for all z satisfying |z|E[0,1]. Here E(1,) is some subset that has finite logarithmic measure.

By (3.8) we can see that ψ is reduced to a constant, say ψ =c1. This together with (3.7) gives

(f(z)a)(L(z)b)(f(z)b)(L(z)a)=A1ec1(ab)z,

where A10 is a constant. By noting that

Tr,A1ec1(ab)z=|c1(ab)|rπ(1+o(1))+O(1),as|z|=r,

we deduce by (3.1), (3.5) and (3.10) that

Tr,A1ec1(ab)z=o(T(r,L))andTr,A1ec1(ab)z=o(T(r,f)),

as r.

By (3.9) we consider the following two cases:

Case 1. Suppose that there exists a subset I(0,+) with infinite linear measure such that

limrIrN¯r,1fcr=+.

Next we prove

A1ec1(ab)z1.

Indeed, if

A1ec1(ab)z1,

by (3.9), (3.11), (3.14) and the assumption that f and L share c IM we have

N¯r,1fcN¯r,1A1ec1(ab)z1=Tr,A1ec1(ab)z+O(1)=|c1(ab)|rπ(1+o(1))+O(1),

which contradicts (3.12), and so (3.13) is valid. By (3.9) and (3.13) we get the conclusion of Theorem 1.1.

Case 2. Suppose that at most there exists a subset E(0,+) with finite linear measure such that

limrErN¯r,1fcr<+.

Then, by (3.15) we have

N¯r,1fcA2r,

as r and rE, where A2>0 is a constant. Now we set

F=fafcbcba,G=LaLcbcba.

Noting the assumption that f and L share a, b CM, and c IM, we have by (3.17) that F and G share 0, 1 CM, and ∞ IM. Moreover, by (2.4), (3.1), (3.4), (3.5), (3.6), (3.16), (3.17) and Lemma 2.2 we deduce

N¯(r,F)=S(r,F),N¯(r,G)=S(r,F).

We discuss the following two subcases:

Subcase 2.1. Suppose that F is a Möbius transformation of G. Then, by Lemma 2.3 we can see that F and G satisfy one of the six relations (i)-(vi) of Lemma 2.3. We consider the following two subcases:

Subcase 2.1.1. Suppose that F and G satisfy one of (i), (ii) and (vi) of Lemma 2.3. We discuss this as follows:

Suppose that F and G satisfy (i) of Lemma 2.3. Then, 0 and ∞ are exceptional values of F and G. Therefore,

F=eα,G=eα,

where α is a nonconstant entire function. By the right formulae of (3.17) and (3.19) we have

LaLcbcba=eα.

By (3.20) and Lemma 2.2 we have

T(r,L)=T(r,eα)+O(1).

By (3.6), (3.21) and the definition of the order of a meromorphic function we have ρ(eα)=1, which implies that α is a polynomial of degree degα=1, say α(z)=A3z+B1, where A30 and B1 are complex constants. This implies that (3.21) can be rewritten as

T(r,L)=T(r,eα)+O(1)=|A3|rπ(1+o(1))+O(1),

which contradicts (3.1).

uppose that F and G satisfy one of (ii) and (vi) of Lemma 2.3. Then, in the same manner as the above discussion we can get a contradiction.

Subcase 2.1.2. Suppose that F and G satisfy one of (iii), (iv) and (v) of Lemma 2.3, say F and G satisfy (iii) of Lemma 2.3. Then, 0 and 1 are Picard exceptional values of F and G. Therefore

GG1=eβ,

where β is an entire function. By (3.17), (3.22) and Lemma 2.2 we have

T(r,L)=T(r,eβ)+O(1).

By (3.6), (3.23) and the definition of the order of a meromorphic function we have ρ(eβ)=1, which implies that β is a polynomial of degree degβ=1, say β(z)=A4z+B2, where A4≠ 0 and B2 are complex constants. This implies that (3.23) can be rewritten as

T(r,L)=T(r,eβ)+O(1)=|A4|rπ(1+o(1))+O(1),

which contradicts (3.1).

Suppose that F and G satisfy one of (iv) and (v) of Lemma 2.3. Then, in the same manner as the above discussion we can get a contradiction.

Subcase 2.2. Suppose that F is not a Möbius transformation of G. In the same manner as in the proof of Lemma 2.9 we have (2.1)-(2.4) and (2.7)-(2.9). By (2.8) and (2.9) we have

FG=(h11)(1h21)h01.

By (2.1)-(2.4), (2.7)-(2.9) and (3.24) we deduce

N0(r)= N¯0(r,1;h1,h0)+S(r,f)= N¯0(r,1;h1,h2)+S(r,F).

We consider the following two subcases:

Subcase 2.2.1. Suppose that

N0(r)S(r,F).

Then, by (3.25) and (3.26) we have

N¯0(r,1;h1,h2)S(r,F).

By (2.7), (3.27) and Lemma 2.6 we know that there exist two integers s and t (|s|+|t|>0), such that

h1sh2t=1.

By substituting (2.1) and (2.2) into (3.28) we get

Ft(F1)s=Gt(G1)s.

By noting that F is not a Möbius transformation of G, we deduce by (3.29) that s, t are nonzero integers such that |s||t|, this together with the assumption that F and G share ∞ IM implies that F and G share ∞ CM. Combining this with Lemma 2.7 and the assumption that F and G share 0, 1 CM, we can see that F and G satisfy one of the three relations (i)-(iii) of Lemma 2.7, say F and G satisfy (i). Then

F=e(k+1)γ1esγ1andG=e(k+1)γ1esγ1,

where γ is a nonconstant entire function, k ≥ 2 and s are positive integers such that s and k+1 are relatively prime and 1sk. Moreover, N0(r) is such that N0(r)12T(r,F)+S(r,F).

By (3.17), (3.30), Lemma 2.2 and Lemma 2.8 we have

T(r,f)=T(r,F)+O(1)=T(r,G)+O(1)=T(r.L)+O(1)=kT(r,eγ)+O(1).

By (3.6) and (3.31) we have ρ(eγ)=1, this implies that γ is a polynomial of degree degγ=1, say γ=A5z+B3, where A50 and B3 are complex constants. This implies that (3.23) can be rewritten as

T(r,L)=kT(r,eγ)+O(1)=k|A5|rπ(1+o(1))+O(1),

which contradicts (3.1).

Subcase 2.2.2. Suppose that

N0(r)=S(r,F).

By noting that L has a pole z=1 in the complex plane at most, we have by (3.3) and (3.17) that

Nr,1Ga1logr+O(1),

where a1=bcba{0,1,}. Therefore, by (2.4), (3.18), (3.32), (3.33), the conclusion (ii) of Lemma 2.9 and the second fundamental theorem we have

2T(r,G)N¯(r,G)+Nr,1G+Nr,1G1+Nr,1Ga1Nr,1 G+O(logr)=N¯(r,G)+Nr,1G+Nr,1G1+Nr,1Ga1+N0(r)T(r,F)+O(logr)Nr,1G+Nr,1G1T(r,F)+O(logr)+S(r,F)2T(r,G)T(r,F)+O(logr)+S(r,F),

i.e.,

T(r,F)=O(logr)+S(r,F).

By (3.34) we deduce that F is a rational function. This together with (3.6) and the left equality of (3.17) and Lemma 2.2 implies that T(r,f)=T(r,F)+O(1)=O(logr). Combining this with (3.5), we have T(r,L)=O(logr), which contradicts (3.1). Theorem 1.1 is thus completely proved.

The authors wish to express their thanks to the referee for his/her valuable suggestions and comments.

  1. E. Bombieri and A. Perelli, Distinct zeros of L-functions, Acta Arith., 83(3)(1998), 271-281.
    CrossRef
  2. G. G. Gundersen, Meromorphic functions that share three or four values, J. London Math. Soc., s2-20(3)(1979), 457-466.
    CrossRef
  3. G. G. Gundersen, Estimates for the logarithmic derivative of a meromorphic function, plus similar estimates, J. London Math. Soc., s2-37(1)(1988), 88-104.
    CrossRef
  4. W. K. Hayman, Meromorphic Functions, Clarendon Press, Oxford, 1964.
  5. W. K. Hayman and J. Miles, On the growth of a meromorphic function and its derivatives, Complex Var. Theory Appl., 12(1-4)(1989), 245-260.
    CrossRef
  6. P. C. Hu and B. Q. Li, A simple proof and strengthening of a uniqueness theorem for L-functions, Canadian math. Bull., 59(1)(2016), 119-122.
    CrossRef
  7. H. Ki, A remark on the uniqueness of the Dirichlet series with a Riemann-type function equation, Adv. Math. 231(5)(2012), 2484-2490.
    CrossRef
  8. I. Lahiri and A. Sarkar, On uniqueness theorem of Tohge, Arch. Math., 84(5)(2005), 461-469.
    CrossRef
  9. I. Laine, Nevanlinna Theory and Complex Differential Equations, Walter de Gruyter, Berlin, 1993.
    CrossRef
  10. B. Q. Li, A result on value distribution of L-functions, Proc. Amer. Math. Soc., 138(6)(2009), 2071-2077.
    CrossRef
  11. B. Q. Li , A uniqueness theorem for Dirichlet series satisfying a Riemann type functional equation, Adv. Math., 226(5)(2011), 4198-4211.
    CrossRef
  12. B. Q. Li, On common zeros of L-functions, Math. Z., 272(3-4)(2012), 1097-1102.
    CrossRef
  13. W.C. Lin and K. Tohge, On shared-value properties of Painlev´e Transcendents, Comput. methods Funct.theory, 7(2)(2007), 477-499.
    CrossRef
  14. A. Z. Mokhon-ko, On the Nevanlinna characteristics of some meromorphic functions, in: Theory of Functions, Functional Analysis and Their Applications, vol.14, Izd-vo Khar’kovsk. Un-ta, 1971, pp. 83-87.
  15. R. Nevanlinna, Einige Eindeutigkeitss¨atze in der Theorie der meromorphen Funktionen, Acta Math., 48(3)(1926), 367-391.
    CrossRef
  16. A. Selberg, Old and new conjectures and results about a class of Dirichlet series, Proceedings of the Amalfi Conference on Analytic Number Theory (Maiori, 1989), E. Bombieri et al. (eds.), Collected papers, Vol. II, Springer-Verlag, 1991, pp. 47-63.
  17. J. Steuding, Value distribution of L-functions, Lecture Notes in Mathematics, 1877, Springer-Verlag, Berlin, 2007.
  18. C. C. Yang and H. X. Yi, Uniqueness Theory of Meromorphic Functions, Kluwer Academic Publishers, Dordrecht/Boston/London, 2003.
  19. L. Yang, Value Distribution Theory, Springer-Verlag, Berlin Heidelberg, 1993.
  20. H. X. Yi, Unicity theorems for meromorphic functions that share three values, Kodai Math. J., 18(2)(1995), 300-314.
    CrossRef
  21. Q. C. Zhang, Meromorphic functions sharing three values, Indian J. Pure Appl. Math., 30(7)(1999), 667-682.
  22. Q. C. Zhang, Meromorphic functions sharing values or sets, Doctoral Thesis, Shan-dong University, 1999.