• Ei tuloksia

Lower order and Baker wandering domains of solutions to differential equations with coefficients of exponential growth

N/A
N/A
Info
Lataa
Protected

Academic year: 2022

Jaa "Lower order and Baker wandering domains of solutions to differential equations with coefficients of exponential growth"

Copied!
17
0
0

Kokoteksti

(1)

DSpace https://erepo.uef.fi

Rinnakkaistallenteet Luonnontieteiden ja metsätieteiden tiedekunta

2019

Lower order and Baker wandering domains of solutions to differential equations with coefficients of

exponential growth

Korhonen, Risto

Elsevier BV

Tieteelliset aikakauslehtiartikkelit

© Elsevier Inc.

CC BY-NC-ND https://creativecommons.org/licenses/by-nc-nd/4.0/

http://dx.doi.org/10.1016/j.jmaa.2019.07.007

https://erepo.uef.fi/handle/123456789/7696

Downloaded from University of Eastern Finland's eRepository

(2)

Lower order and Baker wandering domains of solutions to differential equations with coefficients of exponential growth

Risto Korhonen, Jun Wang, Zhuan Ye

PII: S0022-247X(19)30568-2

DOI: https://doi.org/10.1016/j.jmaa.2019.07.007 Reference: YJMAA 23317

To appear in: Journal of Mathematical Analysis and Applications Received date: 28 August 2018

Please cite this article in press as: R. Korhonen et al., Lower order and Baker wandering domains of solutions to differential equations with coefficients of exponential growth,J. Math. Anal. Appl.(2019), https://doi.org/10.1016/j.jmaa.2019.07.007

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

(3)

SOLUTIONS TO DIFFERENTIAL EQUATIONS WITH COEFFICIENTS OF EXPONENTIAL GROWTH

RISTO KORHONEN, JUN WANG, AND ZHUAN YE

Abstract. We investigate transcendental entire solutions of complex differen- tial equationsf+A(z)f =H(z), where the entire function A(z) has a growth property similar to the exponential functions, andH(z) is an entire function of order less than that ofA. We first prove that the lower order of the entire solu- tion to the equation is infinity. By using our result on the lower order, we prove the entire solution does not bear any Baker wandering domains.

1. Introduction and main results

Let f : C C be a transcendental meromorphic function, where C is the complex plane and C=C∪ {∞}. We use ρ(f) and μ(f) to denote the order and the lower order off, respectively, which are defined as [20, Definition 1.6]

ρ(f) = lim sup

r→∞

log+T(r, f)

logr , μ(f) = lim inf

r→∞

log+T(r, f) logr .

Nevanlinna theory is an important tool in this paper, its usual notations and basic results come mainly from [7, 9, 13, 20].

The nth iterate fn of f is defined for all z C except for a countable set of poles of f, f2,· · · , fn1. The Fatou set F(f) of f is the set where the family of iterates{fn} is well defined and forms a normal family. The complement of F(f) is called the Julia set J(f) of f. It is well-known that F(f) is open and forward invariant. A basic theory of complex dynamics can be found in, for instance, [5]

for meromorphic functions, and [17] for entire functions.

It is well-known that if U is a Fatou component U then U is connected and fn(U) must be contained in a Fatou component Un. If all Un are different, then U is called a wandering domain of f. By Sullivan’s theorem, rational functions have no wandering domains, but wandering domains do occur for transcendental functions. For example, f(z) = z + sinz + 2π has a bounded simply connected wandering domain, see [3] and [17, p. 311]. Baker [1, 2] also gave examples of multiply connected wandering domains, which are now known as Baker wandering domains.

For the convenience of the reader, we give the following definition.

2010 Mathematics Subject Classification. 34M10, 37F10, 30D35.

Key words and phrases. Lower order, Baker wandering domain, entire function, linear differ- ential equation.

1

(4)

Definition 1. Let f be a meromorphic function in C. For a wandering domain U, if all Un are multiply connected components ofF(f) which surround the origin, and the Euclidean distance dist(0, Un) + as n +, then U is called a Baker wandering domain.

Clearly, transcendental entire functions without Baker wandering domains have only simply connected Fatou components. There are already some criteria of non- existence of Baker wandering domains [3, 5, 21]. For example, if the entire function f has a path to on whichf is bounded, then all components ofF(f) are simply connected (see [3, Corollary]). On the other hand, if f is entire with a finite deficient value of Nevanlinna, then every component of F(f) is simply connected [21, Corollary 4].

We recall a growth property of complex exponential functionsep(z), wherep(z) = αnzn+· · · is a polynomial of degree n, [15, p. 254]. The complex plane is divided into 2n equal open sectors by the rays

argz =argαn

n + (2j1)π

2n (j = 0,1,· · · ,2n).

In each of these sectors, ep(z) has two possible growth types: it either (1) blows up exponentially; or (2) decays to zero exponentially. In addition,ρ(ep(z)) =μ(ep(z)) = n. By [3, Corollary] and the radial behaviour of ep(z), it is easy to see that ep(z) has no Baker wandering domains. Inspired by the observation, we want to extend our studies to any transcendental entire functions with nice radial growth or decay properties.

Definition 2. Let A(z) be transcendental entire with ρ(A) = μ(A) < , and δA(θ) be a real-valued function defined on [0,2π) which is continuous outside an exceptional set F of finitely many points. Further, let c, d be positive constants.

Then for any given θ [0,2π)\F, there are a constant τ, and positive constants R=R(θ) andM =M(θ) such that when |z|=r > R,

(A1) |A(re)| ≥exp{cδA(θ)rd} if δA(θ)>0, (A2) |A(re)| ≤M rτ if δA(θ)<0,

where τ < 2(ρ(A)1). Especially, for any θ [0,2π)\F with δA(θ) > 0, there existsl =l(θ)>0 such that sup{R(˜θ),θ˜−l, θ+l)}<∞. We call such A(z) as a function of exponential growth and denote the set of all these functions byA. It is also worth noting that there exists τ > 1 for ρ(A) > 1/2, and when ρ(A) > 1, we can even take positive τ. The growth type (A1) must always exist because A(z) is transcendental. Clearly, ep(z) belongs to A, and A contains many elements from the class of exponential polynomials, which consists of functions of the formm

j=1Pj(z)eαjz, where Pj(z) are nonzero polynomials and αj are distinct constants. The study of this class of functions is a classical topic in complex analysis [4, Chapter 3].

(5)

If there really exists one θ such that δA(θ) < 0 and τ < 0, then in this case, it follows from [3, Corollary] again that there is no Baker wandering domains for suchA(z)∈ A.

Example 1. Letnbe a positive integer andp(z) = αnzn+· · ·0be a polynomial of degree n. We take A(z) = ak(z)ekp(z)+· · ·+as(z)esp(z), which is a polynomial inep(z) of degree k with polynomial coefficients, and 0< s≤k. Set

δ(p, θ) = Re(αn) cos(nθ)Im(αn) sin(nθ).

It is known that ep(z)= exp{δ(p, θ)|z|n+O(|z|n1)}. Define δA(θ) = max{kδ(p, θ), sδ(p, θ)}.

Thus, it is clear that A satisfies A1. Then δA(θ) = sδ(p, θ) for θ with δA(θ) < 0 and further, for any τ R, we have

|A(re)| ≤rdegasexp{s

2δ(p, θ)rn} ≤rτ, as r→ ∞. It follows thatA satisfies A2, consequently, A ∈ A.

Example 2. Suppose that A(z) =b1(z)ep1(z)+b2(z)ep2(z) wherepi(z) =αk,izk+

· · ·,(i = 1,2) are two polynomials of degree k > 0, and bi(z)(i = 1,2) are also polynomials. Now we take δA(θ) = max{δ(p1, θ), δ(p2, θ)}. When αk,1k,2 is not negative, we know thatG:={θ, δ(pi, θ)<0(i= 1,2)} =∅.Thus for θ ∈G, both δ(p1, θ) and δ(p2, θ) are negative, so it follows that for any τ R,

|A(re)| ≤rdegb1exp{1

2δ(p1, θ)rn}+rdegb2exp{1

2δ(p2, θ)rn} ≤rτ, asr tends to .

Example 3. Recall Mittag-Leffler function Eα(z) =

n=0

zn

Γ(α1n+ 1), 0< α <∞, has the uniform asymptotic behavior [7, Chapter 1, (5.40)],

Eα(z) =

⎧⎨

αexp (zα) +O

|z|1

, |arg(z)| ≤ π, O

|z|1

, π

≤ |arg(z)| ≤π.

Clearly,A(z) =Eα(z)∈ A with α >1/2, whereδA(θ) =1 for θ (π , π ), and δA(θ) = cos(αθ) for [0,2π)\(π , π ).

For solutions of complex differential equations with rational coefficients, Zheng [22] obtained that every transcendental meromorphic function satisfying linear dif- ferential equation with rational coefficients must have no Baker wandering domains and its Julia set has an unbounded component.

As for complex differential equations with transcendental coefficients, Huang and Wang [11] recently considered linear differential equations

f+A(z)f = 0, (1.1)

(6)

where entire coefficientA(z) has radial growth behavior similar to functions in our family A, but A(z) decays to zero exponentially in some rays argz = θ. They proved thatE(z) have no Baker wandering domains whereE(z) = f1f2, andf1, f2 are two linearly independent solutions of (1.1).

In this paper, we will apply the generalized Fuchs’ small arcs lemma [16, Lemma 5] to prove that E(z) to (1.1) with A∈ A and every nonzero solution to

f+A(z)f =H(z), A(z)∈ A, (1.2)

whereH is entire withρ(H)< ρ(A), are of infinite lower order. Based on this fact, we further consider the non-existence of Baker wandering domains for the solutions and the product E, even for their derivatives and primitives.

Note that the lower order can be quite different from the order, since there even exists an entire function of order ρ= and lower order μ= 0, see [7, p.238]. As we know, there are only some attempts to associate the lower order with complex differential equations, see [14], and references therein. We feel a need to state the results on the lower order separately.

Theorem 1. The lower order of any non-zero solution f of (1.2) is infinite.

Theorem 2. The lower order ofE is infinite, whereE(z) =f1(z)f2(z), andf1, f2 are linearly independent solutions of (1.1) withA∈ A.

Next, with these two theorems in hand, we will prove the non-existence of Baker wandering domains for the solutions f to (1.2) and E to (1.1). Before stating our results, we use f(n)(n Z) to denote the derivatives for n N and the anti- derivatives for −n∈N, respectively, and f(0) =f.

Theorem 3. If there exists one θ F such that δA(θ) < 0, then every non-zero solution f of (1.2), along with its f(n)(nZ), have no Baker wandering domains.

Theorem 4. If there exists one θ ∈F such thatδA(θ)<0, then allE(n)(z)(nZ) have no Baker wandering domains, whereE(z) =f1(z)f2(z), andf1f2 are linearly independent solutions of (1.1) with A∈ A.

Theorems 1-4 cover a broader range of differential equations than that in [11, Theorem 1.1]. Besides the two examples given in [11], we have another two exam- ples to illustrate some equations that can occur in Theorems 1-4.

Example 4.Letm be any positive integer and f0 = exp(zm1+ 0zetmdt). Then μ(f0) = and f0 satisfies the equation (1.1) with

−A(z) = (m−1)(m2)zm3+(m1)2z2(m2)+[2(m1)zm2−mzm1]ezm+e2zm. Clearly, ρ(A) =m, and δA(θ) = cos(mθ). Thus, for θ with δA(θ)<0,

|A(re)| ≤2(m1)2r2(m2), for all large r.

Also, 2(m2)<2(ρ(A)1). Thus, A∈ A.

Example 5.The function f0 = exp(aez2 +bz) satisfies the equation f(4a2z2e2z2 + 2a(2z2 + 2bz+ 1)ez2 +b2)f = 0.

(7)

2. lower order of the solutions and E(z) To prove Theorems 1 and 2, we need the following lemmas.

Lemma 1. [11, Lemma 2.7] Let g(z) be an entire function, and ν(r, g) be the central index of g. Then

ρ(g) = lim sup

r→∞

log+ν(r, g)

logr , μ(g) = lim inf

r→∞

log+ν(r, g) logr .

Before introducing a version of Fuchs’ small-arc lemma, we need recall the upper and the lower logarithmic densities of a given set G [0,). The logarithmic measure ofGisml(G) = G(1/t)dt. The upper and the lower logarithmic densities of Gare defined by

logdens(G) = lim sup

r→∞

ml(G[1, r])

logr , logdens(G) = lim inf

r→∞

ml(G[1, r]) logr . It is easy to see 0logdens(G)≤logdens(G)≤1.

The following lemma is a routine consequence of Fuchs’ small-arc lemma [6, Lemma 1], the version here could be found in [16, Lemma 5].

Lemma 2. Let g(z) be a non-constant meromorphic function with finite lower order and let 0 < η < 1. Then there exists a positive constant L and a subset Gη [0,) of upper logarithmic density at least 1 −η such that if r Gη is sufficiently large andIr is a subinterval of [0,2π] of length m then

r

Ir

g(re) g(re)

dθ < Lmlog2πe m

T(r, g).

Lemma 3. Let f(z) be an entire function of finite lower order μ, M(r, f) =

|f(rer)| for every r, and η (0,1). Given s (0,1), there exist a constant l0 (0,1/2) and a set Fη with logdens(Fη)>1−η such that

eM(r, f)1s ≤ |f(re)|, (2.1) for sufficiently larger ∈Fη and all θ such that |θ−θr| ≤l0.

Proof. Following the argument in [18, Lemma 2.4] and Lemma 2, we obtain logM(r, f)−Lmlog2πe

m

T(r, f)≤ |logf(re)|+ 2π log|f(re)|+ 5π, (2.2)

for large r Fη with logdens(Fη) 1−η and |θ−θr| ≤ m. Obviously, we can take m so small that Lmlog2πe

m

< s. At the same time, (2.1) will follow from

(2.2).

Before the next lemma, we recall one more definition: a meromorphic function α(z) is of growth S(r, f) if T(r, α) = o(T(r, f)) as r → ∞, possibly outside a set of r with finite logarithmic measure.

(8)

Lemma 4. ([19, Lemma 2.4]) Let f(z) be a non-constant entire function of lower order μ(f) = μ < ∞. Suppose that αj(z)(j = 1,2,· · · , m) are meromorphic functions of growth S(r, f). Then there exists a setE (1,)of lower logarithmic density 1, such that

log+M(r, αj)

log+M(r, f) 0, (j = 1,2,· · · , m) (2.3) holds simultaneously for r∈E, r→ ∞.

Remark 1. Since the entire functionf in Lemma 4 is not constant, it follows that M(r, f) tends to infinity as r→ ∞. Making use of this fact and (2.3) yields

M(r, αj)

M(r, f)1s 0 (2.4)

forr ∈E, r→ ∞, and any arbitrary constants [0,1).

Lemma 5. Suppose f(z) and g(z) are two non-constant meromorphic functions.

If ρ(f)< μ(g), then T(r, f) = o(T(r, g)).

Proof. From the definition of order and lower order, for 0 < ε < (μ(g)−ρ(f))/3, there existsR >0 such that we have

T(r, f)≤rρ(f)+ε ≤rμ(g) ≤rεT(r, g), for r≥R.

This immediately leads to T(r, f) = o(T(r, g)).

Proof of Theorem 1. Suppose first that f is a nonzero solution of (1.2) with μ(f)<∞. It is easy to see ρ(f)≥ρ(A)> ρ(H).

Claim 1. Moreover, we also have μ(f) ρ(A). Since ρ(H) < ρ(A) = μ(A), Lemma 5 gives

T(r, H) = o(T(r, A)).

We rewrite (1.2) as

f(z)

f(z) =−A(z) + H(z)

f(z). (2.6)

Then by the first fundamental theorem, it is easy to see T(r, A)≤T(r,f

f ) +T(r,H

f )≤T(r, f) + 2T(r, f) +T(r, H) +O(1).

This implies

T(r, A)5T(r, f) +o(T(r, f)) +o(T(r, A)),

outside a set of r with finite linear measure. By [13, Lemma 1.1.1], a lemma from real analysis, we know that for α >1, there exists r0 >0 such that

T(r, A)6T(αr, f), r ≥r0.

This leads that μ(f)≥μ(A) =ρ(A). Then, Claim 1 is proved.

(9)

Applying Lemma 4 to H(z) yields that there is a set F1 with logdens(F1) = 1 such that for any given constant s∈[0,1)

|z2H(z)|

M(r, f)1s 0, for r ∈F1 → ∞. (2.7) By Gundersen’s estimation for logarithmic derivatives [8], we have

f(z) f(z)

≤B

T(2r, f)

r log2rlogT(2r, f) 2

≤B(T(2r, f))4, (2.8) for all z =re with|z| ∈F2[0,1], whereml(F2)<∞, and for a positive constant B. The Wiman-Valiron theorem (see [12, p. 187-199] and [13, p. 51]) states that there is a setF3 with ml(F3)<∞ such that

f(k)(z) f(z) =

ν(r, f) z

k

(1 +o(1)), (2.9)

for z satisfying |f(z)| = M(r, f) and |z| = r F3. We may take θr such that

|f(rer)|=M(r, f). Recalling Lemma 3, there exists l0 (0,12) and a set Fη with logdens(Fη) = 1−η such that

eM(r, f)1s ≤ |f(re)|, (2.10) for all sufficiently large r∈Fη and |θ−θr| ≤l0.

Since the characteristic functions of F1 and Fη satisfy χF1Fη(t) =χF1(t) +χFη(t)−χF1Fη(t), we obtain

logdens(F1) + logdens(Fη)logdens(F1 ∩Fη)logdens(F1∪Fη)1, which implies logdens(F1 ∩Fη) 1−η. Thus, we can take a sequence of points zn=rnen satisfying |f(zn)|=M(rn, f) and

rn ∈F def= (F1∩Fη)\(F2∪F3).

Clearly, logdens(F)1−η. Passing to a subsequence of n}, if needed, we may assume that limn→∞θn=θ0.

Further we can assume that the sequence {rn} is increasing and

logrn+1 ≤rd/2n , (2.11)

where d is also from Definition 2. If not, then there exists a sequence {rn} with rn → ∞ such that [rn,exp(rd/2n )]∩F =. Therefore,

logdens(R\F) lim

n→∞

exp(rd/2n ) rn

dt t

log exp(rnd/2) = 1.

This is impossible since

1 = logdens(R)logdens(F) + logdens(R\F)>2−η.

Summarizing what we have, we obtain the following claim.

(10)

Claim 2. There are l0 >0 and a sequence zn=rnen such that

|f(zn)|=M(rn, f), and lim

n→∞θn=θ0,

and further, (2.7)-(2.11) holds for allrn and any θ n−l0, θn+l0).

In the following, we consider two cases on whether A has the growth type (A1) in someθ∈0−l0/2, θ0+l0/2)\F or not, where F contains finitely many points.

For simplicity, we denoteδ(θ) =δA(θ) in the following.

Case 1.First assume that there is noθ 0−l0/2, θ0+l0/2)\F such that (A1) holds in this direction. We can take ˜θ1 0−l0/2, θ0) and ˜θ2 0, θ0+l0/2) such that

max0−θ˜1˜2−θ0} ≤ π 2(ρ(A) + 1). It follows from (A2) that

|A(reiθ˜j)| ≤Mjrτ, (j = 1,2).

Taking suitable branch ofzτ makes sure thatA(z)zτ is analytic in Ω(r0; ˜θ1˜2) = {z : argz θ1˜2), |z| ≥r0}. It is easy to see that for 0< δ <1, we have

|zτA(z)| ≤exp{|z|κ2δ}, where κ2 =π/(˜θ2−θ˜1)

in Ω(r0; ˜θ1˜2). Applying Phragm´en-Lindel¨of theorem tozτA(z), we conclude that

|zτA(z)| ≤M˜ := max{M1, M2, r0τM(r0, A)}, forz Ω(r0; ˜θ1˜2), and so we also have

|A(z)| ≤M r˜ τ, for z Ω(r0; ˜θ1˜2) (2.12) where Ω(r0; ˜θ1˜2) denotes the closure of Ω(r0; ˜θ1˜2).

Applying (2.7) in (2.6), we obtain that for large enough n, 1

2

ν(rn, f) rn

2

≤ |A(rnen)|+ |H(zn)| M(rn, f). Then by (2.7) and (2.12), it is easy to see

ν(rn, f)2 2 ˜M r2+τn + 2rn2|H(zn)|

M(rn, f) 4 ˜M rn2+τ,

which impliesμ(f)≤1 +τ /2< μ(A). This contradicts the fact that μ(f)≥μ(A) in Claim 1.

Case 2.Assume that there is an θ 0 −l0/2, θ0 +l0/2)\ F such that (A1) holds in this direction. This also means thatδ(θ)>0. Without loss of generality, we assume thatθ 0−l0/2, θ0]. We will construct another sequence of points from {zn} by taking zn = rnen with θn = θn0 −θ). Here, we can choose l0 < min{|θ −θ|, θ F}, so θn F for sufficiently large n. It follows from the continuity ofδ(θ) outside F that for large enough n,

1

2δ(θ)≤δ(θn) 3

2δ(θ). (2.13)

(11)

At the same time, from (2.7) and (2.10), it is obvious that H(zn)

f(zn)

≤e |H(zn)|

M(rn, f)1s 0, (2.14) asn → ∞. Then inserting (A1), (2.8) and (2.14) into (2.6) implies

exp1

2c1δ(θ)rnd

≤ |A(rnen)| ≤f(zn) f(zn)

+|H(zn)|

|f(zn)| 2B(T(2rn, f))4. Consequently,

1

2c1δ(θ0)rnd 4 logT(2rn, f) +O(1), as n → ∞. Thus, for any r [rn, rn+1],

logT(2r, f)

log(2r) logT(2rn, f) log(2rn+1)

1

6c1δ(θ0)rnd+O(1) log 2 +rnd/2

→ ∞. Thusμ(f) = is proved. Therefore the proof of Theorem 1 is completed.

Proof of Theorem 2. We know from [13, p.77] that

4A= c

E 2

−E E

2

+ 2E

E , (2.15)

where c is the Wronskian determinant of f1 and f1, thus c is a constant. Clearly, E(z) must be transcendental entire, so for any given constant s [0,1)

|c|

M(r, E)1s 0, (2.16)

asr→ ∞. Applying Wiman-Valiron theory as in (2.9), we obtain 2E

E −E E

2

=

ν(r, E) z

2

(1 +o(1)), (2.17)

for z satisfying |f(z)| = M(r, f) and |z| = r outside an exceptional set of finite logarithmic measure. On the other hand, again by Gundersen’s estimate [8], we have −E(z)

E(z) 2

+ 2E(z) E(z)

≤E(z) E(z)

2+ 2E(z) E(z)

≤C(T(2r, f))4, (2.18) for |z|=routside another exceptional set of finite logarithmic measure, and where C is a constant. Then, similar to the proof of Theorem 1, we can conclude that μ(E) =∞.

3. Baker wandering domain of the solutions

In this section, we will use Nevanlinna theory in angular domains along with our results in Section 2 to prove Theorem 3. For the convenience of the reader, we recall some basic definitions here (e.g. see [7, 23]).

(12)

Let g(z) be meromorphic in the closure of Ω(α, β) def= {z C : argz (α, β)} where β−α∈(0,2π]. Define

Aα, β(r, g) = ω π

r 1

1 tω tω

r

log+|g(te)|+ log+|g(te)|dt t , Bα, β(r, g) = 2ω

πrω β

α

log+|g(re)|sinω(θ−α)dθ, Cα,β(r, g) = 2

1<|bν|<r

1

|bν|ω |bν|ω r

sinω(βν −α),

whereω =π/(β−α), andbν =|bν|eν are the poles of g in the closure of Ω(α, β) appearing according to their multiplicities. The Nevanlinna’s angular characteristic of g is defined by

Sα, β(r, g) = Aα, β(r, g) +Bα, β(r, g) +Cα,β(r, g). (3.1) Lemma 6. ([23, Theorem 2.5.1]) Let f(z) be a meromorphic function in Ω(α ε, β+ε) for ε >0 and 0< α < β <2π. Then there exists a constant K >0 such that

Aα, β

r, f f

+Bα, β

r,f f

≤K

log+Sαε, β+ε(r, f) + logr+ 1 , for r >1, except for a set with finite linear measure.

Lemma 7. ([22, Corollary 1]) Let f(z) be a transcendental meromorphic function with at most finitely many poles. If J(f) has only bounded components, then for any complex number a, there exists a constant 0< d < 1 and two sequences {rn} and {Rn} of positive numbers with rn → ∞ and Rn/rn → ∞(n → ∞) such that

M(r, a, f)d≤L(r, a, f), r ∈G,

where M(r, a, f) = max{|f(z)|:|z−a|=r}, L(r, a, f) = min{|f(z)|:|z−a|=r}, and G=n=1{r:rn< r < Rn} which has an infinite logarithmic measure.

Lemma 8. ([10, Satz 1]) Let pj(x)(j = 1,2,· · · , n) and f(x) be complex-valued functions on the interval [a, b). Suppose that Pj(x)(j = 1,2,· · · , n) and F(x) are continuous, non-negative functions defined on [a, b) such that |pj(x)| ≤ Pj(x) and

|f(x)| ≤F(x). Let v(x) satisfy the differential equation v(n)

n j=1

pj(x)v(nj) =f(x), and V(x) satisfy the differential equation

V(n) n

j=1

Pj(x)V(nj) =F(x).

Moreover, we assume that |v(k)(a)| ≤ V(k)(a) for k = 0,1,· · · , n 1. Then

|v(k)(x)| ≤V(k)(x) holds for x∈[a, b).

(13)

Proof of Theorem 3. We now assume that g = f(n)(n Z) has a Baker wandering domain, and complete the proof by reduction to absurdity. From [22, Remark (A)], the Julia set of a transcendental meromorphic function with at most finitely many poles has only bounded components if and only if it has a Baker wandering domain. Since g is transcendental entire, so J(g) only has bounded components. Now, by Lemma 7, there exists 0< d <1 such that

|g(z)| ≥M(r, g)d, r∈G0, (3.1) where G0 is a set of infinite logarithmic measure.

Since there is oneθ [0,2π)\F withδ(θ) =δA(θ)<0, then by the continuity of δ(θ) outside the exceptional setF, we can findθ1 < θ2 such thatδ(θj)<0(j = 1,2) and furthermore,θ2−θ1 ≤π/(ρ(A)+1). Thus similarly to the the proof of Theorem 1, we have

|A(z)| ≤M r˜ τ, for z Ω(r0;θ1, θ2), (3.2) where ˜M > 1 is a constant.

We first consider f(n)(z) when n 0. It follows from (1.2) that g(z) =f(n)(z) satisfies the equation

g(m)+A(z)g(m2) =H(z), (3.3) where m = −n + 2. For any fixed θ 1, θ2], we set h(r) = g(re). Then, h(k)(r) = eikθg(k)(re) fork N. Substituting this into (3.3) yields

h(m)+A(re)ei2θh(m2) =eimθH(re). (3.4) By a simple computation for a positive integer l 2, we have

exp(rl)(k)

=pk(l1)(r) exp(rl), k N, (3.5) wherepk(l1)(r) is a polynomial of degreek(l−1) in r with the highest order term aslkrk(l1). Set

M0 def= max{M(r0, g(k)), k = 0,1,· · · , m−1}, F(r)def= M0

pm(l1)(r)−M p˜ (m2)(l1)(r)r[τ]+1

exp(rl).

By the definition of order, there exists sufficiently larger0 such that for r≥r0,

|emiθH(re)| ≤exp(rl)≤F(r),

where we takel = [ρ(H)] + [τ] + 2, and [x] is the integer part of x≥0. Define

V(r) =M0exp(rl), (3.6)

and it is easy to check that

|h(k)(r0)| ≤M0 ≤M0exp((r0)l)≤V(r0), (3.7) for k = 0,1,· · · , m−1, and that V(r) defined in (3.6) satisfies the equation

V(m)−M r˜ [τ]+1V(m2) =F(r). (3.8)

Then, applying Lemma 8 to equations (3.4) and (3.8), we get

|g(re)|=|h(r)| ≤V(r) =M0exp(rl), (3.9)

(14)

for all z = re Ω(r0;θ1, θ2), and if needed, we can take θ2 a little smaller than before.

Specially for n = 0, combining the estimate (3.9) with (3.1) immediately yields M(r, f)d≤M0exp(rl),

which impliesμ(f)≤l, a contradiction.

By using (3.9), we obtain

Aθ1, θ2(r, g) = O(rlω), Bθ1, θ2(r, g) =O(rlω)

for ω = π/(θ2 −θ1), and for all r r0. Since f(z) is entire, then Cθ12(r, g) = 0.

Summarizing the above argument, so for n≤0, we have

Sθ1, θ2(r, f(n)) =Aθ12(r, f(n)) +Bθ12(r, f(n)) =O(rlω). (3.10) Now we consider f(n)(z) when n >0. For any ε >0, Lemma 6 gives

Sθ1+ε, θ2ε

r,f(n) f

n1

j=0

Sθ1+ε, θ2ε

r,f(j+1) f(j)

≤K n1

j=0

log+Sθ1, θ2(r, f(j)) + logr+ 1

, (3.11)

with an exceptional set F1 of finite linear measure. Whenn= 1, (3.10) and (3.11) imply

Sθ1+ε,θ2ε

r,f f

≤K

log+Sθ12(r, f) + logr+ 1

=O(logr), forr ∈F1, which leads to

Sθ1+ε,θ2ε(r, f)≤Sθ1+ε,θ2ε

r,f f

+Sθ1+ε,θ2ε(r, f) =O(rlω), r∈F1. Thus, by induction, we can conclude that forr ∈F1 and n >0,

Sθ1+ε, θ2ε

r,f(n) f

=O(logr), Sθ1+ε, θ2ε(r, f(n)) = O(rlω). (3.12) On account of this fact and (3.1), we obtain

Sα, β(r, g)≥Bα, β(r, g) = 2ω πrω

β α

log+|g(re)|sinω(θ−α)dθ

2

πrωdlog+M(r, g) β

α

sinω(θ−α)dω(θ−α)

= 4d

πrω logM(r, g), r ∈G0\F1

(15)

whereα=θ1+ε, β =θ2−ε, andω =π/(θ2−θ12ε) forn >0, whileα=θ1, β =θ2, and ω = π/(θ2 θ1) for n 0. Substituting (3.10) and (3.12) into the above inequality yields

logM(r, g) rω

4dSα, β(r, g) = rω

4dSα, β(r, f(n)) = O(rl), r∈G0\F1,

which implies μ(g) < . Therefore, μ(f) < since μ(g) = μ(f(n)) = μ(f) for any n Z (e.g. see [20, Theorem 4.2]). However, Theorem 1 gives μ(f) = . Therefore, f(n) has no Baker wandering domains for any n∈Z.

4. Baker wandering domain of E(z)

First we consider E(n) for any non-negative integer n. Following from Leibniz formula for higher-order derivatives, we have

E(n)= n k=0

Cnkf1(nk)f2(k), where Cnk = n!

k!(n−k)!. (4.1) Recall that (1.1) is a special case of (1.2) with H(z) 0. From the proof of Theorem 3., there exists one smaller angular domain Ω(α, β)Ω(θ1, θ2) such that

Sα,β(r, f(k)) =O(rlω), for k Z,

for r outside a set of finite logarithmic measure. Combining this with (4.1) leads to

Sα,β(r, E(n)) =O(r2(lω)),

for ω = π/(β −α), and for r outside a set of finite linear measure. Similar to the proof of Theorem 3, ifE(n) has a Baker wandering domain, then μ(E(n))<∞ which contradicts μ(E(n)) =μ(E) =∞ by Theorem 2. Therefore, for any integer n≥0,E(n) has no Baker wandering domains.

Now we consider g(z) = E(n) for n <0. We know from [13, p. 77] that E+ 4A(z)E+ 2A(z)E = 0,

which implies thatg satisfies the differential equation

g(m)+ 4A(z)g(m2)+ 2A(z)g(m3) = 0, (4.2) where m = −n + 3. Similarly to the proof of Theorem 3, we have (3.2) for z Ω(r0;θ1, θ2). By the estimate for logarithmic derivatives in [8] again, it follows that we can find another two ˜θ1˜1 1, θ2)\E1 such that for sufficiently large r,

|A(reiθ˜j)|=A(reiθ˜j) A(reiθ˜j)

|A(reiθ˜j)| ≤rρ(A)+τ, (4.3) where the set E1 [0,2π) has linear measure zero, and j = 1,2. Since ˜θ1 and ˜θ2 are chosen to be very close to each other, so by the Phragm´en-Lindel¨of principle, there exist two positive constants r0, M1 such that

|A(z)| ≤M1rρ(A)+τ ≤M1rN, for z Ω(r0˜1˜2), (4.4) whereN = [ρ(A) +τ] + 1.

(16)

For any fixed θ θ1˜2], we still take h(r) = g(re), so

h(m)+ 4A(re)e2iθh(m2)+ 2A(re)e3iθh(m3) = 0. (4.6) Thus, we have from (3.2) and (4.4) that

|A(re)e2iθ| ≤M r˜ [τ]+1, |A(re)e3iθ| ≤M1rN. (4.7) Recalling (3.5), set M0 = max{|g(k)(0)|, k = 0,1,· · · , m−1.}, and

F(r) =M0

pmN(r)4 ˜M r[τ]+1p(m2)N(r)2M1rNp(m3)N(r)

exp(rN+1), we easily see that

V(r) =M0exp(rN+1), satisfies the differential equation

V(m)4 ˜M r[τ]+1V(m2)2M1rNV(m3) =F(r).

Moreover, there exists a r0 large enough such that 0 < F(r) and |g(k)(0)| ≤ V(k)(0), k = 0,1,· · · , m−1.Applying Lemma 8 again leads to

|g(re)|=|h(r)| ≤V(r) = M0exp(rN+1),

for all z Ω(r0˜1˜2). Hence, log+|E(n)(z)|=O(rl) in Ω(˜θ1˜2), which implies Sθ1, θ2(r, E(n)) = O(rl).

IfE(n)has a Baker wandering domain, similar to the proof of Theorem 3, we obtain μ(E(n))<∞, which contradicts Theorem 2. Thus, for any negative integern,E(n) also has no Baker wandering domains.

Therefore, E(n) has no Baker wandering domains for every n∈Z, and Theorem 4 is completely proved.

Acknowledgment: The first author was supported in part by the Academy of Finland (Grant No. 286877). The second author was supported by the National Natural Science Foundation of China (Grant No. 11771090) and Natural Sciences Foundation of Shanghai (Grant No. 17ZR1402900).

References

[1] I. N. Baker, Multiply connected domains of normality in iteration theory,Math.Z., 81(1963), 206-214.

[2] I. N. Baker, An entire function which has wandering domains,J. Aust. Math. Soc., Ser. A, 22(1976), 173-176.

[3] I. N. Baker, Wandering domain in the iteration of entire functions, Proc. Lond. Math.

Soc.,(3)49(1984), 563-576.

[4] Carlos A. Berestein and Roger Gay, Complex Analysis and Special Topics in Harmonics Analysis, New York, Springer-Verlag, 1995.

[5] W. Bergweiler, Iteration of meromorphic functions,Bull. Amer. Math. Soc., 29 (1993), 151- 188.

[6] W. Fuchs, Proof of conjecture of G.P´olya concerning gap series, Illinois J. Math, 7(1963), 661-667.

[7] A. A. Goldberg, I. V. Ostrovskii, Value distribution of meromorphic functions, AMS Trans- lations of Mathematical Monographs series, 2008.

Viittaukset

LIITTYVÄT TIEDOSTOT

In this section, we experiment with a set of RIRs to explore the im- provements of modeling source directivity and reflection coefficients over R-DSB. In addition, the order

Finally we show the existence of weak solutions to mean-field stochastic differential equa- tions under bounded, measurable and continuous coefficients by show- ing that there exists

Keywords: Deficient values, entire functions, Frei’s theorem, linear difference equations, linear differential equations, Wittich’s theorem.. 2010 MSC: Primary 34M10;

The author mainly deals with the value distribution of meromorphic solutions of complex functional difference equations in the complex plane, investigates the existence of finite

Keywords: Asymptotic growth, asymptotic integration, deficient values, Frei’s theorem, growth of solutions, linear differential equation, Liouville’s transformation, nonlinear

Keywords: ϕ-order, deficient values, difference operators, Frei’s theorem, growth of solu- tions, linear difference equations, linear differential equations, logarithmic

Korhonen et al., Lower order and Baker wandering domains of solutions to differential equations with coefficients of exponential growth, J.. As a service to our customers we

The aim of this paper is to consider certain conditions on the coef- ficient A of the differential equation f 00 + Af = 0 in the unit disc, which place all normal solutions f to