Abstract

Let 𝑋 and 𝐾 be compact plane sets with 𝐾𝑋. We define 𝐴(𝑋,𝐾)={𝑓𝐶(𝑋)𝑓|𝐾𝐴(𝐾)}, where 𝐴(𝐾)={𝑔𝐶(𝑋)𝑔 is analytic on int(𝐾)}. For 𝛼(0,1], we define Lip(𝑋,𝐾,𝛼)={𝑓𝐶(𝑋)𝑝𝛼,𝐾(𝑓)=sup{|𝑓(𝑧)𝑓(𝑤)|/|𝑧𝑤|𝛼𝑧,𝑤𝐾,𝑧𝑤}<} and Lip𝐴(𝑋,𝐾,𝛼)=𝐴(𝑋,𝐾)Lip(𝑋,𝐾,𝛼). It is known that Lip𝐴(𝑋,𝐾,𝛼) is a natural Banach function algebra on 𝑋 under the norm ||𝑓||Lip(𝑋,𝐾,𝛼)=||𝑓||𝑋+𝑝𝛼,𝐾(𝑓), where ||𝑓||𝑋=sup{|𝑓(𝑥)|𝑥𝑋}. These algebras are called extended analytic Lipschitz algebras. In this paper we study unital homomorphisms from natural Banach function subalgebras of Lip𝐴(𝑋1,𝐾1,𝛼1) to natural Banach function subalgebras of Lip𝐴(𝑋2,𝐾2,𝛼2) and investigate necessary and sufficient conditions for which these homomorphisms are compact. We also determine the spectrum of unital compact endomorphisms of Lip𝐴(𝑋,𝐾,𝛼).

1. Introduction and Preliminaries

We let , 𝔻={𝑧|𝑧|<1}, 𝔻={𝑧|𝑧|1}, 𝔻(𝜆,𝑟)={𝑧|𝑧𝜆|<𝑟}, and 𝔻(𝜆,𝑟)={𝑧|𝑧𝜆|𝑟} denote the field of complex numbers, the open unit disc, the closed unit disc, and the open and closed discs with center at 𝜆 and radius 𝑟, respectively. We also denote 𝔻(0,𝑟) by 𝔻𝑟.

Let 𝐴 and 𝐵 be unital commutative semisimple Banach algebras with maximal ideal spaces (𝐴) and (𝐵). A homomorphism 𝑇𝐴𝐵 is called unital if 𝑇1𝐴=1𝐵. If 𝑇 is a unital homomorphism from 𝐴 into 𝐵, then 𝑇 is continuous and there exists a norm-continuous map 𝜑(𝐵)(𝐴) such that 𝑇𝑓=𝑓𝜑 for all 𝑓𝐴, where ̂𝑔 is the Gelfand transform 𝑔. In fact, 𝜑 is equal the adjoint of 𝑇𝐵𝐴 restricted to (𝐵). Note that 𝑇 is a weak*-weak* continuous map from 𝐵 into 𝐴. Thus 𝜑 is a continuous map from (𝐵) with the Gelfand topology into (𝐴) with the Gelfand topology.

Let 𝐴 be a unital commutative semisimple Banach algebra, and let 𝑇 be an endomorphism of 𝐴, a homomorphism from 𝐴 into 𝐴. We denote the spectrum of 𝑇 by 𝜎(𝑇) and define𝜎(𝑇)={𝜆𝜆𝐼𝑇isnotinvertible}.(1.1)

For a compact Hausdorff space 𝑋, we denote by 𝐶(𝑋) the Banach algebra of all continuous complex-valued functions on 𝑋.

Definition 1.1. Let 𝑋 be a compact Hausdorff space. A Banach function algebra on 𝑋 is a subalgebra 𝐴 of 𝐶(𝑋) which contains 1𝑋, the constant function 1 on 𝑋, separates the points of 𝑋, and is a unital Banach algebra with an algebra norm . If the norm of a Banach function algebra on 𝑋 is 𝑋, the uniform norm on 𝑋, it is called a uniform algebra on 𝑋.

Let 𝐴 and 𝐵 be Banach function algebras on 𝑋 and 𝑌, respectively. If 𝜑𝑌𝑋 is a continuous mapping such that 𝑓𝜑𝐵 for all 𝑓𝐴 and if 𝑇𝐴𝐵 is defined by 𝑇𝑓=𝑓𝜑, then 𝑇 is a unital homomorphism, which is called the induced homomorphism from 𝐴 into 𝐵 by 𝜑. In particular, if 𝑌=𝑋 and 𝐵=𝐴, then 𝑇 is called the induced endomorphism of 𝐴 by the self-map 𝜑 of 𝑋.

Let 𝐴 be a Banach function algebra on a compact Hausdorff space 𝑋. For 𝑥𝑋, the map 𝑒𝑥𝐴, defined by 𝑒𝑥(𝑓)=𝑓(𝑥), is an element of (𝐴) and is called the evaluation homomorphism on 𝐴 at 𝑥. This fact implies that 𝐴 is semisimple and 𝑓𝑋𝑓(𝐴) for all 𝑓𝐴. Note that the map 𝑥𝑒𝑥𝑋(𝐴) is a continuous one-to-one mapping. If this map is onto, we say that 𝐴 is natural.

Proposition 1.2. Let 𝑋 and 𝑌 be compact Hausdorff spaces, and let 𝐴 and 𝐵 be natural Banach function algebras on 𝑋 and 𝑌, respectively. Then every unital homomorphism 𝑇𝐴𝐵 is induced by a unique continuous map 𝜑𝑌𝑋. In particular, if 𝑋 is a compact plane set and the coordinate function 𝑍 belongs to 𝐴, then 𝜑=𝑇𝑍 and so 𝜑𝐵.

Proof. Let 𝑇𝐴𝐵 be a unital homomorphism. Since 𝐴 and 𝐵 are unital commutative semisimple Banach algebras, there exists a continuous map 𝜓(𝐵)(𝐴) such that 𝑇𝑓=𝑓𝜓 for all 𝑓𝐴. The naturality of the Banach function algebra 𝐴 on 𝑋 implies that the map 𝐽𝐴𝑋(𝐴), defined by 𝐽𝐴(𝑥)=𝑒𝑥, is a homeomorphism and so 𝐽𝐴1(𝐴)𝑋 is continuous. Since 𝐵 is a Banach function algebra on 𝑌, the map 𝐽𝐵𝑌(𝐵), defined by 𝐽𝐵(𝑦)=𝑒𝑦, is continuous. We now define the map 𝜑𝑌𝑋 by 𝜑=𝐽𝐴1𝜓𝐽𝐵. Clearly, 𝜑 is continuous. Let 𝑓𝐴. Since 𝑒(𝑇𝑓)(𝑦)=𝑇𝑓𝑦=𝐽𝑓𝜓𝐵=(𝑦)𝑓𝐽𝐴=𝑓𝑒(𝜑(𝑦))𝜑(𝑦)=𝑒𝜑(𝑦)(=𝑓)=𝑓(𝜑(𝑦))(𝑓𝜑)(𝑦),(1.2) for all 𝑦𝑌, we have 𝑇𝑓=𝑓𝜑. Therefore, 𝑇 is induced by 𝜑.
Now, let 𝑋 be a compact plane set, and let 𝑍𝐴. Then 𝜑=𝑍𝜑=𝑇𝑍, and so 𝜑𝐵.

Corollary 1.3. Let 𝑋 be a compact Hausdorff space, and let 𝐴 be a natural Banach function algebra on 𝑋. Then every unital endomorphism 𝑇 of 𝐴 is induced by a unique continuous self-map 𝜑 of 𝑋. In particular, if 𝑋 is a compact plane set and 𝐴 contains the coordinate function 𝑍, then 𝜑=𝑇𝑍 and so 𝜑𝐴.

Definition 1.4. Let 𝑋 be a compact plane set which is connected by rectifiable arcs, and let 𝛿(𝑧,𝑤) be the geodesic metric on 𝑋, the infimum of the length of the arcs joining 𝑧 and 𝑤. 𝑋 is called uniformly regular if there exists a constant 𝐶 such that, for all 𝑧,𝑤𝑋, 𝛿(𝑧,𝑤)𝐶|𝑧𝑤|.

The following lemma occurs in [1] but it is important and we will be using it in the sequel.

Lemma 1.5 (see [1, Lemma 1.5]). Let 𝐻 and 𝐾 be two compact plane sets with 𝐻int(𝐾). Then there exists a finite union of uniformly regular sets in int(𝐾) containing 𝐻, namely 𝑌, and then a positive constant 𝐶 such that for every analytic complex-valued function 𝑓 on int(𝐾) and any 𝑧,𝑤𝐻, ||𝑓||(𝑧)𝑓(𝑤)𝐶|𝑧𝑤|𝑓𝑌+𝑓𝑌.(1.3)

Let 𝑋 be a compact plane set. We denote by 𝐴(𝑋) the algebra of all continuous complex-valued functions on 𝑋 which are analytic on int(𝑋), the interior of 𝑋, and call it the analytic uniform algebra on 𝑋. It is known that 𝐴(𝑋) is a natural uniform algebra on 𝑋.

Let 𝑋 and 𝐾 be compact plane sets such that 𝐾𝑋. We define 𝐴(𝑋,𝐾)={𝑓𝐶(𝑋)𝑓|𝐾𝐴(𝐾)}. Clearly, 𝐴(𝑋,𝐾)=𝐴(𝑋) if 𝐾=𝑋, and 𝐴(𝑋,𝐾)=𝐶(𝑋) if int(𝐾) is empty. We know that 𝐴(𝑋,𝐾) is a natural uniform algebra on 𝑋 (see [2]) and call it the extended analytic uniform algebra on 𝑋 with respect to 𝐾.

Let (𝑋,𝑑) be a compact metric space. For 𝛼(0,1], we denote by Lip(𝑋,𝛼) the algebra of all complex-valued functions 𝑓 for which 𝑝𝛼,𝑋(𝑓)=sup{|𝑓(𝑧)𝑓(𝑤)|/𝑑𝛼(𝑧,𝑤)𝑧,𝑤𝑋,𝑧𝑤}<. For 𝑓Lip(𝑋,𝛼), we define the 𝛼-Lipschitz norm 𝑓 by 𝑓Lip(𝑋,𝛼)=𝑓𝑋+𝑝𝛼,𝑋(𝑓). Then (Lip(𝑋,𝛼),Lip(𝑋,𝛼)) is a unital commutative Banach algebra. For 𝛼(0,1), we denote by lip(𝑋,𝛼) the algebra of all complex-valued functions 𝑓 on 𝑋 for which |𝑓(𝑧)𝑓(𝑤)|/𝑑𝛼(𝑧,𝑤)0 as 𝑑(𝑧,𝑤)0. Then lip(𝑋,𝛼) is a unital closed subalgebra of Lip(𝑋,𝛼). These algebras are called Lipschitz algebras of order 𝛼 and were first studied by Sherbert in [3, 4]. We know that the Lipschitz algebras Lip(𝑋,𝛼) and lip(𝑋,𝛼) are natural Banach function algebras on 𝑋.

Let (𝑋,𝑑) be a compact metric space, and let 𝐾 be a compact subset of 𝑋. For 𝛼(0,1], we denote by Lip(𝑋,𝐾,𝛼) the algebra of all complex-valued functions 𝑓 on 𝑋 for which 𝑝𝛼,𝐾(𝑓)=sup{|𝑓(𝑧)𝑓(𝑤)|/𝑑𝛼(𝑧,𝑤)𝑧,𝑤𝐾,𝑧𝑤}<. In fact, Lip(𝑋,𝐾,𝛼)={𝑓𝐶(𝑋)𝑓|𝐾Lip(𝐾,𝛼)}. For 𝑓Lip(𝑋,𝐾,𝛼), we define 𝑓Lip(𝑋,𝐾,𝛼)=𝑓𝑋+𝑝𝛼,𝐾(𝑓). Then Lip(𝑋,𝐾,𝛼) under the algebra norm Lip(𝑋,𝐾,𝛼) is a unital commutative Banach algebra. Moreover, Lip(𝑋,𝛼) is a subalgebra of Lip(𝑋,𝐾,𝛼); Lip(𝑋,𝐾,𝛼)=Lip(𝑋,𝛼) if 𝑋𝐾 is finite, and Lip(𝑋,𝐾,𝛼)=𝐶(𝑋) if 𝐾 is finite. For 𝛼(0,1), we denote by lip(𝑋,𝐾,𝛼) the algebra of all complex-valued functions 𝑓 on 𝑋 for which |𝑓(𝑧)𝑓(𝑤)|/𝑑𝛼(𝑧,𝑤)0 as 𝑑(𝑧,𝑤)0 with 𝑧,𝑤𝐾. In fact, lip(𝑋,𝐾,𝛼)={𝑓𝐶(𝑋)𝑓|𝐾lip(𝐾,𝛼)}. Clearly, lip(𝑋,𝐾,𝛼) is a closed unital subalgebra of Lip(𝑋,𝐾,𝛼). Moreover, lip(𝑋,𝛼) is a subalgebra of lip(𝑋,𝐾,𝛼); lip(𝑋,𝐾,𝛼)=lip(𝑋,𝛼) if 𝑋𝐾 is finite, and lip(𝑋,𝐾,𝛼)=𝐶(𝑋) if 𝐾 is finite. The Banach algebras Lip(𝑋,𝐾,𝛼) and lip(𝑋,𝐾,𝛼) are Banach function algebras on 𝑋 and were first introduced by Honary and Moradi in [5].

Let 𝑋 be a compact plane set. We define Lip𝐴(𝑋,𝛼)=Lip(𝑋,𝛼)𝐴(𝑋) for 𝛼(0,1] and lip𝐴(𝑋,𝛼)=Lip(𝑋,𝛼)𝐴(𝑋) for 𝛼(0,1). These algebras are called analytic Lipschitz algebras. We know that analytic Lipschitz algebras Lip𝐴(𝑋,𝛼) and lip𝐴(𝑋,𝛼) under the norm Lip(𝑋,𝛼) are natural Banach function algebras on 𝑋 (see [6]).

Let 𝑋 and 𝐾 be compact plane sets with 𝐾𝑋. We define Lip𝐴(𝑋,𝐾,𝛼)=Lip(𝑋,𝐾,𝛼)𝐴(𝑋,𝐾) for 𝛼(0,1] and lip𝐴(𝑋,𝐾,𝛼)=lip(𝑋,𝐾,𝛼)𝐴(𝑋,𝐾) for 𝛼(0,1). Then Lip𝐴(𝑋,𝐾,𝛼) and lip(𝑋,𝐾,𝛼) are closed unital subalgebras of Lip(𝑋,𝐾,𝛼) and lip(𝑋,𝐾,𝛼) under the norm Lip(𝑋,𝐾,𝛼), respectively. Moreover, Lip𝐴(𝑋,𝐾,𝛼)=Lip𝐴(𝑋,𝛼)[lip𝐴(𝑋,𝐾,𝛼)=Lip𝐴(𝑋,𝛼)] if 𝐾=𝑋, and Lip𝐴(𝑋,𝐾,𝛼)=Lip(𝑋,𝐾,𝛼)[lip𝐴(𝑋,𝐾,𝛼)=lip(𝑋,𝐾,𝛼)] if int(𝐾) is empty.

The algebras Lip𝐴(𝑋,𝐾,𝛼) and lip(𝑋,𝐾,𝛼) are called extended analytic Lipschitz algebras and were first studied by Honary and Moradi in [5]. They showed that the extended analytic Lipschitz algebras Lip𝐴(𝑋,𝐾,𝛼) and lip𝐴(𝑋,𝐾,𝛼) under the norm Lip(𝑋,𝐾,𝛼) are natural Banach function algebras on 𝑋 [5, Theorem 2.4].

Behrouzi and Mahyar in [1] studied endomorphisms of some uniform subalgebras of 𝐴(𝑋) and some Banach function subalgebras of Lip𝐴(𝑋,𝛼) and investigated some necessary and sufficient conditions for these endomorphisms to be compact, where 𝑋 is a compact plane set and 𝛼(0,1].

In Section 2, we study unital homomorphisms from natural Banach function subalgebras of Lip𝐴(𝑋1,𝐾1,𝛼1) to natural Banach function subalgebras of Lip𝐴(𝑋2,𝐾2,𝛼2) and investigate necessary and sufficient conditions for which these homomorphisms are compact. In Section 3, we determine the spectrum of unital compact endomorphisms of Lip𝐴(𝑋,𝐾,𝛼).

2. Unital Compact Homomorphisms

We first give a sufficient condition for which a continuous map 𝜑𝑋2𝑋1 induces a unital homomorphism 𝑇 from a subalgebra 𝐵1 of 𝐴(𝑋1,𝐾1) into a subalgebra 𝐵2 of 𝐴(𝑋2,𝐾2).

Proposition 2.1. Let 𝑋𝑗 and 𝐾𝑗 be compact plane sets with int(𝐾𝑗) and 𝐾𝑗𝑋𝑗, and let 𝐵𝑗 be a subalgebra of 𝐴(𝑋𝑗,𝐾𝑗) which is a natural Banach function algebra on 𝑋𝑗 under an algebra norm 𝑗, where 𝑗{1,2}. If 𝜑𝐵2 with 𝜑(𝑋2)int(𝐾1), then 𝜑 induces a unital homomorphism 𝑇𝐵1𝐵2. Moreover, if 𝑍𝐵1, then 𝜑=𝑇𝑍.

Proof. The naturality of Banach function algebra 𝐵2 on 𝑋2 implies that 𝜎𝐵2()=(𝑋2), where 𝜎𝐴() is the spectrum of 𝐴 in the Banach algebra 𝐴. Let 𝑓𝐵1. Since 𝜑𝐵2, 𝜑(𝑋2)int(𝐾1), and 𝑓 is analytic on int(𝐾1), we conclude that 𝑓 is analytic on an open neighborhood of 𝜎𝐵2(𝜑). By using the Functional Calculus Theorem [2, Theorem  5.1 in Chapter I], there exists 𝑔𝐵2 such that ̂𝑔=𝑓𝜑 on (𝐵2). It follows that 𝑔(𝑧)=𝑒𝑧𝑒(𝑔)=̂𝑔𝑧𝑒=𝑓𝜑𝑧𝑒=𝑓𝑧(𝜑)=𝑓(𝜑(𝑧))=(𝑓𝜑)(𝑧),(2.1) for all 𝑧𝑋2 and so 𝑔=𝑓𝜑. Therefore, 𝑓𝜑𝐵2. This implies that the map 𝑇𝐵1𝐵2 defined by 𝑇𝑓=𝑓𝜑 is a unital homomorphism from 𝐵1 into 𝐵2, which is induced by 𝜑. Now let 𝑍𝐵1. Then 𝜑=𝑇𝑍 by Proposition 1.2.

Corollary 2.2. Let 𝑋 and 𝐾 be compact plane sets with int(𝐾) and 𝐾𝑋. Let 𝐵 be a subalgebra of 𝐴(𝑋,𝐾) which is a natural Banach function algebra on 𝑋 under an algebra norm 𝐵. If 𝜑𝐵 with 𝜑(𝑋)int(𝐾), then 𝜑 induces a unital endomorphism 𝑇 of 𝐵. Moreover, if 𝑍𝐵, then 𝜑=𝑇𝑍.

Proposition 2.3. Suppose that 𝛼𝑗(0,1], 𝑧𝑗, 0<𝑟𝑗<𝑅𝑗, 𝐺𝑗=𝔻(𝑧𝑗,𝑅𝑗), Ω𝑗=𝔻(𝑧𝑗,𝑟𝑗), 𝑋𝑗=𝐺𝑗, and 𝐾𝑗=Ω𝑗, where 𝑗{1,2}. Then for each 𝜌(𝑟1,𝑅1] there exists a continuous map 𝜑𝜌𝑋2𝑋1 with 𝜑𝜌(𝑋2)=𝔻(𝑧1,𝜌) such that 𝜑𝜌Lip𝐴(𝑋2,𝐾2,𝛼2) and 𝜑𝜌 does not induce any homomorphism from Lip𝐴(𝑋1,K1,𝛼1) to Lip𝐴(𝑋2,𝐾2,𝛼2).

Proof. Let 𝜌(𝑟1,𝑅1]. We define the map 𝜑𝜌𝑋2𝑋1 by 𝜑𝜌𝑧(𝑧)=1+𝜌𝑧𝑧2𝑟2||𝑧𝑧2||𝑟2,𝑧1+𝜌𝑧𝑧1||𝑧𝑧2||𝑟2<||𝑧𝑧2||𝑅2.(2.2) Clearly, 𝜑𝜌 is a continuous mapping, 𝜑𝜌(𝑋2)=𝔻(𝑧1,𝜌), and 𝜑𝜌Lip𝐴(𝑋2,𝐾2,𝛼2). We now define the function 𝑓𝜌𝑋1 by 𝑓𝜌𝜌(𝑧)=𝑧𝑧1𝑟1||𝑧𝑧1||𝑟1,𝜌𝑧𝑧1||𝑧𝑧1||𝑟1<||𝑧𝑧1||𝑅1.(2.3) Then, 𝑓𝜌Lip𝐴(𝑋1,𝐾1,𝛼1). Since 0<𝑟1𝑟2/𝜌<𝑟2 and 𝑓𝜌𝜑𝜌𝜌(𝑧)=2𝑟1𝑟2𝑧𝑧2||𝑧𝑧2||𝑟1𝑟2𝜌,𝜌𝑧𝑧2||𝑧𝑧2||𝑟1𝑟2𝜌<||𝑧𝑧2||𝑅2,(2.4) we conclude that 𝑓𝜌𝜑𝜌Lip𝐴(𝑋2,𝐾2,𝛼2). Therefore, 𝜑𝜌 does not induce any homomorphism from Lip𝐴(𝑋1,𝐾1,𝛼1) to Lip𝐴(𝑋2,𝐾2,𝛼2). Hence, the proof is complete.

Corollary 2.4. Suppose that 𝛼(0,1], 𝜆, 0<𝑟<𝑅, 𝐺=𝔻(𝜆,𝑅), Ω=𝔻(𝜆,𝑟), 𝑋=𝐺, and 𝐾=Ω. Then for each 𝜌(𝑟,𝑅], there exists a continuous self-map 𝜑𝜌 of 𝑋 with 𝜑𝜌(𝑋)=𝔻(𝜆,𝜌) such that 𝜑𝜌Lip𝐴(𝑋,𝐾,𝛼) and 𝜑𝜌 does not induce any endomorphism of Lip𝐴(𝑋,𝐾,𝛼).

We now give a sufficient condition for a unital homomorphism from a subalgebra 𝐵1 of Lip𝐴(𝑋1,𝐾1,𝛼1) into a subalgebra 𝐵2 of Lip𝐴(𝑋2,𝐾2,𝛼2) to be compact.

Theorem 2.5. Suppose that 𝛼𝑗(0,1], 𝑋𝑗 and 𝐾𝑗 are compact plane sets with int(𝐾𝑗) and 𝐾𝑗𝑋𝑗, and 𝐵𝑗 is a subalgebra of Lip𝐴(𝑋𝑗,𝐾𝑗,𝛼𝑗) which is a natural Banach function algebra on 𝑋𝑗 under the norm Lip(𝑋𝑗,𝐾𝑗,𝛼𝑗), where 𝑗{1,2}. Let 𝜑𝑋2𝑋1 be a continuous mapping. If 𝜑 is constant or 𝜑𝐵2 with 𝜑(𝑋2)int(𝐾1), then 𝜑 induces a unital compact homomorphism 𝑇𝐵1𝐵2.

Proof. If 𝜑𝑋2𝑋1 is constant, then the map 𝑇𝐵1𝐵2 defined by 𝑇𝑓=𝑓𝜑 is a unital homomorphism from 𝐵1 into 𝐵2 with dim𝑇(𝐵1)1, and so it is compact.
Let 𝜑𝑋2𝑋1 be a nonconstant mapping with 𝜑𝐵2 and 𝜑(𝑋2)Ω1. Then the map 𝑇𝐵1𝐵2 defined by 𝑇𝑓=𝑓𝜑 is a unital homomorphism from 𝐵1 to 𝐵2 by Proposition 2.1. To prove the compactness of 𝑇, let {𝑓𝑛}𝑛=1 be a bounded sequence in 𝐵1 with 𝑓𝑛Lip(𝑋1,𝐾1,𝛼1)1 for all 𝑛. This implies that {𝑓𝑛|𝐾1}𝑛=1 is a bounded sequence in 𝐶(𝐾1) which is equicontinuous on (𝐾1,𝑑𝛼11). By Arzela-Ascoli’s theorem, {𝑓𝑛}𝑛=1 has a subsequence {𝑓𝑛𝑗}𝑗=1 such that {𝑓𝑛𝑗|𝐾1}𝑗=1 is convergent in 𝐶(𝐾1). Since 𝑓𝑛𝑗|𝐾1𝐴(𝐾1) for all 𝑗, {𝑓𝑛𝑗|𝐾1}𝑗=1 is convergent in 𝐴(𝐾1). By Montel's theorem, the sequences {𝑓𝑛𝑗}𝑗=1 and {𝑓𝑛𝑗}𝑗=1 are uniformly convergent on the compact subsets of int(𝐾1). Since 𝜑(𝑋2) and 𝐾1 are compact sets in the complex plane and 𝜑(𝑋2)int(𝐾1), by using Lemma 1.5, we deduce that there exists a finite union of uniformly regular sets in int(𝐾1) containing 𝜑(𝑋2), namely 𝑌, and then a positive constant 𝐶 such that for every analytic complex-valued function 𝑓 on int(𝐾1) and any 𝑧,𝑤𝜑(𝑋2)||𝑓||(𝑧)𝑓(𝑤)𝐶|𝑧𝑤|𝑓𝑌+𝑓𝑌.(2.5) Therefore, there exists a positive constant 𝐶 such that |||𝑓𝑛𝑗(𝜑(𝑧))𝑓𝑛𝑗|||||||𝑓(𝜑(𝑤))𝐶𝜑(𝑧)𝜑(𝑤)𝑛𝑗𝑌+𝑓𝑛𝑗𝑌,(2.6) for all 𝑗 and any 𝑧,𝑤𝑋2. Let 𝑗,𝑘. Then, for all 𝑧,𝑤𝐾2 with 𝜑(𝑧)𝜑(𝑤), we have |||𝑓𝑛𝑗𝑓𝜑𝑛𝑘𝑓𝜑(𝑧)𝑛𝑗𝑓𝜑𝑛𝑘|||𝜑(𝑤)|𝑧𝑤|𝛼2=|||𝑓𝑛𝑗𝑓𝑛𝑘𝑓(𝜑(𝑧))𝑛𝑗𝑓𝑛𝑘|||(𝜑(𝑤))||||||||𝜑(𝑧)𝜑(𝑤)𝜑(𝑧)𝜑(𝑤)|𝑧𝑤|𝛼2𝐶𝑝𝛼2,𝐾2𝑓(𝜑)𝑛𝑗𝑓𝑛𝑘𝑌+𝑓𝑛𝑗𝑓𝑛𝑘𝑌.(2.7) The above inequality is certainly true for all 𝑧,𝑤𝐾2 with 𝑧𝑤 and 𝜑(𝑧)=𝜑(𝑤). Therefore, 𝑝𝛼2,𝐾2𝑓𝑛𝑗𝑓𝜑𝑛𝑘𝜑𝐶𝑝𝛼2,𝐾2𝑓(𝜑)𝑛𝑖𝑓𝑛𝑗𝑌+𝑓𝑛𝑖𝑓𝑛𝑗𝑌,(2.8) and so 𝑓𝑛𝑗𝑓𝜑𝑛𝑘𝜑Lip(𝑋2,𝐾2,𝛼2)1+𝐶𝑝𝛼2,𝐾2𝑓(𝜑)𝑛𝑗𝑓𝑛𝑘𝑌+𝑓𝑛𝑗𝑓𝑛𝑘𝑌.(2.9)
Since 𝑌 is a compact subset of int(𝐾1), we deduce that the sequences {𝑓𝑛𝑗}𝑗=1 and {𝑓𝑛𝑗}𝑗=1 are convergent uniformly on 𝑌. Therefore, {𝑓𝑛𝑗𝜑}𝑗=1 is a Cauchy sequence on Lip(𝑋2,𝐾2,𝛼2), that is {𝑇𝑓𝑛𝑗}𝑗=1 is convergent in Lip(𝑋2,𝐾2,𝛼2). Hence, 𝑇 is compact.

Corollary 2.6. Suppose that 𝛼(0,1], 𝑋 and 𝐾 are compact plane sets with int(𝐾), and 𝐾𝑋. Let 𝐵 be a subalgebra of Lip𝐴(𝑋,𝐾,𝛼) which is a natural Banach function algebra on 𝑋 with the norm Lip(𝑋2,𝐾2,𝛼2), and let 𝜑 be a self-map of 𝑋. If 𝜑 is constant or 𝜑𝐵 with 𝜑(𝑋)int(𝐾), then 𝜑 induces a unital compact endomorphism of 𝐵.

Definition 2.7. (a)A sector in 𝔻(𝑧0,𝑟) at a point 𝜔𝜕𝔻(𝑧0,𝑟) is the region between two straight lines in 𝔻(𝑧0,𝑟) that meet at 𝜔 and are symmetric about the radius to 𝜔. (b)If 𝑓 is a complex-valued function on 𝔻(𝑧0,𝑟) and 𝜔𝜕𝔻(𝑧0,𝑟), then lim𝑧𝜔𝑓(𝑧)=𝐿 means that 𝑓(𝑧)𝐿 as 𝑧𝜔 through any sector at 𝜔. When this happens, we say that 𝐿 is angular (or non-tangential) limit of 𝑓 at 𝜔. (c)An analytic map 𝜑𝔻(𝑧0,𝑟)𝔻𝜌 has an angular derivation at a point 𝜔𝜕𝔻𝑟(𝑧0,𝑟) if for some 𝜂𝜕𝔻𝜌lim𝑧𝜔𝜂𝑓(𝑧)𝜔𝑧(2.10) exists (finitely). We call the limit the angular derivative of 𝜑 at 𝜔 and denote it by 𝜑(𝜔).

Lemma 2.8. Let 0<𝑟1, and let 𝜑𝔻(𝑧0,𝑟)𝔻𝜌 be an analytic function and 𝜓𝔻𝔻 defined by 𝜓(𝑧)=(1/𝜌)𝜑(𝑧0+𝑟𝑧). Then 𝜑 has angular derivation at 𝜔𝜕𝔻(𝑧0,𝑟) if and only if 𝜓 has angular derivation at (𝜔𝑧0)/𝑟𝜕𝔻. Moreover, 𝜑𝑟(𝜔)=𝜌𝜓𝜔𝑧0𝑟.(2.11)

The following result is a modification of Julia-Caratheodory’s theorem. For further details and proof of Julia-Caratheodory’s theorem, see [7, pages 295–300].

Theorem 2.9. Take 0<𝑟1. Let 𝜑𝔻(𝑧0,𝑟)𝔻 be a nonconstant analytic function and 𝜔𝜕𝔻(𝑧0,𝑟). Then the following are equivalent: (i)liminf𝑧𝜔(𝜑𝔻𝑟|𝜑(𝑧)|)/(𝑟|𝑧|)=𝛿<, (ii)lim𝑧𝜔(𝜂𝜑(𝑧))/(𝜔𝑧) exists for some 𝜂𝜕𝔻, (iii)lim𝑧𝜔𝜑(𝑧) exists and lim𝑧𝜔𝜑(𝑧)=𝜂𝜕𝔻.

The boundary point 𝜂 in (ii) and (iii) is the same, and 𝛿>0 in (i). Also the limit of the difference quotients in (ii) coincides with the limit of the derivative in (iii), and both are equal to 𝜔𝜂𝛿.

Note that the existence of the angular derivative 𝜑 at 𝜔𝜕𝔻(𝑧0,𝑟), according to Theorem 2.9, is equivalent to liminf𝑧𝜔(𝜑𝔻(𝑧0,𝑟)|𝜑(𝑧)|)/(𝑟|𝑧𝑧0|)<. In this case the angular derivative of 𝜑 at 𝜔 is nonzero.

Proposition 2.10. Let 𝑋 be a compact plane set, and let 𝔻(𝑧0,𝑟)𝑋 and 𝐾=𝔻(𝑧0,𝑟). Suppose that 𝑐𝜕𝔻(𝑧0,𝑟) and 𝜑Lip𝐴(𝑋,𝐾,1) is a nonconstant function such that |𝜑(𝑐)|=𝜑𝔻(𝑧0,𝑟). Then the angular derivative of 𝜑 at 𝑐 exists and is nonzero.

Proof. Let Γ={𝑧𝔻(𝑧0,𝑟)|𝑧𝑐|/(𝑟|𝑧𝑧0|)<2}. For every 𝑧Γ we have 𝜑𝔻(𝑧0,𝑟)||||𝜑(𝑧)||𝑟𝑧𝑧0||=||||||||𝜑(𝑐)𝜑(𝑧)||𝑟𝑧𝑧0|||𝑧𝑐|||𝑟𝑧𝑧0||||||𝜑(𝑧)𝜑(𝑐)|𝑧𝑐|<2𝑝1,𝐾(𝜑).(2.12) Therefore, liminf𝑧𝜔(𝜑𝔻(𝑧0,𝑟)|𝜑(𝑧)|)/(𝑟|𝑧𝑧0|)<, and, by Theorem 2.9, the proof is complete.

Definition 2.11. (a)A plane set 𝑋 at 𝑐𝜕𝑋 has an internal circular tangent if there exists a disc 𝐷 in the complex plane such that 𝑐𝜕𝐷 and 𝐷{𝑐}int(𝑋). (b)A plane set 𝑋 is called strongly accessible from the interior if it has an internal circular tangent at each point of its boundary. Such sets include the closed unit disc 𝔻 and 𝔻(𝑧0,𝑟)𝑛𝑘=1𝔻(𝑧𝑘,𝑟𝑘), where closed discs 𝔻(𝑧𝑘,𝑟𝑘) are mutually disjoint in 𝔻(𝑧0,𝑟). (c)A compact plane set 𝑋 has peak boundary with respect to 𝐵𝐶(𝑋) if for each 𝑐𝜕𝑋 there exists a nonconstant function 𝐵 such that 𝑋=(𝑐)=1.

Example 2.12. The closed unit disc 𝔻 has peak boundary with respect to 𝐴(𝔻) because, if 𝑐𝜕𝔻, then the function 𝔻𝐶 defined by (𝑧)=(1/2)(1+𝑐𝑧) belongs to 𝐴(𝔻) and satisfies 𝔻=(𝑐)=1.

Let 𝑋 be a compact plane set. The algebra 𝑅(𝑋) consists of all functions in 𝐶(𝑋) which can be approximated by rational functions with poles off 𝑋. It is known that 𝑅(𝑋) is a natural uniform algebra on 𝑋.

Example 2.13. Let 𝑋 be a compact plane set such that 𝑋 is strongly accessible from the interior. If 𝑅(𝑋)𝐵𝐶(𝑋), then 𝑋 has a peak boundary with respect to 𝐵.

Proof. Let 𝑧0𝑋. Since 𝑋 is strongly accessible from the interior, for each 𝑐𝜕(𝑋), there exists a 𝛿>0 such that |𝑐𝑧0|=𝛿 and 𝔻(𝑧0,𝛿)int(𝑋). Now, we define the function 𝑋 by 𝛿(𝑧)=2𝑐𝑧0𝑧𝑧0.(2.13) Then 𝐵, 𝑋=(𝑐)=1.

Theorem 2.14. Let 𝑋1 be a compact plane set such that 𝐺1=int(𝑋1) is connected, and 𝐺1=𝑋1. Suppose that 𝑋1 has peak boundary with respect to Lip𝐴(𝑋1,1). Let Ω1𝐺1 be a bounded connected open set in the complex plane, and let 𝐾1=Ω1. Let Ω2 be a bounded connected open set in the complex plane, and let 𝐾2=Ω2 such that 𝐾2 is strongly accessible from the interior. Suppose that 𝑋2 is a compact plane set such that 𝐾2𝑋2. If 𝑇Lip𝐴(𝑋1,𝐾1,1)Lip𝐴(𝑋2,𝐾2,1) is a unital compact homomorphism, then 𝑇 is induced by a continuous mapping 𝜑𝑋2𝑋1 such that 𝜑 is constant on 𝐾2 or 𝜑(𝐾2)𝐺1=int(𝑋1). Moreover, 𝜑=𝑇𝑍.

Proof. Since Lip𝐴(𝑋1,𝐾1,1) and Lip𝐴(𝑋2,𝐾2,1) are, respectively, natural Banach function algebras on 𝑋1 and 𝑋2, 𝑇Lip𝐴(𝑋1,𝐾1,1)Lip𝐴(𝑋2,𝐾2,1) is a unital homomorphism, 𝑋1 is a compact plane set, and 𝑍Lip𝐴(𝑋1,𝐾1,1), we conclude that 𝑇 is induced by 𝜑=𝑇𝑍 and so 𝜑Lip𝐴(𝑋2,𝐾2,1) by Proposition 1.2. Suppose that 𝜑 is nonconstant on Ω2. Since 𝜑 is analytic on Ω2, we deduce that 𝜑(Ω2) is an open subset of 𝑋1 and so 𝜑(Ω2)𝐺1. We now show that 𝜑(𝐾2)𝐺1. Suppose that 𝜑(𝐾2)̸𝐺1. Then there exists 𝑐𝜕𝐾2 such that 𝜑(𝑐)𝜕𝑋1. Since 𝑋1 has peak boundary with respect to Lip𝐴(𝑋1,1), there exists a nonconstant function Lip𝐴(𝑋1,1) such that 𝑋1=(𝜑(𝑐))=1. We now define the sequence {𝑓𝑛}𝑛=1 of complex-valued functions on 𝑋1 by 𝑓𝑛(𝑧)=(1/𝑛)𝑛(𝑧). Let 𝑛. Then 𝑓𝑛𝑋1=1𝑛𝑋1𝑛=1𝑛,𝑝(2.14)1,𝐾1𝑓𝑛||=sup𝑛(𝑧)𝑛||(𝑤)𝑛|𝑧𝑤|𝑧,𝑤𝐾1||||,𝑧𝑤sup(𝑧)(𝑤)|𝑧𝑤|𝑧,𝑤𝐾1||||,𝑧𝑤sup(𝑧)(𝑤)|𝑧𝑤|𝑧,𝑤𝑋1,𝑧𝑤𝑝1,𝑋1().(2.15) Thus 𝑓𝑛Lip(𝑋1,𝐾1,1)1𝑛+𝑝1,𝑋1()1+𝑝1,𝑋1(),(2.16) by (2.14) and (2.15). This implies that {𝑓𝑛}𝑛=1 is a bounded sequence in Lip𝐴(𝑋1,𝐾1,1). The compactness of homomorphism 𝑇 implies that there exists a subsequence {𝑓𝑛𝑗}𝑗=1 of {𝑓𝑛}𝑛=1 and a function 𝑔 in Lip𝐴(𝑋2,𝐾2,1) such that lim𝑗𝑇𝑓𝑛𝑗𝑔Lip(𝑋2,𝐾2,1)=0.(2.17) This implies that lim𝑗𝑇𝑓𝑛𝑗𝑔𝑋2=0.(2.18) On the other hand, we have 𝑇𝑓𝑛𝑗𝑋21/𝑛𝑗 for all 𝑗 by (2.14). Hence, lim𝑗𝑇𝑓𝑛𝑗𝑋2=0.(2.19) By (2.18) and (2.19), 𝑔=0. Therefore, by (2.17) we have lim𝑗𝑇𝑓𝑛𝑗Lip(𝑋2,𝐾2,1)=0.(2.20) This implies that lim𝑗𝑝1,𝐾2𝑓𝑛𝑗𝜑=0.(2.21) Assume that 𝜀>0. By (2.21), there exists a natural number 𝑁 such that for each 𝑗 with 𝑗𝑁|||𝑓sup𝑛𝑗𝑓𝜑(𝑧)𝑛𝑗|||𝜑(𝑤)|𝑧𝑤|𝑧,𝑤𝐾2,𝑧𝑤<𝜀.(2.22) In particular, ||sup((𝜑)(𝑧))𝑛𝑁((𝜑)(𝑤))𝑛𝑁||𝑛𝑁|𝑧𝑤|𝑧,𝑤𝐾2,𝑧𝑤<𝜀.(2.23) This implies that 1𝑛𝑁||sup((𝜑)(𝑧))𝑛𝑁((𝜑)(𝑐))𝑛𝑁|||𝑧𝑐|𝑧𝐾2,𝑧𝑐<𝜀.(2.24) Since 𝑐𝜕𝐾2 and 𝐾2 is strongly accessible from the interior, there exists an open disc 𝐷=𝔻(𝑧0,𝑟) such that 𝑐𝜕𝐷 and 𝐷{𝑐}int(𝐾2). Since 𝜑 is analytic on int(𝐷)int(𝐾2) and is analytic on 𝜑(𝐷)int(𝑋1), we deduce that 𝜑 is analytic on int(𝐷). On the other hand, we can easily show that 𝑝1,𝐷(𝜑)𝑝1,𝑋1()𝑝1,𝐾2(𝜑)<.(2.25) Therefore, 𝜑Lip𝐴(𝑋2,𝐷,1). Since 𝑋1=(𝜑(𝑐))=1, we conclude that (𝜑)(𝑐)=𝜑𝐷=1.(2.26) We claim that 𝜑 is constant on 𝐷. If 𝜑 is nonconstant on 𝐷, then, by Proposition 2.10, (𝜑)(𝑐) exists and is nonzero and since (𝜑)𝑛𝑁(𝑐)=1, (𝜑)𝑛𝑁(𝑐)𝜕𝐷. If Γ is a sector in 𝐷 at 𝑐𝜕𝐷, then 1𝑛𝑁lim𝑧𝑐𝑧Γ||||(𝜑)𝑛𝑁(𝑧)(𝜑)𝑛𝑁(𝑐)||||𝑧𝑐𝜀(2.27) by (2.24). Thus 1𝑛𝑁(𝜑)𝑛𝑁(𝑐)𝜀.(2.28) But (𝜑)𝑛𝑁(𝑐)=𝑛𝑁(𝜑)𝑛𝑁1(𝑐)(𝜑)(𝑐).(2.29) Hence, by (2.28) we have (𝜑)1(𝑐)=𝑛𝑁(𝜑)𝑛𝑁(𝑐)𝜀.(2.30) Since 𝜀 is assumed to be a positive number, we conclude that (𝜑)(𝑐)=0, contradicting to (𝜑)(𝑐)0. Hence, our claim is justified. Since 𝜑 is nonconstant on 𝐾2, 𝜑 is a nonconstant analytic function on connected open set 𝐷. This implies that 𝜑(𝐷) is a connected open set in the complex plane. This implies that is constant on connected open set 𝐺1. The continuity of on 𝑋1=𝐺1 follows that is constant on 𝐺1=𝑋1. This contradicting to is nonconstant on 𝑋1. Therefore, 𝜑(𝐾2)𝐺1.

Corollary 2.15. Let 𝑋 be a compact plane set such that 𝐺=int(𝑋) is connected and 𝐺=𝑋. Let Ω𝐺 be a bounded connected open set in the complex plane, and let 𝐾=Ω. Suppose that 𝐾 is strongly accessible from the interior and 𝑋 has peak boundary with respect to Lip𝐴(𝑋,1). If 𝑇 is a unital compact endomorphism of Lip𝐴(𝑋,𝐾,1), then 𝑇 is induced by a continuous self-map 𝜑 of 𝑋 such that 𝜑 is constant on 𝐾 or 𝜑(𝐾)𝐺=int(𝑋). Moreover, 𝜑=𝑇𝑍.

Lemma 2.16. Let 𝐺 and Ω be bounded connected open sets in the complex plane with Ω𝐺, and let 𝑋=𝐺 and 𝐾=Ω. Then for each 𝑐𝐺𝐾 there exists a function 𝑓𝑐Lip𝐴(𝑋,𝐾,1) such that 𝑓𝑐 is not analytic at 𝑐.

Proof. Let 𝑐𝐺𝐾. Then there is a positive number 𝑟 such that {𝑧|𝑧𝑐|𝑟}𝐺𝐾.(2.31) We now define the function 𝑓𝑐𝑋 by 𝑓𝑐(𝑧)=𝑧𝑐𝑧𝑋,|𝑧𝑐|𝑟,(1+𝑟)(𝑧𝑐)1+|𝑧𝑐|𝑧𝑋,|𝑧𝑐|<𝑟.(2.32) It is easily seen that 𝑓𝑐Lip𝐴(𝑋,𝐾,1) and 𝑓𝑐 is not analytic at 𝑐.

Definition 2.17. Let 𝑋 and 𝐾 be compact plane sets such that 𝐾𝑋. We say that 𝐾 has peak 𝐾-boundary with respect to 𝐵𝐴(𝑋,𝐾) if for each 𝑐𝜕𝐾 there is a function 𝐵 such that is nonconstant on 𝐾 and 𝑋=(𝑐)=1.

Example 2.18. Let 𝑟(0,1] and 𝐾=𝔻𝑟. Suppose that Lip𝐴(𝔻,𝐾,1)𝐵𝐴(𝔻,𝐾). Then 𝐾 has peak 𝐾-boundary with respect to 𝐵.

Proof. We first assume that 𝑟=1. If for each 𝑐𝜕𝔻 the function 𝔻 defined by (𝑧)=(1/2)(1+𝑐𝑧), then 𝐵, is nonconstant on 𝐾=𝔻 and (𝑐)=1=𝔻.
We now assume that 0<𝑟<1. For each 𝑐𝜕𝐾, set 𝑧0=(1+𝑟)𝑐/𝑟. Then 𝑧0𝔻. Define the function 𝔻 by 𝑟(𝑧)=𝑐𝑧𝑧0𝑧||𝔻,𝑧𝑧0||𝑟||1,𝑧𝑧0||𝑐𝑧𝑧0𝑧||𝔻,𝑧𝑧0||<1.(2.33) It is easily seen that Lip𝐴(𝔻,𝐾,1) and 𝔻=1=(𝑐).

Lemma 2.19. Let Ω be a connected open set in the complex plane, and let 𝜑 be a one-to-one analytic function on Ω. If 𝑓 is a continuous complex-valued function on 𝜑(Ω) and 𝑓𝜑 is analytic on Ω, then 𝑓 is an analytic function on 𝜑(Ω).

Proof. By [8, Theorem 7.5 and Corollary 7.6 in Chapter IV], we deduce that 𝜑(Ω) is a connected open set in the complex plane, 𝜑(𝑧)0 for all 𝑧Ω, and 𝜑1𝜑(Ω)Ω is an analytic function on 𝜑(Ω). Since 𝑓=𝑓𝜑𝜑1 on 𝜑(Ω), we conclude that 𝑓 is analytic on 𝜑(Ω).

Theorem 2.20. Let 𝑋1 be a compact plane set such that 𝐺1=int(𝑋1) is connected and 𝐺1=𝑋1. Suppose that 𝐾1 is a compact subset of 𝑋1 such that Ω1=int(𝐾1) is connected, 𝐾1=Ω1, and 𝐾1 has peak 𝐾1-boundary with respect to Lip𝐴(𝑋1,𝐾1,1). Let Ω2 be a bounded connected open set in the complex plane, and let 𝐾2=Ω2 such that 𝐾2 is strongly accessible from the interior. Suppose that 𝑋2 is a compact plane set such that 𝐾2𝑋2. If 𝑇Lip𝐴(𝑋1,𝐾1,1)Lip𝐴(𝑋2,𝐾2,1) is a unital compact homomorphism and 𝑇𝑍 is one-to-one on Ω2, then 𝑇 is induced by a continuous mapping 𝜑𝑋2𝑋1 such that 𝜑=𝑇𝑍 and 𝜑(𝐾2)Ω1=int(𝐾1).

Proof. Since Lip𝐴(𝑋1,𝐾1,1) and Lip𝐴(𝑋2,𝐾2,1) are, respectively, natural Banach function algebras on 𝑋1 and 𝑋2, 𝑇Lip𝐴(𝑋1,𝐾1,1)Lip𝐴(𝑋2,𝐾2,1) is a unital homomorphism, 𝑋1 is a compact plane set, and 𝑍Lip𝐴(𝑋1,𝐾1,1), we conclude that 𝑇 is induced by 𝜑=𝑇𝑍 and so 𝜑Lip𝐴(𝑋2,𝐾2,1) by Proposition 1.2.
To prove 𝜑(𝐾2)Ω1, we first show that 𝜑(Ω2)𝐾1. Since 𝜑 is a one-to-one analytic mapping on Ω2, we conclude that 𝜑(Ω2) is an open set in the complex plane. This follows that 𝜑(Ω2)int(𝑋1)=𝐺1 since 𝜑(Ω2)𝑋1. Suppose that 𝜑(Ω2)̸𝐾1. Then there exists 𝜆Ω2 such that 𝜑(𝜆)𝐺1𝐾1. By Lemma 2.16, there exists a function 𝑓Lip𝐴(𝑋1,𝐾1,1) such that 𝑓 is not analytic at 𝜑(𝜆). But 𝑓𝜑=𝑇𝑓Lip𝐴(𝑋2,𝐾2,1), so that 𝑓𝜑 is analytic on Ω2. Since 𝑓 is continuous on 𝜑(Ω2) and 𝜑 is a one-to-one analytic function on Ω2, we conclude that 𝑓 is analytic on 𝜑(Ω2) by Lemma 2.19. This contradicts to the fact 𝑓 is not analytic at 𝜑(𝜆)𝜑(Ω2). Therefore, 𝜑(Ω2)𝐾1 so 𝜑(Ω2)int(𝐾1)=Ω1 since 𝜑(Ω2) is an open set in the complex plane. Since 𝜑 is continuous on 𝐾2, 𝜑(Ω2)Ω1, 𝐾2=Ω2, and 𝐾1=Ω1, we can easily show that 𝜑(𝐾2)𝐾1. We now show that 𝜑(𝐾2)Ω1. Suppose that 𝜑(𝐾2)̸Ω1. Then there exists 𝑐𝜕𝐾2 such that 𝜑(𝑐)𝜕𝐾1. Since 𝐾1 has peak 𝐾1-boundary with respect to Lip𝐴(𝑋1,𝐾1,1), there exists a function Lip𝐴(𝑋1,𝐾1,1) such that is not constant on 𝐾1 and 𝑋1=(𝜑(𝑐))=1.(2.34) Applying the similar argument used in the proof of Theorem 2.14, we can prove that is constant on 𝐾1. This contradiction shows that 𝜑(𝐾2)Ω1.

Corollary 2.21. Let 𝑋 be a compact plane set such that 𝐺=int(𝑋) is connected and 𝐺=𝑋. Let 𝐾 be a compact subset of 𝑋 such that Ω=int(𝐾) is connected and 𝐾=Ω. Suppose that 𝐾 has peak 𝐾-boundary with respect to Lip𝐴(𝑋,𝐾,1) and 𝐾 is strongly accessible from the interior. If 𝑇 is a unital compact endomorphism of Lip𝐴(𝑋,𝐾,1) and 𝑇𝑍 is a one-to-one mapping on Ω, then 𝑇 is induced by a continuous self-map 𝜑 of 𝑋 such that 𝜑=𝑇𝑍 and 𝜑(𝐾)Ω=int(𝐾).

3. Spectrum of Unital Compact Endomorphisms

In this section we determine the spectrum of a unital compact endomorphism of a subalgebra of Lip𝐴(𝑋,𝐾,𝛼) which is a natural Banach function algebra with the norm Lip(𝑋,𝐾,𝛼).

The following result is a modification of [9, Theorem 1.7] for unital compact endomorphisms of natural Banach function algebras.

Theorem 3.1. Let 𝑋 be a compact Hausdorff space and 𝐵 a natural Banach function algebra on 𝑋. If 𝑇 is a unital compact endomorphism of 𝐵 induced by a self-map 𝜑𝑋𝑋, then 𝑛=0𝜑𝑛(𝑋) is finite, and if 𝑋 is connected, 𝑛=0𝜑𝑛(𝑋) is singleton where 𝜑𝑛 is the 𝑛th iterate of 𝜑, that is, 𝜑0(𝑥)=𝑥 and 𝜑𝑛(𝑥)=𝜑(𝜑𝑛1(𝑥)). If 𝑛=0𝜑𝑛(𝑋)={𝑥0}, then 𝑥0 is a fixed point for 𝜑. In fact, if 𝐹=𝑛=0𝜑𝑛(𝑋), then 𝜑(𝐹)=𝐹.

Theorem 3.2. Suppose that 𝑋 is a compact plane set with int(𝑋), Ω is a connected open set in the complex plane with Ωint(𝑋), and 𝐾=Ω. Let 𝐵 be a subalgebra of 𝐴(𝑋,𝐾) containing the coordinate function 𝑍 which is a natural Banach function algebra on 𝑋 with an algebra norm 𝐵. Let 𝑇 be a unital compact endomorphism of 𝐵 induced by a self-map 𝜑 of 𝑋. If 𝜑(𝑋)int(𝐾) and 𝑧0 is a fixed point of 𝜑, then 𝜑𝜎(𝑇)={0,1}𝑧0𝑛.𝑛(3.1)

Proof. Clearly 0 and also 1𝜎(𝑇) since 𝑇(1𝑋)=1𝑋. If 𝜑 is constant then the proof is complete. Let 𝜆𝜎(𝑇){0,1}. The compactness of 𝑇 implies that there exists 𝑓𝐵{0} such that 𝑇𝑓=𝑓𝜑=𝜆𝑓. Since 𝜑(𝑧0)=𝑧0int(𝐾), 𝑓(𝑧0)=0. We claim that 𝑓(𝑗)(𝑧0)0 for some 𝑗. If 𝑓(𝑛)(𝑧0)=0 for all 𝑛, then 𝑓=0 on an open disc with center 𝑧0 and so on Ω. By maximum modules principle, it follows that 𝑓=0 on 𝑋 since 𝜑(𝑋)Ω, 𝜆{0}, and 𝜆𝑓(𝑧)=𝑓(𝜑(𝑧)) for all 𝑧𝑋. This contradicts to 𝑓0. Hence, our claim is justified. Let 𝑚=min{𝑛𝑓(𝑛)(𝑧0)0}. Then 𝑓(𝑘)(𝑧0)=0 for all 𝑘{0,,𝑛1} and 𝑓(𝑚)(𝑧0)0. By 𝑚 times differentiation of 𝑓𝜑=𝜆𝑓, we have (𝜑(𝑧0))𝑚𝑓(𝑚)(𝜑(𝑧0))=𝜆𝑓(𝑚)(𝑧0), and therefore 𝜆=(𝑓(𝑧0))𝑚. Then 𝜎(𝑇){0,1}{(𝜑(𝑧0))𝑛𝑛}.
Conversely, first we show that, if 𝜆𝜎(𝑇) with |𝜆|=1, then 𝜆=1. Let 𝜆𝜎(𝑇) and |𝜆|=1. The compactness of 𝑇 implies that there exists 𝑔𝐵{0} such that 𝑔𝜑=𝜆𝑔. It follows that |𝑔𝜑|=|𝑔|. Since 𝜑(𝐾)int(𝐾)=Ω and 𝑔 is analytic on the connected open set Ω, we conclude that 𝑔 is constant on Ω by maximum modules principle. Since 𝜑(𝑋)Ω, 𝑔𝜑=𝜆𝑔, and 𝜆{0}, we deduce that 𝑔 is constant on 𝑋. Applying again 𝑔𝜑=𝜆𝑔 implies that 𝜆=1 since 𝑔𝐵{0}.
We now claim that 𝜑(𝑧0)𝜎(𝑇). If 𝜑(𝑧0)𝜎(𝑇), then there exists a nonzero linear operator 𝑆𝐵𝐵 such that 𝑇𝜑𝑧0𝐼𝑆=𝐼.(3.2) Since 𝑍𝑧01𝑋𝐵, =𝑆(𝑍𝑧01𝑋)𝐵 and so 𝜑𝜑𝑧0=𝑍𝑧01𝑋,(3.3) by (3.2). By differentiation at 𝑧0, we have 0=𝜑𝑧0𝜑𝑧0𝜑𝑧0𝑧0=1,(3.4) this is a contradiction. Hence, our claim is justified.
We now show that (𝜑(𝑧0))𝑛𝜎(𝑇) for all 𝑛. If 𝜑(𝑧0)=0 or |𝜑(𝑧0)|=1, the proof is complete. Suppose that 𝜑(𝑧0)0 and |𝜑(𝑧0)|1. If (𝜑(𝑧0))𝑗𝜎(𝑇) for some 𝑗 with 𝑗>1, then there exists a nonzero linear operator 𝑆𝑗𝐵𝐵 such that 𝜑𝑇𝑧0𝑗𝐼𝑆𝑗=𝐼.(3.5) Since (𝑍𝑧01𝑋)𝑗𝐵, 𝑗=𝑆𝑗(𝑍𝑧01𝑋)𝑗𝐵 and so 𝑗𝜑𝜑𝑧0𝑗𝑗=𝑍𝑧01𝑋𝑗,(3.6) by (3.5). By 𝑗1 times differentiation at 𝑧0, we have 𝑗𝑧0=𝑗𝑧0==𝑗(𝑗1)𝑧0=0,(3.7) and by 𝑗 times differentiation at 𝑧0, we have 𝜑0=𝑧0𝑗𝑗(𝑗)𝜑𝑧0𝜑𝑧0𝑗𝑗(𝑗)𝑧0=𝑗!,(3.8) this is a contradiction. Thus, (𝜑(𝑧0))𝑛𝜎(𝑇) for all 𝑛. This completes the proof.

Corollary 3.3. Let 𝐵 and 𝑇 satisfy the conditions of Theorem 3.2, and let 𝐵 be a natural Banach function algebra with the norm 𝛼,𝐾. If 𝐹 is a finite set such that 𝜑(𝐹)=𝐹, then there exist 𝑧0𝐹 and 𝑚 such that {𝜆𝑚𝜑𝜆𝜎(𝑇)}={0,1}𝑚𝑧0𝑛.𝑛(3.9)

Proof. Since 𝐹 is a finite set and 𝜑(𝐹)=𝐹, there exist 𝑧0𝐹 and 𝑚 such that 𝜑𝑚(𝑧0)=𝑧0. Since 𝜑(𝑋)int(𝐾), so 𝑧0int(𝐾). If 𝜑 is constant, then the proof is complete. When 𝜑 is nonconstant, we define 𝑇𝐵𝐵 by 𝑇𝑓=𝑓𝜑𝑚. Then 𝑇 is a compact endomorphism of 𝐵 induced by 𝜑𝑚 by Corollary 2.6 and 𝜑𝑚(𝑧0)=𝑧0. Therefore, 𝜎𝑇𝜑={0,1}𝑚𝑧0𝑛𝑛(3.10) by Theorem 3.2. Since 𝑇𝑓=𝑓𝜑 and 𝑇𝑓=𝑓𝜑𝑚 for all 𝑓𝐵, we have 𝑇=𝑇𝑚. By Spectral Mapping Theorem, 𝜎(𝑇𝑚)={𝜆𝑚𝜆𝜎(𝑇)}. Therefore, {𝜆𝑚𝜑𝜆𝜎(𝑇)}={0,1}𝑚𝑧0𝑛.𝑛(3.11)

Corollary 3.4. Suppose that 𝑋 is a compact plane set with int(𝑋), Ω is a connected open set in the complex plane with Ωint(𝑋), and 𝐾=Ω. Take 𝛼(0,1]. Let 𝜑 be a self-map of 𝑋 with 𝜑(𝑋)Ω such that 𝜑Lip𝐴(𝑋,𝐾,𝛼), and let 𝜑(𝑧0)=𝑧0 for some 𝑧0Ω. If 𝑇 is a unital endomorphism of Lip𝐴(𝑋,𝐾,𝛼) induced by 𝜑, then 𝑇 is compact and 𝜑𝜎(𝑇)={0,1}𝑧0𝑛.𝑛(3.12)