Abstract

We study the normality of families of meromorphic functions related to a Hayman conjecture. We consider whether a family of meromorphic functions is normal in if, for every pair of functions and in , and share the value for , and 3, where and are two finite complex numbers. Some examples show that the conditions in our results are the best possible.

1. Introduction and Main Results

Let and be two nonconstant meromorphic functions in a domain , and let be a finite complex value. We say that and share CM (or IM) in provided that and have the same zeros counting (or ignoring) multiplicity in . When the zeros of mean the poles of (see [1]). It is assumed that the reader is familiar with the standard notations and the basic results of Nevanlinna’s value-distribution theory ([24] or [1]).

Bloch’s principle [5] states that every condition which reduces a meromorphic function in the plane to be a constant forces a family of meromorphic functions in a domain normal. Although the principle is false in general (see [6]), many authors proved the normality criterion for families of meromorphic functions corresponding to Liouville-Picard-type theorem (see [7] or [4]).

It is also more interesting to find normality criteria from the point of view of shared values. In this area, Schwick [8] first proved an interesting result that a family of meromorphic functions in a domain is normal if it in every function in it shares three distinct finite complex numbers with its first derivative. Later, more results about normality criteria concerning shared values can be found, for instance, see [911] and so on. In recent years, this subject has attracted the attention of many researchers worldwide.

We now first introduce a normality criterion related to a Hayman normal conjecture [12].

Theorem 1.1. Let be a family of holomorphic (meromorphic) functions defined in a domain , . If for each function and , then is normal in .

The results for the holomorphic case are due to Drasin [7] for , Pang [13] for , and Chen and Fang [14] for , Ye [15] for , and Chen and Gu [16] for the generalized result with and replaced by meromorphic functions. The results for the meromorphic case are due to Li [17], Li [18], and Langley [19] for , Pang [13] for , Chen and Fang [14] for , and Zalcman [20] for , obtained independently.

When and is meromorphic, Theorem 1.1 is not valid in general. Fang and Yuan [21] gave an example to this, and moreover a result added other conditions below.

Example 1.2. The family of meromorphic functions is not normal in . This is deduced by , as , and Marty’s criterion [2], although, for any .

Here denotes the spherical derivative

Theorem 1.3. Let be a family of meromorphic functions in a domain and . If and the poles of are of multiplicity for each , then is normal in .

In 2008, by the ideas of shared values, Zhang [11] proved the following.

Theorem 1.4. Let be a family of meromorphic (holomorphic) functions in , a positive integer, and , two finite complex numbers such that . If and, for every pair of functions and in , and share the value , then is normal in .

Example 1.5 (see [11]). The family of meromorphic functions is not normal in . Obviously . So, for each pair and share the value 0 in , but is not normal at the point since , as .

Remark 1.6. Example 1.5 shows that Theorem 1.4 is not valid when , and the condition is the best possible for the meromorphic case.

It is natural to ask under what conditions Theorem 1.4 holds for . In this paper, we answer the above question and prove the following results.

Theorem 1.7. Let be a family of meromorphic functions in and and two finite complex numbers such that . Suppose that each has no simple pole. If and share the value for every pair of functions and in , then is normal in .

Remark 1.8. Example 1.5 shows that the condition added in Theorem 1.7 about the multiplicity of poles of is the best possible.

Theorem 1.9. Let be a family of meromorphic functions in and and two finite complex numbers such that . Suppose that admits the zeros of multiple and the poles of multiplicity for each . If and share the value for every pair of functions and in , then is normal in .

Remark 1.10. Example 1.2 shows that the condition added in Theorem 1.9 about the multiplicity of poles and zeros of is the best possible.

Theorem 1.11. Let be a family of meromorphic functions in and and two nonzero finite complex numbers. Suppose that and that its poles are multiple for each . If and share the value for every pair of functions and in , then is normal in .

Corollary 1.12. Let be a family of holomorphic functions in and and two finite complex numbers such that . Suppose that for each . If and share the value for every pair of functions and in , then is normal in .

Example 1.13. The family of holomorphic functions is not normal in . Obviously . So, for each pair and share the value in . On the other hand, , as . This implies that the family fails to be equicontinuous at 0, and thus is not normal at 0.

Theorem 1.14. Let be a family of meromorphic functions in and and two finite complex numbers such that . Suppose that and for each . Then is normal in .

Example 1.15. The family of holomorphic functions is normal in . Obviously and . So, for each pair and share the value 1 in . Corollary 1.12 implies that the family is normal in .

Example 1.16. The family of meromorphic functions is normal in . The reason is that the conditions of Theorem 1.14 hold that and in .

Remark 1.17. Example 1.13 shows that Theorem 1.4 is not valid when and in the holomorphic case and the condition is necessary in Theorem 1.11, Corollary 1.12. Both Examples 1.15 and 1.16 tell us that Corollary 1.12 and Theorem 1.14 occur.

2. Preliminary Lemmas

In order to prove our result, we need the following lemmas. The first one extends a famous result by Zalcman [22] concerning normal families.

Lemma 2.1 (see [23]). Let be a family of meromorphic functions on the unit disc satisfying all zeros of functions in   which have multiplicity and all poles of functions in which have multiplicity . Let be a real number satisfying . Then is not normal at 0 if and only if there exist (a)a number ,(b)points with ,(c)functions ,(d)positive numbers such that converges spherically uniformly on each compact subset of to a nonconstant meromorphic function , whose all zeros have multiplicity and all poles have multiplicity and order is at most 2.

Remark 2.2. If is a family of holomorphic functions on the unit disc in Lemma 2.1, then is a nonconstant entire function whose order is at most 1.

The order of is defined by using Nevanlinna’s characteristic function :

Lemma 2.3 (see [24] or [25]). Let be a meromorphic function and . If has neither simple zero nor simple pole and , then is constant.

Lemma 2.4 (see [26]). Let be a transcendental meromorphic function of finite order in , with no simple zero; then assumes every nonzero finite value infinitely often.

Lemma 2.5 (see [3]). Let be a meromorphic function in ; then

Remark 2.6. Both (2.2) and (2.3) are called the Hayman inequality and Milloux inequality, respectively.

3. Proof of the Results

Proof of Theorem 1.7. Suppose that is not normal in . Then there exists at least one point such that is not normal at the point . Without loss of generality we assume that . By Lemma 2.1, there exist points , positive numbers , and functions such that locally uniformly with respect to the spherical metric, where is a nonconstant meromorphic function in . Moreover, the order of is no greater than 2 and the poles of are of multiplicity .
From (3.1) we know that in locally uniformly with respect to the spherical metric, where is the set of all poles of .
If , then , where is a constant. This contradicts with the idea that the poles of are of multiplicity . So .
If , then . Set ; then . By Lemma 2.3, is a constant, so is also a constant which is a contradiction with being a nonconstant. Hence, is a nonconstant meromorphic function and has at least one zero.
Next we prove that has just a unique zero. By contraries, let and be two distinct zeros of , and choose small enough such that , where and . From (3.3), by Hurwitz's theorem, there exist points , such that for sufficiently large
By the hypothesis that, for each pair of functions and in , and share in , we know that, for any positive integer
Fix , take , and note that , ; then . Since the zeros of have no accumulation point, Hence, . This contradicts with , and . So has just a unique zero, which can be denoted by .
Set again; then . So has only a unique zero . Therefore, is a multiple pole of , or else a zero of . If is a multiple pole of , since has only one zero , then . By Lemma 2.3 again, is a constant which contradicts with the idea that is a nonconstant.
So has no multiple pole and has just a unique zero . By Lemma 2.3, is not any transcendental function.
If is a nonconstant polynomial, then , where A is a nonzero constant and is a positive integer because the poles of are of multiplicity . So the zeros of are of multiplicity , and thus, . Set ; then and . Note that the zeros of are of multiplicity . But has only one zero , and so has only the same zero . Hence which contradicts with . Therefore and are rational functions which are not polynomials and has just a unique zero .
Next we prove that there exists no rational function such as . Noting that , has no multiple pole, and the zeros of are of multiplicity , we can set where is a nonzero constant, . For stating briefly, denote From (3.7), where are polynomials. Since has only a unique zero , set where is a nonzero constant, and so where is a polynomial. From (3.9), we also have where is also a polynomial.
Let denote the degree of a polynomial .
From (3.9) and (3.10), Similarly from (3.12) and (3.13) and noting (3.14),
Note that . It follows from (3.9) and (3.11) that and . Thus , and then is a factor of . Hence we get that . Combining (3.12) and (3.13) we also have . By (3.15) we obtain
Since , we know by (3.17) that
If , by (3.16), then . Noting (3.18), we obtain , a contradiction.
If , from (3.9) and (3.11), . Noting that , so . From (3.10), then , and then . Noting that , . By (3.17), . From (3.18), we obtain , a contradiction.
The proof of Theorem 1.7 is complete.

Proof of Theorem 1.9. Similarly with the proof of Theorem 1.7, we assume that is not normal at . Then, by Lemma 2.1, there are a sequence of complex numbers , , a sequence of functions , and a sequence of positive numbers such that converges uniformly with respect to the spherical metric to a nonconstant mermorphic function and all zeros and poles of are multiple. Moreover, is of order 2 at most.
Thus also locally uniformly with respect to the spherical metric.
If , then where is a constant. This contradicts with the idea that the poles of are multiple. So .
If , then . By Lemma 2.3, is a constant which contradicts with our conclusion. Hence, is a nonconstant meromorphic function and has at least one zero.
As the same argument in the proof of Theorem 1.7, we obtain that has only one distinct zero denoted by .
Set ; then . So has only a unique zero . Therefore is a multiple pole of or a zero of . If is a multiple pole of , since has only one zero , then . By Lemma 2.3 again, is a constant, which is a contradiction.
Hence is an entire function with no simple zero and growth order at most 2 and has just a unique zero . By Lemma 2.4, is not any transcendental function. Therefore is a nonconstant polynomial and has the form that , where is a nonzero constant and is a positive integer because the poles of are of multiplicity . So the zeros of are of multiplicity ; thus, . Note that has only one zero , and so has only the same zero too. Hence which contradicts with .
The proof of Theorem 1.9 is complete.

Proof of Theorem 1.11. Similarly with the proof of Theorem 1.7, we assume that is not normal at . Then, by Lemma 2.1, there are a sequence of complex numbers , , a sequence of functions , and a sequence of positive numbers such that converges uniformly with respect to the spherical metric to a nonconstant meromorphic function whose poles are multiple and . Moreover, is of order 2 at most.
Thus also locally uniformly with respect to the spherical metric.
If , then , where is a constant. This contradicts with . So .
If , then by Milloux inequality (2.3) of Lemma 2.5 we have From (3.23) we know that is a constant which contradicts with our conclusion. Hence, is a nonconstant meromorphic function and has at least one zero.
As the same argument in the proof of Theorem 1.7, we obtain that has only one distinct zero denoted by . Thus the Hayman inequality (2.2) of Lemma 2.5 implies that is a rational function of degree 4 at most. Noting that and has no simple pole, we obtain that has at most two distinct poles. Using the Milloux inequality (2.3) of Lemma 2.5 again we get that has at most one distinct pole. Hence we can write , where and are two finite complex numbers. Simple calculation shows that has at least three distinct zeros. This is impossible.
The proof of Theorem 1.11 is complete.

Proof of Theorem 1.14. Similarly with the proof of Theorem 1.11, we have that (3.22) also holds. Moreover, and is of order 2 at most.
If , then , where is a constant. This contradicts with .
If , then by the Hayman inequality (2.2) of Lemma 2.5 we have From (3.24) we know that is a constant which contradicts with our conclusion.
The proof of Theorem 1.8 is complete.

Acknowledgments

The authors would like to express their hearty thanks to Professor Qingcai Zhang for supplying them with his helpful reprint. The authors wish to thank the managing editor and referees for their very helpful comments and useful suggestions. This work was completed with the support of the NSF of China (10771220) and Doctorial Point Fund of National Education Ministry of China (200810780002).