Skip to main content

A note on entire functions and their differences

Abstract

In this paper, we prove that for a transcendental entire function f(z) of finite order such that λ(fa(z))<σ(f), where a(z) is an entire function and satisfies σ(a(z))<1, n is a positive integer and if Δ η n f(z) and f(z) share the function a(z) CM, where η (C) satisfies Δ η n f(z)0, then

a(z)0andf(z)=c e c 1 z ,

where c, c 1 are two nonzero constants.

MSC:39A10, 30D35.

1 Introduction and results

In this paper, we assume that the reader is familiar with the fundamental results and the standard notations of Nevanlinna’s value distribution theory of meromorphic functions (see [13]). In addition, we use the notation λ(f) for the exponent of convergence of the sequence of zeros of a meromorphic function f, and σ(f) to denote the order growth of f. For a nonzero constant η, the forward differences Δ η n f(z) are defined (see [4, 5]) by

Δ η f ( z ) = Δ η 1 f ( z ) = f ( z + η ) f ( z ) and Δ η n + 1 f ( z ) = Δ η n f ( z + η ) Δ η n f ( z ) , n = 1 , 2 , .

Throughout this paper, we denote by S(r,f) any function satisfying S(r,f)=o(T(r,f)) as r, possibly outside a set of r of finite logarithmic measure. A meromorphic function α(z) is said to be a small function of f(z) if T(r,α(z))=S(r,f), and we denote by S(f) the set of functions which are small compared to f(z).

Let f and g be two nonconstant meromorphic functions, and let aC. We say that f and g share the value a CM (IM) provided that fa and ga have the same zeros counting multiplicities (ignoring multiplicities), that f and g share the value ∞ CM (IM) provided that f and g have the same poles counting multiplicities (ignoring multiplicities). Using the same method, we can define that f and g share the function a(z) CM (IM), where a(z)S(f)S(g). Nevanlinna’s four values theorem [6] says that if two nonconstant meromorphic functions f and g share four values CM, then fg or f is a Möbius transformation of g. The condition ‘f and g share four values CM’ has been weakened to ‘f and g share two values CM and two values IM’ by Gundersen [7, 8], as well as by Mues [9]. But whether the condition can be weakened to ‘f and g share three values IM and another value CM’ is still an open question.

In the special case, we recall a well-known conjecture by Brück [10].

Conjecture Let f be a nonconstant entire function such that hyper order σ 2 (f)< and σ 2 (f) is not a positive integer. If f and f share the finite value a CM, then

f a=c(fa),

where c is a nonzero constant.

The notation σ 2 (f) denotes hyper-order (see [11]) of f(z) which is defined by

σ 2 (f)= lim ¯ r log log T ( r , f ) log r .

The conjecture has been verified in the special cases when a=0 [10], or when f is of finite order [12], or when σ 2 (f)< 1 2 [13].

Recently, many authors [1417] started to consider sharing values of meromorphic functions with their shifts. Heittokangas et al. proved the following theorems.

Theorem A (See [15])

Let f be a meromorphic function with σ(f)<2, and let cC. If f(z) and f(z+c) share the values a (C) andCM, then

f(z+c)a=τ ( f ( z ) a )

for some constant τ.

In [15], Heittokangas et al. give the example f(z)= e z 2 +1 which shows that σ(f)<2 cannot be relaxed to σ(f)2.

Theorem B (See [16])

Let f be a meromorphic function of finite order, let cC. If f(z) and f(z+c) share three distinct periodic functions a 1 , a 2 , a 3 S ˆ (f) with period c CM (where S ˆ (f)=S(f){}), then f(z)=f(z+c) for all zC.

Recently, many results of complex difference equations have been rapidly obtained (see [1825]). In the present paper, we utilize a complex difference equation to consider uniqueness problems.

The main purpose of this paper is to utilize a complex difference equation to study problems concerning sharing values of meromorphic functions and their differences. It is well known that Δ η f(z)=f(z+η)f(z) (where η (C) is a constant satisfying f(z+η)f(z)0) is regarded as the difference counterpart of f . So, Chen and Yi [20] considered the problem that Δ η f(z) and f(z) share one value a CM and proved the following theorem.

Theorem C (See [20])

Let f be a finite order transcendental entire function which has a finite Borel exceptional value a, and let η (C) be a constant such that f(z+η)f(z). If Δ η f(z)=f(z+η)f(z) and f(z) share the value a CM, then

a=0and f ( z + η ) f ( z ) f ( z ) =A,

where A is a nonzero constant.

Question 1 What can be said if we consider the forward difference Δ η n f(z) and f(z) share one value or one small function?

In this paper, we answer Question 1 and prove the following theorem.

Theorem 1.1 Let f(z) be a finite order transcendental entire function such that λ(fa(z))<σ(f), where a(z) is an entire function and satisfies σ(a)<1. Let n be a positive integer. If Δ η n f(z) and f(z) share a(z) CM, where η (C) satisfies Δ η n f(z)0, then

a(z)0andf(z)=c e c 1 z ,

where c, c 1 are two nonzero constants.

In the special case, if we take a(z)a in Theorem 1.1, we can get the following corollary.

Corollary 1.1 Let f(z) be a finite order transcendental entire function which has a finite Borel exceptional value a. Let n be a positive integer. If Δ η n f(z) and f(z) share value a CM, where η (C) satisfies Δ η n f(z)0, then

a=0andf(z)=c e c 1 z ,

where c, c 1 are two nonzero constants.

Remark 1.1 From Corollary 1.1, we know that Δ η n f ( z ) f ( z ) = ( e c 1 η 1 ) n and it shows that the quotient of Δ η n f(z) and f(z) is related to η, n and c 1 , but not related to c. On the other hand, Corollary 1.1 shows that if f has a nonzero finite Borel exceptional value b , then, for any constant η satisfying Δ η n f(z)0, the value b is not shared CM by Δ η n f(z) and f(z). See the following example.

Example 1.1 Suppose that f(z)= e z + b , where b is a nonzero finite value. Then f has a nonzero finite Borel exceptional value b . For any η2kπi, kZ, the value b is not shared CM by Δ η n f(z) and f(z). Observe that Δ η n f(z)= j = 0 n ( 1 ) j C n j f(z+(nj)η), where C n j are the binomial coefficients. Thus, for any η2kπi, kZ, we have Δ η n f(z)= ( e η 1 ) n e z . Thus, we can see that f(z) b = e z has no zero, but Δ η n f(z) b = ( e η 1 ) n e z b has infinitely many zeros. Hence, the value b is not shared CM by Δ η n f(z) and f(z).

In the special case, if we take n=1 in Theorem 1.1 and n=1 in Corollary 1.1, we can obtain the following corollaries.

Corollary 1.2 Let f(z) be a finite order transcendental entire function such that λ(fa(z))<σ(f), where a(z) is an entire function and satisfies σ(a)<1. If Δ η f(z)=f(z+η)f(z) and f(z) share a(z) CM, where η (C) satisfies f(z+η)f(z), then

a(z)0andf(z)=c e c 1 z ,

where c, c 1 are two nonzero constants.

Corollary 1.3 Let f(z) be a finite order transcendental entire function which has a finite Borel exceptional value a. If Δ η f(z)=f(z+η)f(z) and f(z) share value a CM, where η (C) satisfies f(z+η)f(z), then

a=0andf(z)=c e c 1 z ,

where c, c 1 are two nonzero constants.

Remark 1.2 The Corollary 1.2 shows that if a nonzero polynomial a(z) satisfies λ(fa)<σ(f), then a(z) is not shared CM by Δf(z) and f(z). For example, if we take a(z)z, and λ(fz)<σ(f) holds, then Δf(z) and f(z) do not have any common fixed point (counting multiplicities). See the following example.

Example 1.2 Suppose that f(z)= e z +z. Then f(z) satisfies λ(f(z)z)=0<1=σ(f) and has no fixed point. But for any η2kπi, kZ, the function Δ η f(z)=f(z+η)f(z)=( e η 1) e z +η has infinitely many fixed points by Milloux’s theorem (see [1, 3]). Hence, the nonzero polynomial a(z)z is not shared CM by Δ η f(z) and f(z).

Remark 1.3 From Corollary 1.3, we can see that under the hypothesis of Theorem C, we can get the expression of f(z), that is, f(z)=c e c 1 z . Thus, we know that the constant A in Theorem C is related to η and c 1 , but not related to c. Actually, from the proof of Lemma 2.9, we have A= e c 1 η 1 (obviously, we can obtain A1). Hence, Corollary 1.3 contains and improves Theorem C. Obviously, Theorem 1.1 generalizes Theorem C.

2 Lemmas for the proof of theorems

Lemma 2.1 (See [21])

Let f be a meromorphic function with a finite order σ, η be a nonzero constant. Let ε>0 be given, then there exists a subset E(1,) with finite logarithmic measure such that for all z satisfying |z|=rE[0,1], we have

exp { r σ 1 + ε } | f ( z + η ) f ( z ) |exp { r σ 1 + ε } .

Lemma 2.2 (See [11, 26])

Suppose that n2 and let f 1 (z),, f n (z) be meromorphic functions and g 1 (z),, g n (z) be entire functions such that

  1. (i)

    j = 1 n f j (z)exp{ g j (z)}=0;

  2. (ii)

    when 1j<kn, g j (z) g k (z) is not constant;

  3. (iii)

    when 1jn, 1h<kn,

    T(r, f j )=o { T ( r , exp { g h g k } ) } (r,rE),

where E(1,) has finite linear measure or logarithmic measure.

Then f j (z)0, j=1,,n.

ε-set Following Hayman [1], we define an ε-set to be a countable union of open discs not containing the origin and subtending angles at the origin whose sum is finite. If E is an ε-set, then the set of r1, for which the circle S(0,r) meets E, has finite logarithmic measure, and for almost all real θ, the intersection of E with the ray argz=θ is bounded.

Lemma 2.3 (See [4])

Let f be a function transcendental and meromorphic in the plane of order <1. Let h>0. Then there exists an ε-set E such that

f(z+c)f(z)=c f (z) ( 1 + o ( 1 ) ) as z in CE,

uniformly in c for |c|h.

Lemma 2.4 (See [25])

Let f be a transcendental meromorphic solution of finite order ρ of a difference equation of the form

U(z,f)P(z,f)=Q(z,f),

where U(z,f), P(z,f), Q(z,f) are difference polynomials such that the total degree degU(z,f)=n in f(z) and its shifts, and degQ(z,f)n. Moreover, we assume that U(z,f) contains just one term of maximal total degree in f(z) and its shifts. Then, for each ε>0,

m ( r , P ( z , f ) ) =O ( r ρ 1 + ε ) +S(r,f),

possibly outside of an exceptional set of finite logarithmic measure.

Remark 2.1 From the proof of Lemma 2.4 in [25], we can see that if the coefficients of U(z,f), P(z,f), Q(z,f), namely α λ (z), satisfy m(r, α λ )=S(r,f), then the same conclusion still holds.

Lemma 2.5 (See [27])

Let P n (z),, P 0 (z) be polynomials such that P n P 0 0 and satisfy

P n (z)++ P 0 (z)0.
(2.1)

Then every finite order transcendental meromorphic solution f(z) (0) of the equation

P n (z)f(z+n)+ P n 1 (z)f(z+n1)++ P 0 (z)f(z)=0
(2.2)

satisfies σ(f)1, and f(z) assumes every nonzero value aC infinitely often and λ(fa)=σ(f).

Remark 2.2 If equation (2.2) satisfies condition (2.1) and all P j (z) are constants, we can easily see that equation (2.2) does not possess any nonzero polynomial solution.

Lemma 2.6 (See [27])

Let F(z), P n (z),, P 0 (z) be polynomials such that F P n P 0 0. Then every finite order transcendental meromorphic solution f(z) (0) of the equation

P n (z)f(z+n)+ P n 1 (z)f(z+n1)++ P 0 (z)f(z)=F
(2.3)

satisfies λ(f)=σ(f)1.

Remark 2.3 From the proof of Lemma 2.5 in [27], we can see that if we replace f(z+j) by f(z+jη) (j=1,,n) in equation (2.2) or (2.3), then the corresponding conclusion still holds.

Lemma 2.7 (See [4])

Let f be a function transcendental and meromorphic in the plane which satisfies lim ̲ r T ( r , f ) r =0. Then g(z)=f(z+1)f(z) and G(z)= f ( z + 1 ) f ( z ) f ( z ) are both transcendental.

Remark 2.4 From the proof of Lemma 2.7 in [4], we can see that, under the same hypotheses of Lemma 2.7, we can obtain the following conclusion: Δ η f(z)=f(z+η)f(z) and G(z)= Δ η f ( z ) f ( z ) = f ( z + η ) f ( z ) f ( z ) are both transcendental.

Lemma 2.8 Let f(z)=H(z) e c 1 z , where H(z) (0) is an entire function such that σ(H)<1 and c 1 is a nonzero constant. If Δ η n f(z)0 for some constant η, and

Δ η n f ( z ) f ( z ) =A
(2.4)

holds, where A is a constant, then H(z) is a constant.

Proof From Δ η n f(z)0, we can see that A0. In order to prove that H(z) is a constant, we only need to prove H (z)0. Substituting f(z)=H(z) e c 1 z into (2.4), we can obtain

j = 0 n 1 ( 1 ) j C n j e ( n j ) c 1 η H ( z + ( n j ) η ) + ( ( 1 ) n A ) H(z)=0.
(2.5)

First, we assert that the sum of all coefficients of equation (2.5) is equal to zero, that is,

e n c 1 η C n 1 e ( n 1 ) c 1 η ++ ( 1 ) n 1 C n n 1 e c 1 η + ( ( 1 ) n A ) =0.
(2.6)

On the contrary, we suppose that

e n c 1 η C n 1 e ( n 1 ) c 1 η ++ ( 1 ) n 1 C n n 1 e c 1 η + ( ( 1 ) n A ) 0.

Thus, applying Lemma 2.5 and Remarks 2.2-2.3 to (2.5), we have σ(H)1, a contradiction. Hence, (2.6) holds. Thus, by (2.6) and (2.5), we have

j = 0 n 1 ( 1 ) j C n j e ( n j ) c 1 η ( H ( z + ( n j ) η ) H ( z ) ) =0.
(2.7)

By Lemma 2.3, we see that there exists an ε-set E such that for j=1,2,,n,

H(z+jη)H(z)=jη H (z) ( 1 + o ( 1 ) ) as z in CE.
(2.8)

Substituting (2.8) into (2.7), we can get

η H (z)K+η H (z)Ko(1)=0as z in CE,
(2.9)

where K is a constant and satisfies

K=n e n c 1 η C n 1 (n1) e ( n 1 ) c 1 η ++ ( 1 ) n 2 C n n 2 2 e 2 c 1 η + ( 1 ) n 1 C n n 1 e c 1 η .

Secondly, we assert that K0. If n=1, then K= e c 1 η 0; if n2, on the contrary, we suppose that K=0. Then, for j=0,1,,n1, noting that

C n j (nj)= n ! ( n j ) ( n j ) ! j ! = ( n 1 ) ! n ( n 1 j ) ! j ! =n C n 1 j ,

we have

j = 0 n 1 ( 1 ) j C n j (nj) e ( n j ) c 1 η =n e c 1 η ( e c 1 η 1 ) n 1 =0.

Thus, we can obtain from the equality above that e c 1 η =1 since n11. By (2.6) we have A= ( e c 1 η 1 ) n =0, which contradicts A0. Hence K0 and (2.9) implies H (z)0. Thus, we can know that H(z) is a nonzero constant.

Thus, Lemma 2.8 is proved. □

Lemma 2.9 Suppose that f(z) is a finite order transcendental entire function such that λ(fa(z))<σ(f), where a(z) is an entire function and satisfies σ(a)<1. Let n be a positive integer. If Δ η n f(z)0 for some constant η (C), and

Δ η n f ( z ) a ( z ) f ( z ) a ( z ) =A
(2.10)

holds, where A is a constant, then

a(z)0,A0andf(z)=c e c 1 z ,

where c, c 1 are two nonzero constants.

Proof Since f(z) is a transcendental entire function of finite order and satisfies λ(fa(z))<σ(f), we can write f(z) in the form

f(z)=a(z)+H(z) e h ( z ) ,
(2.11)

where H (0) is an entire function, h is a polynomial with degh=k (k1), H and h satisfy

λ(H)=σ(H)=λ ( f a ( z ) ) <σ(f)=degh.
(2.12)

First, we assert that a(z)0. Substituting (2.11) into (2.10), we can get that

Δ η n f ( z ) a ( z ) f ( z ) a ( z ) = j = 0 n ( 1 ) j C n j H ( z + ( n j ) η ) e h ( z + ( n j ) η ) + b ( z ) H ( z ) e h ( z ) =A,
(2.13)

where b(z)= Δ η n a(z)a(z). Rewrite (2.13) in the form

j = 0 n 1 ( 1 ) j C n j H ( z + ( n j ) η ) e h ( z + ( n j ) η ) h ( z ) + ( ( 1 ) n A ) H(z)=b(z) e h ( z ) .
(2.14)

Suppose that b(z)0. Then, from σ(H(z+(nj)η))=σ(H(z))<degh(z)=k (j=0,1,,n1), deg(h(z+(nj)η)h(z))=k1 and σ(b(z))σ(a(z))<1k, we can see that the order of growth of the left-hand side of (2.14) is less than k, and the order of growth of the right-hand side of (2.14) is equal to k. This is a contradiction. Hence, b(z) Δ η n a(z)a(z)0. Namely,

a(z+nη) C n 1 a ( z + ( n 1 ) η ) ++ ( 1 ) n 1 C n n 1 a(z+η)+ ( ( 1 ) n 1 ) a(z)=0.
(2.15)

Suppose that a(z)0. Note that the sum of all coefficients of (2.15) does not vanish. Then we can apply Lemma 2.5 and Remarks 2.2-2.3 to (2.15) and obtain σ(a(z))1, which contradicts our hypothesis. Hence, a(z)0. Thus, (2.13) can be rewritten as

Δ η n f ( z ) f ( z ) = j = 0 n ( 1 ) j C n j H ( z + ( n j ) η ) e h ( z + ( n j ) η ) h ( z ) H ( z ) =A.
(2.16)

Secondly, we prove that A0. In fact, if A=0, we obtain from (2.16) that Δ η n f(z)0, which contradicts our hypothesis.

Thirdly, we prove that σ(f)=k=1. On the contrary, we suppose that σ(f)=k2. Thus, we will deduce a contradiction for cases A= ( 1 ) n and A ( 1 ) n , respectively.

Case 1. Suppose that A= ( 1 ) n . Thus, for a positive integer n, there are three subcases: (1) n=1; (2) n=2; (3) n3.

Subcase 1.1. Suppose that n=1. Then, by A=1, we can obtain from (2.16) that

e h ( z + η ) h ( z ) =(1+A) H ( z ) H ( z + η ) 0,

a contradiction.

Subcase 1.2. Suppose that n=2. Then, by A= ( 1 ) 2 =1 and (2.16), we have

e h ( z + 2 η ) h ( z + η ) = 2 H ( z + η ) H ( z + 2 η ) .
(2.17)

Set Q 1 (z)= 2 H ( z + 2 η ) H ( z + η ) . Then, from (2.17), we can know that Q 1 (z) is a nonconstant entire function. Set σ(H)= σ 1 . Then σ 1 <σ(f)=k. By Lemma 2.1, we see that for any given ε 1 (0<3 ε 1 <k σ 1 ), there exists a set E 1 (1,) of finite logarithmic measure such that for all z satisfying |z|=r[0,1] E 1 , we have

exp { r σ 1 1 + ε 1 } | 2 H ( z + η ) H ( z + 2 η ) |exp { r σ 1 1 + ε 1 } .
(2.18)

Since Q 1 (z) is an entire function, by (2.18), we have

T ( r , Q 1 ( z ) ) =m ( r , Q 1 ( z ) ) m ( r , 2 H ( z + η ) H ( z + 2 η ) ) +O(1) r σ 1 1 + ε 1 ,

so that σ( Q 1 (z)) σ 1 1+ ε 1 <k1. Thus, by deg(h(z+η)h(z))=k1 and σ( Q 1 )<k1, we can see that the order of growth of the left-hand side of (2.17) is equal to k1, and the order of growth of the right-hand side of (2.17) is less than k1. This is a contradiction.

Subcase 1.3. Suppose that n3. Then we can obtain from (2.16) that

j = 0 n 2 ( 1 ) j C n j H ( z + ( n j ) η ) H ( z + η ) e h ( z + ( n j ) η ) h ( z + η ) + ( 1 ) n 1 C n n 1 =0.
(2.19)

Set Q 2 (z)= e h ( z + 2 η ) h ( z + η ) . Then Q 2 (z) is a transcendental entire function since σ( Q 2 (z))=k11. For j=3,4,,n, we have

e h ( z + j η ) h ( z + η ) = Q 2 ( z + ( j 2 ) η ) Q 2 ( z + ( j 3 ) η ) Q 2 (z).

Thus, (2.19) can be rewritten as

U 2 ( z , Q 2 ( z ) ) Q 2 (z)= ( 1 ) n C n n 1 ,
(2.20)

where

U 2 ( z , Q 2 ( z ) ) = H ( z + n η ) H ( z + η ) Q 2 ( z + ( n 2 ) η ) Q 2 ( z + ( n 3 ) η ) Q 2 ( z + η ) C n 1 H ( z + ( n 1 ) η ) H ( z + η ) Q 2 ( z + ( n 3 ) η ) Q 2 ( z + ( n 4 ) η ) Q 2 ( z + η ) + + ( 1 ) n 2 C n n 2 H ( z + 2 η ) H ( z + η ) .

Noting that ( 1 ) n C n n 1 0, we can see that U 2 (z, Q 2 (z))0. Set σ(H)= σ 2 . Then σ 2 <k. Since Q 2 (z) is of regular growth and σ( Q 2 (z))=k1, for any given ε 2 (0<3 ε 2 <k σ 2 ) and all r> r 0 (>0), we have

T ( r , Q 2 ( z ) ) > r k 1 ε 2 .
(2.21)

By Lemma 2.1, we see that for ε 2 , there exists a set E 2 (1,) of finite logarithmic measure such that for all z satisfying |z|=r[0,1] E 2 , we have

exp { r σ 2 1 + ε 2 } | H ( z + j η ) H ( z + η ) |exp { r σ 2 1 + ε 2 } (j=2,3,,n).
(2.22)

Thus, from (2.21) and (2.22), we can get that for j=2,3,,n,

m ( r , H ( z + j η ) H ( z + η ) ) T ( r , Q 2 ( z ) ) r σ 2 1 + ε 2 r k 1 ε 2 0 ( r  and  r [ 0 , 1 ] E 2 ) ,

that is,

m ( r , H ( z + j η ) H ( z + η ) ) =S(r, Q 2 )(j=2,3,,n).
(2.23)

Noting that deg Q 2 U(z, Q 2 )=n21 and by Lemma 2.4 and Remark 2.1, we have

T(r, Q 2 )=m(r, Q 2 )=S(r, Q 2 ),

a contradiction.

Case 2. Suppose that A ( 1 ) n . Thus, for a positive integer n, there are two subcases: (1) n=1; (2) n2.

Subcase 2.1. Suppose that n=1. Thus, (2.16) can be rewritten as

H ( z + η ) H ( z ) = ( A ( 1 ) n ) e h ( z ) h ( z + η ) =(A+1) e h ( z ) h ( z + η ) .

Noting the A+10, we can use the same method as in the proof of Subcase 1.2 and deduce a contradiction.

Subcase 2.2. Suppose that n2. Then we can obtain from (2.16) that

j = 0 n 1 ( 1 ) j C n j H ( z + ( n j ) η ) H ( z ) e h ( z + ( n j ) η ) h ( z ) + ( 1 ) n A=0.
(2.24)

Set Q 3 (z)= e h ( z + η ) h ( z ) . Then Q 3 (z) is a transcendental entire function since σ( Q 3 (z))=k11. For j=1,2,,n, we have

e h ( z + j η ) h ( z ) = Q 3 ( z + ( j 1 ) η ) Q 3 ( z + ( j 2 ) η ) Q 3 (z).

Thus, (2.24) can be rewritten as

U 3 ( z , Q 3 ( z ) ) Q 3 (z)=A ( 1 ) n ,
(2.25)

where

U 3 ( z , Q 3 ( z ) ) = H ( z + n η ) H ( z ) Q 3 ( z + ( n 1 ) η ) Q 3 ( z + ( n 2 ) η ) Q 3 ( z + η ) C n 1 H ( z + ( n 1 ) η ) H ( z ) Q 3 ( z + ( n 2 ) η ) Q 3 ( z + ( n 3 ) η ) Q 3 ( z + η ) + + ( 1 ) n 1 C n n 1 H ( z + η ) H ( z ) .

We can see that U 3 (z, Q 3 (z))0 since A ( 1 ) n 0. Noting that deg Q 3 U 3 (z, Q 3 (z))=n11, we can use the same method as in the proof of Subcase 1.3 and deduce a contradiction.

Thus, we have proved that σ(f)=k=1. And f(z) can be written as

f(z)=H(z) e c 1 z + c 0 = H (z) e c 1 z ,
(2.26)

where c 0 , c 1 (≠0) are two constants and H (z)= e c 0 H(z) (0) is an entire function and satisfies

σ ( H ( z ) ) =λ ( H ( z ) ) =λ(f)<σ(f)=1.
(2.27)

Thus, by (2.26), (2.27), (2.16) and Lemma 2.8, we can get that H (z) is a nonzero constant, and so, f(z) can be written as

f(z)=c e c 1 z ,

where c, c 1 are two nonzero constants.

Thus, Lemma 2.9 is proved. □

Remark 2.5 From the proof of Lemma 2.9 or Remark 1.3, we can see that A1 in Lemma 2.9 when n=1 and Theorem C. Unfortunately, we cannot obtain A ( 1 ) n when n2 in Lemma 2.9. This is because we can get a contradiction from the equality e c 1 η 1=1, but we cannot obtain a contradiction from the equality ( e c 1 η 1 ) n = ( 1 ) n when n2.

3 Proof of Theorem 1.1

By the hypotheses of Theorem 1.1, we can write f(z) in the form (2.11), and (2.12) holds. Since Δ η n f(z) and f(z) share an entire function a(z) CM, then

Δ η n f ( z ) a ( z ) f ( z ) a ( z ) = j = 0 n ( 1 ) n j C n j H ( z + j η ) e h ( z + j η ) + b ( z ) H ( z ) e h ( z ) = e P ( z ) ,
(3.1)

where P(z) is a polynomial and b(z)= Δ η n a(z)a(z). Obviously, σ(b(z))σ(a(z))<1.

First step. We prove

Δ η n f ( z ) a ( z ) f ( z ) a ( z ) =A,
(3.2)

where A (≠0) is a constant. If P(z)0, then, by (3.1), we see that (3.2) holds and A=1.

Now suppose that P(z)0 and degP(z)=s. Set

h(z)= a k z k + a k 1 z k 1 ++ a 0 ,P(z)= b s z s + b s 1 z s 1 ++ b 0 ,
(3.3)

where k=σ(f)1, a k (0), a k 1 ,, a 0 , b s (0), b s 1 ,, b 0 are constants. By (3.1), we can see that 0degP=sdegh=k.

In this case, we prove that P(z) is a constant, that is, s=0. To this end, we will deduce a contradiction for the cases s=k and 1s<k, respectively.

Case 1. Suppose that 1s=k. Thus, there are two subcases: (1) a(z)0; (2) a(z)0.

Subcase 1.1. Suppose that a(z)0. First we suppose that b k a k . Then (3.1) is rewritten as

g 11 (z) e P ( z ) + g 12 e h ( z ) + g 13 e h 0 ( z ) =0,
(3.4)

where h 0 (z)0 and

g 11 (z)=H(z); g 12 (z)=b(z); g 13 (z)= j = 0 n ( 1 ) n j C n j H(z+jη) e h ( z + j η ) h ( z ) .

Since σ(H)<k, σ(b)<1k and deg(h(z+jη)h(z))=k1 (j=1,2,,n), we can see that σ( g 1 m (z))<k (m=1,2,3). On the other hand, by b k a k , we can see that deg(P(h))=deg(P h 0 )=deg(h h 0 )=k. Since e P ( h ) , e P h 0 and e h h 0 are of regular growth, and σ( g 1 m )<k (m=1,2,3), we can see that for m=1,2,3,

T(r, g 1 m )=o ( T ( r , e P ( h ) ) ) =o ( T ( r , e P h 0 ) ) =o ( T ( r , e h h 0 ) ) .
(3.5)

Thus, applying Lemma 2.2 to (3.4), by (3.5), we can obtain g 1 m (z)0 (m=1,2,3). Clearly, this is a contradiction.

Now we suppose that b k = a k . Then (3.1) is rewritten as

[ H ( z ) e P ( z ) + h ( z ) b ( z ) ] e h ( z ) = j = 0 n ( 1 ) n j C n j H(z+jη) e h ( z + j η ) h ( z ) .
(3.6)

We affirm that H(z) e P ( z ) + h ( z ) b(z)0. In fact, if H(z) e P ( z ) + h ( z ) b(z)0, then, by (3.6), we can obtain

j = 1 n ( 1 ) n j C n j H(z+jη) e h ( z + j η ) h ( z ) + ( 1 ) n H(z)0,
(3.7)

this is the special case of (2.14) when b(z)0 and A=0. Hence, using the same method as in the proof of Case 2 in the proof of Lemma 2.9, we can get that σ(f)=k=1. Hence, substituting h(z)= c 1 z+ c 0 into (3.7), we have

j = 0 n ( 1 ) j C n j e ( n j ) c 1 η H ( z + ( n j ) η ) =0.
(3.8)

On this occasion, we assert that ( e c 1 η 1 ) n =0. On the contrary, we suppose that ( e c 1 η 1 ) n 0. Then the sum of all coefficients of (3.8) is ( e η 1 ) n , which does not vanish. By Lemma 2.5 and Remarks 2.2-2.3, we have σ(H)1, a contradiction. Hence, ( e c 1 η 1 ) n =0. Thus, e c 1 η =1. Substituting it into (3.8), we have

j = 0 n ( 1 ) j C n j H ( z + ( n j ) η ) =0.
(3.9)

First, we suppose that H(z) is transcendental. Then, noting that σ(H)<1 implies lim ̲ r T ( r , H ) r =0, by Lemma 2.7 and Remark 2.4, we know that Δ η H(z)=H(z+η)H(z) is transcendental. Moreover, σ( Δ η H(z))σ(H(z))<1 implies lim ̲ r T ( r , Δ η H ) r =0. Repeating the process above n1 times, we can see that Δ η n H(z) is transcendental. That is, the left-hand side of (3.9) is a transcendental function. Hence (3.9) is impossible.

Secondly, we suppose that H(z) is a nonzero polynomial. Then, noting that b k = a k , we can see that e p ( z ) + h ( z ) is a nonzero constant. Thus, from b(z)=H(z) e p ( z ) + h ( z ) , we can know that b(z) is a nonzero polynomial. Thus, applying Lemma 2.6 to the equation Δ η n a(z)a(z)=b(z) and by Remark 2.3, we have σ(a)1, a contradiction. Hence, H(z) e P ( z ) + h ( z ) b(z)0. Thus, since deg(P+h)k1, deg(h)=k, deg(h(z+jη)h(z))=k1 (j=1,2,,n) and σ(H)<k, we see that the order of growth of the left-hand side of (3.6) is equal to k, and the order of growth of the right-hand side of (3.6) is less than k. This is a contradiction.

Subcase 1.2. Suppose that a(z)0. Then (3.1) is rewritten as

H(z) e P ( z ) = j = 0 n ( 1 ) n j C n j H(z+jη) e h ( z + j η ) h ( z ) .
(3.10)

Since H(z)0, σ(H)<k, degP=s=k and deg(h(z+jη)h(z))=k1 (j=1,2,,n), we can see that the order of growth of the left-hand side of (3.10) is equal to k, and the order of growth of the right-hand side of (3.10) is less than k. This is a contradiction.

Case 2. Suppose that 1s<k. Thus, there are two subcases: (1) a(z)0; (2) a(z)0.

Subcase 2.1. Suppose that a(z)0. Then, by (3.1), we can obtain

j = 0 n ( 1 ) n j C n j H(z+jη) e h ( z + j η ) h ( z ) H(z) e P ( z ) =b(z) e h ( z ) .
(3.11)

We assert that b(z)0. In fact, if b(z)0, then (2.15) obviously holds. Hence, using the same method as in the proof of Lemma 2.9, by Lemma 2.5 and Remarks 2.2-2.3, we can get that σ(a)1, a contradiction. Hence, b(z)0. Since degh=k, deg(h(z+jη)h(z))=k1 (j=1,2,,n), degP=s<k and σ(H)<k, we see that the order of growth of the left-hand side of (3.11) is less than k, and the order of growth of the right-hand side of (3.11) is equal to k. This is a contradiction.

Subcase 2.2. Suppose that a(z)0. Then, by (3.1), we obtain

j = 1 n ( 1 ) n j C n j H ( z + j η ) H ( z ) e h ( z + j η ) h ( z ) + ( 1 ) n = e P ( z ) .
(3.12)

Thus, there are two subcases: (1) n=1; (2) n2.

Subcase 2.2.1. Suppose that n=1. Then (3.12) can be rewritten as

H ( z + η ) H ( z ) e h ( z + η ) h ( z ) 1= e P ( z ) .
(3.13)

By (3.13), we see that H ( z + η ) H ( z ) is a nonzero entire function. Set σ(H)= σ 4 . Then σ 4 <σ(f)=k. By Lemma 2.1, we see that for any given ε 4 (0<3 ε 4 <k σ 4 ), there exists a set E 4 (1,) of finite logarithmic measure such that for all z satisfying |z|=r[0,1] E 4 , we have

exp { r σ 4 1 + ε 4 } | H ( z + η ) H ( z ) |exp { r σ 4 1 + ε 4 } .
(3.14)

Since H ( z + η ) H ( z ) is an entire function, by (3.13), we have

T ( r , H ( z + η ) H ( z ) ) =m ( r , H ( z + η ) H ( z ) ) r σ 4 1 + ε 4 ,

so that

σ ( H ( z + η ) H ( z ) ) σ 4 1+ ε 4 <k1.
(3.15)

Since s<k, we can see that degPk1. If degP<k1, then, by (3.15) and deg(h(z+η)h(z))=k1, we can see that the order of growth of the left-hand side of (3.13) is equal to k1, and the order of growth of the right-hand side of (3.13) is equal to degP which is less than k1. This is a contradiction.

If degP=k1, then since H ( z + η ) H ( z ) is an entire function and deg(h(z+η)h(z))=k11, by (3.15), we can see that the entire function H ( z + η ) H ( z ) e h ( z + η ) h ( z ) has a Borel exceptional value 0, thus the value 1 must be not its Borel exceptional value. Hence, the left-hand side of (3.13), H ( z + η ) H ( z ) e h ( z + η ) h ( z ) 1, has infinitely many zeros, but the right-hand side of (3.13), e P ( z ) , has no zero. This is a contradiction.

Subcase 2.2.2. Now we suppose that n2. Thus, for s (=degP), there are two subcases: (1) s<k1; (2) s=k1.

Subcase 2.2.2.1. Now we suppose that s<k1. Set Q 5 (z)= e h ( z + η ) h ( z ) . Since σ( Q 5 )=k11, Q 5 (z) is a transcendental entire function. Thus, (3.12) can be rewritten as

U 5 ( z , Q 5 ( z ) ) Q 5 (z)= e P ( z ) ( 1 ) n ,
(3.16)

where

U 5 ( z , Q 5 ( z ) ) = H ( z + n η ) H ( z ) Q 5 ( z + ( n 1 ) η ) Q 5 ( z + ( n 2 ) η ) Q 5 ( z + η ) C n 1 H ( z + ( n 1 ) η ) H ( z ) Q 5 ( z + ( n 2 ) η ) Q 5 ( z + ( n 3 ) η ) Q 5 ( z + η ) + + ( 1 ) n 1 C n n 1 H ( z + η ) H ( z ) .
(3.17)

Thus, using the same method as in the proof of Subcase 1.3 in the proof of Lemma 2.9 and noting that σ( e P ( z ) ( 1 ) n )=degP<k1, we have

m ( r , e P ( z ) ( 1 ) n ) =S(r, Q 5 ).

Noting that n2 and so deg U 5 (z, Q 5 )=n11. Using the same method as in the proof of Subcase 1.3 in the proof of Lemma 2.9, we can obtain

T(r, Q 5 )=m(r, Q 5 )=S(r, Q 5 ).

Clearly, this is a contradiction.

Subcase 2.2.2.2. Now we suppose that s=k1. Thus, (3.12) is written as

j = 1 n ( 1 ) n j C n j H ( z + j η ) H ( z ) e T j ( z ) + ( 1 ) n e P ( z ) =0,
(3.18)

where T j (z)=h(z+jη)h(z) (j=1,2,,n). Thus, by (3.3), we have

T j (z)=jkη a k z k 1 + P k 2 , j (z),
(3.19)

where P k 2 , j (z) is a polynomial with degree at most k2. Thus, we have

T j (z) T t (z)=(jt)kη a k z k 1 + P j , t (z)(1jtn),

where P j , t (z) is a polynomial with degree at most k2.

First, we suppose that there is some j 0 (1 j 0 n) such that j 0 kη a k = b k 1 , that is, deg( T j 0 (z)P(z))k2. Thus, (3.18) can be written as

1 j n , j j 0 ( 1 ) n j C n j H ( z + j η ) H ( z ) e h ( z + j η ) h ( z ) + B j 0 (z) e h ( z + j 0 η ) h ( z ) = ( 1 ) n + 1 ,
(3.20)

where

B j 0 (z)= ( 1 ) n j 0 C n n j 0 H ( z + j 0 η ) H ( z ) e P ( z ) + h ( z ) h ( z + j 0 η ) .

Set Q 6 (z)= e h ( z + η ) h ( z ) and σ(H)= σ 6 . Then (3.20) can be rewritten as

U 6 ( z , Q 6 ( z ) ) Q 6 (z)= ( 1 ) n + 1 ,
(3.21)

where

U 6 ( z , Q 6 ( z ) ) = 1 j n , j j 0 ( 1 ) n j C n n j H ( z + j η ) H ( z ) Q 6 ( z + ( j 1 ) η ) Q 6 ( z + ( j 2 ) η ) Q 6 ( z + η ) + B j 0 ( z ) Q 6 ( z + ( j 0 1 ) η ) Q 6 ( z + ( j 0 2 ) η ) Q 6 ( z + η ) ( j 0 2 ) ,
(3.22)

or

U 6 ( z , Q 6 ( z ) ) = 2 j n ( 1 ) n j C n n j H ( z + j η ) H ( z ) Q 6 ( z + ( j 1 ) η ) Q 6 ( z + ( j 2 ) η ) Q 6 ( z + η ) + B j 0 ( z ) ( j 0 = 1 ) .
(3.23)

Noting that ( 1 ) n + 1 0, we can see that U 6 (z, Q 6 (z))0. Since σ(H)<k and σ( e P ( z ) + h ( z ) h ( z + j 0 η ) )k2<k1, using the same method as in the proof of Subcase 1.3 in the proof of Lemma 2.9, we have

m ( r , B j 0 ( z ) ) =S(r, Q 6 ).
(3.24)

Noting that n2 and so deg U 6 (z, Q 6 )=n11. Combining (3.21)-(3.24), using the same method as in the proof of Subcase 1.3 in the proof of Lemma 2.9, we can obtain

T(r, Q 6 )=m(r, Q 6 )=S(r, Q 6 ).

Clearly, this is a contradiction.

Secondly, we suppose that jkη a k b k 1 for any 1jn. Thus, equation (3.18) can be rewritten as

e P ( z ) = e b k 1 z k 1 e P k 2 ( z ) = j = 0 n ( 1 ) n j C n j H ( z + j η ) H ( z ) e h ( z + j η ) h ( z ) ,
(3.25)

where P k 2 (z)=P(z) b k 1 z k 1 = b k 2 z k 2 + b k 3 z k 3 ++ b 0 . For dealing with equation (3.25), we just compare | b k 1 | with nk|η a k | since nk|η a k |>(n1)k|η a k |>>k|η a k |. Without loss of generality, we suppose that nk|η a k || b k 1 |. Let arg b k 1 = θ 1 , arg(η a k )= θ 2 and σ(H)= σ 7 <k. Take θ 0 such that cos((k1) θ 0 + θ 1 )=1. By Lemma 2.1, we see that for any given ε 7 (0<3 ε 7 <k σ 7 ), there exists a set E 7 (1,) of finite logarithmic measure such that for all z=r e i θ 0 satisfying |z|=r[0,1] E 7 , we have

exp { r σ 7 1 + ε 7 } | H ( z + j η ) H ( z ) |exp { r σ 7 1 + ε 7 } (j=1,,n).
(3.26)

Thus, noting that e P k 2 ( z ) is of regular growth, we can deduce from (3.25) and (3.26) that

| e b k 1 z k 1 | = | e P ( z ) e P k 2 ( z ) | | j = 0 n ( 1 ) j C n j H ( z + ( n j ) η ) H ( z ) e h ( z + ( n j ) η ) h ( z ) | | e b k 2 z k 2 + b k 3 z k 3 + + b 0 | ( n + 1 ) n ! exp { r σ 7 1 + ε 7 } exp { n k | η a k | cos ( ( k 1 ) θ 0 + θ 2 ) r k 1 + O ( r k 2 ) } exp { | b k 2 | 2 r k 2 } ,

that is,

exp { | b k 1 | r k 1 } exp { n k | η a k | cos ( ( k 1 ) θ 0 + θ 2 ) r k 1 + r σ 7 1 + ε 7 + O ( r k 2 ) | b k 2 | 2 r k 2 } exp { n k | η a k | cos ( ( k 1 ) θ 0 + θ 2 ) r k 1 + o ( r k 1 ) } .
(3.27)

We assert that

nk|η a k |cos ( ( k 1 ) θ 0 + θ 2 ) <| b k 1 |.

In fact, if nk|η a k |=| b k 1 |, then, by b k 1 nkη a k , we know that cos((k1) θ 0 + θ 2 )1, that is, cos((k1) θ 0 + θ 2 )<1, and hence nk|η a k |cos((k1) θ 0 + θ 2 )<nk|η a k |=| b k 1 |. If nk|η a k |<| b k 1 |, then we have nk|η a k |cos((k1) θ 0 + θ 2 )nk|η a k |<| b k 1 |.

Thus, taking a positive constant ε 8 (0< ε 8 < | b k 1 | n k | η a k | cos ( ( k 1 ) θ 0 + θ 2 ) 3 ), we can deduce from (3.27) that

exp { | b k 1 | r k 1 } exp { n k | η a k | cos ( ( k 1 ) θ 0 + θ 2 ) r k 1 + o ( r k 1 ) } exp { ( | b k 1 | ε 8 ) r k 1 } ,

a contradiction. Thus, we have proved that P is only a constant and (3.2) holds.

Second step. Applying Lemma 2.9 to (3.2), we can obtain the conclusion.

Thus, Theorem 1.1 is proved.

References

  1. Hayman WK: Meromorphic Functions. Clarendon, Oxford; 1964.

    MATH  Google Scholar 

  2. Laine I: Nevanlinna Theory and Complex Differential Equations. de Gruyter, Berlin; 1993.

    Book  MATH  Google Scholar 

  3. Yang L: Value Distribution Theory. Science Press, Beijing; 1993.

    MATH  Google Scholar 

  4. Bergweiler W, Langley JK: Zeros of differences of meromorphic functions. Math. Proc. Camb. Philos. Soc. 2007, 142: 133–147. 10.1017/S0305004106009777

    Article  MathSciNet  MATH  Google Scholar 

  5. Whittaker JM Cambridge Tract 33. In Interpolatory Function Theory. Cambridge University Press, Cambridge; 1935.

    Google Scholar 

  6. Nevanlinna R: Einige Eindentigkeitssätze in der Theorie der meromorphen Funktionen. Acta Math. 1926, 48: 367–391. 10.1007/BF02565342

    Article  MathSciNet  MATH  Google Scholar 

  7. Gundersen G: Meromorphic functions that share four values. Trans. Am. Math. Soc. 1983, 277: 545–567. 10.1090/S0002-9947-1983-0694375-0

    Article  MathSciNet  MATH  Google Scholar 

  8. Gundersen G: Correction to meromorphic functions that share four values. Trans. Am. Math. Soc. 1987, 304: 847–850.

    MathSciNet  MATH  Google Scholar 

  9. Mues E: Meromorphic functions sharing four values. Complex Var. Elliptic Equ. 1989, 12: 167–179.

    Article  MathSciNet  MATH  Google Scholar 

  10. Brück R: On entire functions which share one value CM with their first derivative. Results Math. 1996, 30: 133–147.

    Article  MathSciNet  MATH  Google Scholar 

  11. Yang CC, Yi HX: Uniqueness Theory of Meromorphic Functions. Kluwer Academic, Dordrecht; 2003.

    Book  MATH  Google Scholar 

  12. Gundersen G, Yang LZ: Entire functions that share one values with one or two of their derivatives. J. Math. Anal. Appl. 1998, 223(1):88–95. 10.1006/jmaa.1998.5959

    Article  MathSciNet  MATH  Google Scholar 

  13. Chen ZX, Shon KH: On conjecture of R. Brück, concerning the entire function sharing one value CM with its derivative. Taiwan. J. Math. 2004, 8(2):235–244.

    MathSciNet  MATH  Google Scholar 

  14. Halburd RG, Korhonen R: Nevanlinna theory for the difference operator. Ann. Acad. Sci. Fenn., Math. 2006, 31: 463–478.

    MathSciNet  MATH  Google Scholar 

  15. Heittokangas J, Korhonen R, Laine I, Rieppo J, Zhang J: Value sharing results for shifts of meromorphic functions, and sufficient conditions for periodicity. J. Math. Anal. Appl. 2009, 355: 352–363. 10.1016/j.jmaa.2009.01.053

    Article  MathSciNet  MATH  Google Scholar 

  16. Heittokangas J, Korhonen R, Laine I, Rieppo J: Uniqueness of meromorphic functions sharing values with their shifts. Complex Var. Elliptic Equ. 2011, 56: 81–92. 10.1080/17476930903394770

    Article  MathSciNet  MATH  Google Scholar 

  17. Liu K: Meromorphic functions sharing a set with applications to difference equations. J. Math. Anal. Appl. 2009, 359: 384–393. 10.1016/j.jmaa.2009.05.061

    Article  MathSciNet  MATH  Google Scholar 

  18. Chen ZX: Growth and zeros of meromorphic solutions of some linear difference equation. J. Math. Anal. Appl. 2011, 373: 235–241. 10.1016/j.jmaa.2010.06.049

    Article  MathSciNet  Google Scholar 

  19. Chen ZX, Shon KH: Value distribution of meromorphic solutions of certain difference Painlevé equations. J. Math. Anal. Appl. 2010, 364: 556–566. 10.1016/j.jmaa.2009.10.021

    Article  MathSciNet  MATH  Google Scholar 

  20. Chen ZX, Yi HX: On sharing values of meromorphic functions and their difference. Results Math. 2013, 63(1):557–565.

    Article  MathSciNet  MATH  Google Scholar 

  21. Chiang YM, Feng SJ:On the Nevanlinna characteristic of f(z+η) and difference equations in the complex plane. Ramanujan J. 2008, 16: 105–129. 10.1007/s11139-007-9101-1

    Article  MathSciNet  MATH  Google Scholar 

  22. Halburd RG, Korhonen R: Difference analogue of the lemma on the logarithmic derivative with applications to difference equations. J. Math. Anal. Appl. 2006, 314: 477–487. 10.1016/j.jmaa.2005.04.010

    Article  MathSciNet  MATH  Google Scholar 

  23. Heittokangas J, Korhonen R, Laine I, Rieppo J, Tohge K: Complex difference equations of Malmquist type. Comput. Methods Funct. Theory 2001, 1: 27–39. 10.1007/BF03320974

    Article  MathSciNet  MATH  Google Scholar 

  24. Laine I, Yang CC: Value distribution of difference polynomials. Proc. Jpn. Acad., Ser. A, Math. Sci. 2007, 83: 148–151. 10.3792/pjaa.83.148

    Article  MathSciNet  MATH  Google Scholar 

  25. Laine I, Yang CC: Clunie theorems for difference and q -difference polynomials. J. Lond. Math. Soc. 2007, 76(3):556–566. 10.1112/jlms/jdm073

    Article  MathSciNet  MATH  Google Scholar 

  26. Gross F: Factorization of Meromorphic Functions. G.P.O., Washington; 1972.

    MATH  Google Scholar 

  27. Chen ZX: On growth of meromorphic solution for linear difference equations. Abstr. Appl. Anal. 2013., 2013: Article ID 619296

    Google Scholar 

Download references

Acknowledgements

This research was supported by the National Natural Science Foundation of China (Nos. 11171119, 11226090) and supported by the Natural Science Foundation of Guangdong Province, China (No. S2013040014347).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Zong-Xuan Chen.

Additional information

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

All authors contributed equally to the manuscript and read and approved the final manuscript.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 2.0 International License (https://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

Chen, CX., Chen, ZX. A note on entire functions and their differences. J Inequal Appl 2013, 587 (2013). https://doi.org/10.1186/1029-242X-2013-587

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1029-242X-2013-587

Keywords