1 Introduction

The study of circle packings, as they are understood in this paper, was initiated by Paul Koebe as early as 1936 in the context of conformal mapping, but the real success of the topic began with William Thurston’s talk at the celebration of the proof of the Bieberbach conjecture in 1985. The publication of Ken Stephenson’s book (Stephenson (2005)) inspired further research and made the topic accessible to a wide audience. Since then many classical results in complex analysis found their discrete counterpart in circle packing.

In this paper we consider circle packings embedded in a bounded, simply connected domain. We introduce the concept of crosscuts for domain-filling circle packings, and study the rigidity of packings with respect to maximal crosscuts (for definitions see below). The main result is a discrete version of an identity theorem for analytic functions, which has implications to uniqueness results for boundary value problems for circle packings, and especially to discrete conformal mappings.

To be more specific, we recall that the tangency relations of a circle packing are encoded in a 2-dimensional simplicial complex \(K\), referred to as the combinatorics of the packing. In this paper it is assumed that \(K\) is a finite triangulation of a topological disk.

Circle packings are a mixture of flexibilty and rigidity. Counting the degrees of freedom for the centers and the radii, and comparing this with the number of conditions caused by the tangency relations, we see that the first number exceeds the latter by \(m+3\), where \(m\) is the number of boundary circles. In fact, the set of all circle packings for a fixed complex \(K\) forms a smooth manifold of dimension \(m+3\) (Bauer et al. 2012). So the question arises which sort of conditions are appropriate to eliminate the flexibility of a packing and make it rigid. Motivated by our work on nonlinear Riemann-Hilbert problems, we are interested in boundary value problems for circle packings. These problems involve \(m\) boundary conditions (one for each boundary circle) and three additional conditions, which can be imposed in different form on boundary circles and interior circles as well.

A standard boundary value problem of this kind consists in finding circle packings with (given combinatorics and) prescribed radii of its boundary circles. Somewhat surprisingly, this problem has always a locally univalent solution, and the solution is unique up to a rigid motion of the complete packing (see Stephenson 2005, Sect. 11.4, for details).

The existence of solutions is also known for a related more general problem, the discrete Beurling problem, where the radii of the boundary circles are prescribed as functions of their centers (see Wegert et al. 2012), but the question of uniqueness has not yet been answered satisfactorily.

Last but not least there are several approaches to discrete conformal mapping via circle packing which fall into this category (see  Stephenson 2005, in particular Chap. 19 and 20, with many interesting comments on the history of this topic, also summarizing He and Schramm (1996), Rodin and Sullivan (1987) and Thurston (1985).

In our favorite setting of discrete conformal mapping, the domain packing \(\mathcal {P}\) is a so-called maximal packing, which ‘fills’ the complex unit disk \(\mathbb {D}\), while the range packing \(\mathcal {P}'\) is required to ‘fill’ a bounded, simply connected domain \(G\). That a packing ‘fills a domain \(G\)’ basically means that all its circles lie in the closure \(\overline{G}\) of that domain and all its boundary circles intersect (touch) the boundary \(\partial G\) of \(G\). For domains which are not Jordan this has to be complemented by a more subtle condition (see Definition 2).

In a series of papers, Oded Schramm proved several outstanding results about packings which fill a Jordan domain. His very general existence theorems do not only address packings of circles, but of much more general packable sets (for an explanation see Schramm 1990).

Surprisingly, much less is known about uniqueness. It is clear that uniqueness of a domain-filling (circle) packing can only be expected if one imposes additional conditions which eliminate the (three) remaining degrees of freedom. Whether this works depends on the type of normalization conditions and on the geometry of the domain. For example, in his uniqueness proofs, Schramm needs that the Jordan domain is (as he says) decent (see Schramm 1991).

This paper is devoted to the question which additional conditions are appropriate to make a domain filling circle packing unique. In analogy to the standard normalization of conformal mappings, it seems reasonable to fix the center of a distinguished circle (the so-called alpha-circle) at some point in \(G\) and to require that the center of a neighboring circle lies on a given ray emerging from that point. Keeping the first condition, we have chosen another setting for the second one. This condition, involving crosscuts, is non-standard, more flexible and allows one to address other uniqueness problems too.

In order to give the reader a flavor of the result, we first state an analogous theorem for analytic functions. Recall that a crosscut of a domain \(G\) in the complex plane \(\mathbb {C}\) is an open Jordan arc \(J\) in \(G\) such that \(\overline{J}=J \cup \{a,b\}\) with \(a,b\in \partial G\) (see Pommerenke 1992). Slightly abusing terminology, we shall also denote \(\overline{J}\) as a crosscut in \(G\).

Theorem 1

(Identity Theorem for Analytic Functions) Let \(J\) be a crosscut of a simply connected domain \(G\), with \(G^-\) and \(G^+\) denoting the (simply connected) components of \(G{\setminus }J\). If \(f: G \rightarrow G\) is analytic, \(f(z_0)=z_0\) for some \(z_0\in G^+\), and \(f(G^-)\subset G^-\), then \(f(z)=z\) for all \(z\in G\).

Proof

Let \(g:G \rightarrow \mathbb {D}\) be a conformal mapping of \(G\) onto the unit disk \(\mathbb {D}\) with \(g(z_0)=0\). Then \(g\) maps the crosscut \(J\) of \(G\) to a crosscut of \(\mathbb {D}\) (see Pommerenke 1992, Prop. 2.14) and the composition \(g \circ f \circ g^{-1}\) satisfies the assumptions of the lemma with \(G:=\mathbb {D}\) and \(z_0:=0\). Hence it suffices to consider this special case.

Let \(z_1\) be a point on \(J\) with \(|z_1|=\min _{z\in J} |z|\). Since \(J\) is a crosscut in \(\mathbb {D}\), and \(0=z_0\in G^+\), we have

$$\begin{aligned} 0<|z_1|\le \min \left\{ |z|: z\in \overline{G^-}\right\} <1. \end{aligned}$$

By continuity, \(f(G^-)\subset G^-\) and \(z_1\in \overline{G^-}\) imply that \(f(z_1)\in \overline{G^-}\), and hence \(|f(z_1)|\ge |z_1|\). Invoking Schwarz’ Lemma, we get \(f(z)=cz\) in \(\mathbb {D}\), where \(c\) is a unimodular constant. Finally, the only rotation of \(\mathbb {D}\) which maps \(G^-\) into itself is the identity. \(\square \)

Although Schwarz’ Lemma has already been investigated in the framework of circle packing (see Rodin 1987; Pommerenke 1992, Chap. 13) the following interpretation of Theorem 1 is new. Though precise definitions will be deferred to the next section, we hope that Fig. 1 helps to get an intuitive understanding of the setting. The domain \(G^-\) is the one containing the brighter (yellow) disks.

Fig. 1
figure 1

A domain-filling packing \(\mathcal {P}\) with a crosscut and a maximal crosscut

Theorem 2

(Rigidity of Circle Packings with Crosscuts) Assume that a univalent circle packing \(\mathcal {P}=\{D_v\}\) for a complex \(K\) with vertex set \(V\) fills a bounded, simply connected domain \(G\). Let \(J\) be a (maximal) crosscut of \(\mathcal {P}\) in \(G\), such that \(G^-\) is a simply connected component of \(G{\setminus } J\), and denote by \(V^-\) and \(V^+\) the sets of vertices of \(K\) associated with circles in \(G^-\) and \(G^+:=G{\setminus }\overline{G^-}\), respectively. Let \(D_\alpha \) be an interior circle of \(\mathcal {P}\) which is contained in \(G^+\).

Assume further that a second univalent packing \(\mathcal {P}'=\{D_v'\}\) for \(K\) is contained in \(G\), such that \(D_\alpha \) and \(D_\alpha '\) have the same center, and \(D'_v\subset G^-\) for all \(v\in V^-\). Then \(D_v'=D_v\) for all accessible vertices \(v\in V\).

As follows from a simple topological argument, the condition \(D'_v\subset G^-\) need only be required for those vertices \(v\in V^-\) which are associated with circles \(D_v\) touching the crosscut \(J\).

We point out that everything hinges on the assumption about the common center of the two alpha-circles. Since we do not assume that \(\mathcal {P}'\) fills \(G\), it is solely this condition which prevents \(\mathcal {P}'\) from lying entirely in \(G^-\).

The notion of accessible vertices will be explicated in Definition 1. Here we only note that all vertices \(v\in V\) are accessible if and only if the complex \(K\) is strongly connected, which means \(K\) satisfies the following conditions (1) and (2):

  1. (1)

    Every boundary vertex has an interior neighbor.

  2. (2)

    The interior of \(K\) is connected.

Note that some authors of the circle packing community make the general assumption that the underlying complex \(K\) is strongly connected (see Stephenson 2005). For circle packings with this simpler combinatoric structure the theorem yields complete rigidity with respect to crosscuts, i.e., \(D_v'=D_v\) for all \(v\in V\).

Figure 2 illustrates some effects which can be observed for packings with general combinatorics. The picture on the left shows an Apollonian packing \(\mathcal {P}\) with four generations. The highlighted line is a maximal crosscut, separating the disks in the “lower domain” from the disks in the “upper domain”. The disk with the darkest color is the alpha-disk with fixed center. The accessible disks are those which can be connected with the alpha-disk by a chain of interior disks (see Definition 1).

Fig. 2
figure 2

Some examples illustrating assumptions and assertions of Theorem 2

The packing \(\mathcal {P}'\), depicted in the middle, satisfies the assumptions of the theorem. In this example, only the accessible disks of \(\mathcal {P}'\) (shown in darker colors) coincide with their partners in \(\mathcal {P}\). The non-accessible disks (shown in lighter colors) differ from the corresponding disks in \(\mathcal {P}\).

The example on the right illustrates that the result need not hold if the alpha-disk is a boundary disk. The depicted packing \(\mathcal {P}''\) satisfies all other assumptions (for the same crosscut), but, apart from the alpha disk, it is completely different from the packing \(\mathcal {P}\) shown on the left-hand side.

The result has an intuitive interpretation when we think of circle packings as dynamic structures: Suppose that \(\mathcal {P}\) fills \(G\), and allow its circles to move (change position and size) in such a way, that they all remain in \(G\), the center of the alpha-circle is fixed in \(G^+\), and the circles in \(G^-\) are not allowed to leave \(G^-\). Then only those circles which are not accessible can be moved, while the core part of the packing is rigid.

In order to illustrate the analogies with Theorem 1, we interpret the result in the framework of discrete analytic functions: The circle packing \(\mathcal {P}\) filling \(G\) is the domain packing, the packing \(\mathcal {P}'\) lies in \(G\), so that \(\mathcal {P}\rightarrow \mathcal {P}'\) defines a discrete analytic function from \(G\) into itself. Fixing the centers of the alpha-circles of both packings at the same point \(z_0\) corresponds to the normalization \(f(z_0)=z_0\). Finally, the condition \(D'_v\subset G^-\) for all \(v\in V^-\) expresses the invariance of the subdomain \(G^-\).

Since the packing \(\mathcal {P}\) represents the identity function on \(G\), it is natural to suppose that \(\mathcal {P}\) is univalent. Contrary to the continuous setting of Theorem 1, also \(\mathcal {P}'\) was assumed to be univalent in Theorem 2. It is challenging to investigate what happens when this condition is dropped.

Terminological remark. For our purposes it would be better to work with disk packings instead of circle packings. Though we stay with the traditional notion, we shall often speak of the disks in a circle packing. In order to avoid cumbersome formulations, we also say that a circle \(\partial D\) lies in a domain \(G\) when this holds for the open disk \(D\) bounded by that circle. We already made use of this convention above.

2 Circle packings

In order to make the paper self-contained we recall basic concepts and notions of topology and circle packing (for details we refer to  Henle 1979;  Stephenson 2005).

Some geometry If \(A\) and \(B\) are subsets of the (complex) plane, we say that \(A\) intersects \(B\) if \(A\cap B\not =\emptyset \). If \(A\) is a disk, then the phrase \(A\) touches \(B\) is in general used when \(\overline{A}\cap \overline{B}\not =\emptyset \) and \(A\cap B=\emptyset \). In this case we also say that the circle \(\partial A\) touches \(B\). As usual, the symbol \(\partial \) denotes the boundary operator.

By a curve \(\gamma \) we understand the image of a continuous mapping \(\varphi :[a,b]\rightarrow \mathbb {C}\). The points \(\varphi (a)\) and \(\varphi (b)\) are said to be the initial point and the terminal point of \(\gamma \), respectively; both are referred to as endpoints of \(\gamma \). A Jordan arc and a Jordan curve are the homeomorphic images of a segment and a circle, respectively. By an open Jordan arc we mean a Jordan arc without its endpoints.

Let \(J\) be an oriented Jordan curve. For \(p,q\in J\) with \(p\not = q\) we denote by \(J(p,q)\) the (oriented) open subarc of \(J\) with initial point \(p\) and terminal point \(q\). If \(p,q,r\) are three pairwise different points on \(J\), we say that \(q\) lies between \(p\) and \(r\) on \(J\) if \(q\in J(p,r)\). Corresponding to whether \(q\) lies between \(p\) and \(r\), or \(q\) lies between \(r\) and \(p\), the orientation of the triplet \((p,q,r)\) with respect to \(J\) is said to be positive or negative, respectively.

Let \(G\) be a bounded, simply connected domain in \(\mathbb {C}\). A conformal mapping \(g: \mathbb {D}\rightarrow G\) of \(\mathbb {D}\) onto \(G\) has a continuous extension to \(\overline{\mathbb {D}}\) if and only if \(\partial G\) is a closed curve, i.e., a continuous image of the unit circle \(\mathbb {T}\) (see Pommerenke 1992, Theorem 2.1). This extension (which we again denote by \(g\)) is a homeomorphism between \(\overline{\mathbb {D}}\) and \(\overline{G}\) if (and only if) \(G\) is a Jordan domain, i.e., \(\partial G\) is a Jordan curve (see Pommerenke 1992, Theorem 2.6).

In general, the conformal mapping \(g\) induces a one-to-one correspondence between the points on \(\mathbb {T}\) and certain equivalence classes of open Jordan arcs \(\gamma \) in \(G\) with terminal point \(q\) on \(\partial G\), so called prime ends. For the details we refer to  Pommerenke (1992), Chap. 2, and  Golusin (1957), Sect. 2.3.

If \(G\) contains a disk \(D\) which touches the boundary \(\partial G\) at some point \(p\in \partial D\cap \partial G\), then every Jordan arc with starting point in \(D\) and terminal point \(p\) is contained in the same equivalence class. Hence there is a well-defined prime end of G associated with p by D.

Complexes The skeleton of a circle packing is a simplicial 2-complex \(K\). Throughout this paper it is assumed that \(K\) is a combinatorial closed disk, i.e., it is finite, simply connected and has a nonempty boundary. Simply speaking of a complex, we always mean a complex of this class. Properties of complexes which are relevant in circle packing are summarized in Lemma 3.2 of Stephenson (2005).

We denote the sets of vertices, edges and faces of \(K\) by \(V,E,F\), respectively. The edge adjacent to the vertices \(u\) and \(v\) is denoted by \(e(u,v)\) or \(\langle u,v\rangle \), where the first version stands for the non-oriented edge, while the second means the oriented edge from \(u\) to \(v\). Similarly, a face of \(K\) with vertices \(u,v,w\) is written as \(f(u,v,w)\) (non-oriented) or \(\langle u,v,w\rangle \) (oriented), respectively. Two vertices \(u\) and \(v\) are said to be neighbors if they are connected by an edge \(e(u,v)\) in \(E\). For any vertex \(v\in V\) we denote by \(E(v)\) the set of edges adjacent to \(v\). This set is endowed with a natural cyclic (counterclockwise) ordering, so that for \(e_1,e_2\in E(v)\) definitions like \(\{e\in E(v): e_1<e\le e_2\}\) make sense. Any edge \(e\) of \(K\) is adjacent to one or two faces. In the first case \(e\) is a boundary edge, otherwise it is an interior edge of \(K\). Boundary vertices are those vertices of \(K\) which are adjacent to a boundary edge. The sets of boundary edges and boundary vertices are denoted \(\partial E\) and \(\partial V\), respectively, the vertices in \(V{\setminus }\partial V\) are called interior vertices. We point out that a boundary vertex can be adjacent to many other boundary vertices, and that an edge which connects two boundary vertices need not be a boundary edge (as vertex \(v\) and edge \(e\) in Fig. 3, left). We let \(B(v)\) denote the smallest sub-complex of \(K\) which contains a vertex \(v\) and all its neighbors. If \(v\) is an interior vertex \(B(v)\) is said to be the flower of \(v\) (see Fig. 3, middle), if \(v\) is a boundary vertex we speak of an incomplete flower (see Fig. 3, left). Note that \(B(v)\) need not contain all edges which connect neighbors of \(v\) (see Fig. 3, right).

Since \(K\) is a triangulation with non-void boundary, it must have at least three boundary vertices. The natural cyclic ordering of boundary edges, corresponding to the orientation of the boundary of the triangulated surface, induces a cyclic ordering of the boundary vertices. With respect to this ordering, any boundary vertex has a precursor and a successor which are well-defined.

Speaking of a chain, we mean a finite sequence \((c_1,\ldots ,c_n)\) of vertices, edges or faces, such that neighboring elements \(c_j\) and \(c_{j+1}\) are adjacent to a common edge (if the \(c_j\) are vertices or faces) or a common vertex (if the \(c_j\) are edges), respectively.

Fig. 3
figure 3

The sub-complex of a (incomplete) flower and a corresponding packing

We have illustrated some limitations of Theorem 2 in Figure 2. The reason for the observed effects is the relative independence of some substructures from the rest of the packing. This is described more precisely in the following definition.

Definition 1

Let \(K\) be a complex with a distinguished interior vertex, the alpha-vertex \(v_\alpha \). Then a vertex \(v\in V\) is called accessible (from \(v_\alpha \)) if there is a chain of vertices \((v,v_1,\ldots ,v_n,v_\alpha )\) such that \(v_1,\ldots ,v_n\) are interior vertices. The set of all accessible vertices of \(K\) is denoted by \(V^*\), the set of all edges \(e(u,v)\in E\) with \(u,v\in V^*\) by \(E^*\), and the set of all faces \(f(u,v,w)\in F\) with \(u,v,w\in V^*\) by \(F^*\). The kernel \(K^*\) of \(K\) is defined as the simplicial-2-complex arising from \(V^*,E^*,F^*\), that is \(K^*(V^*,E^*,F^*)\subset K(V,E,F)\).

Recall that a complex \(K\) is strongly connected, if the interior of \(K\) is connected, and every boundary vertex has an interior neighbor. The following lemma establishes a relation between this property and accessible vertices, and summarizes the crucial properties of the kernel. Since the statements are intuitive, we leave the details of the (somewhat tedious) proof to the reader.

Lemma 1

Let \(K\) be a complex with a distinguished interior alpha-vertex \(v_\alpha \). Then the following assertions hold:

  1. (1)

    The kernel \(K^*(V^*,E^*,F^*)\) of \(K(V,E,F)\) is a strongly connected complex with \(\partial V^*=\partial V\cap V^*\).

  2. (2)

    The complex \(K\) coincides with its kernel \(K^*\) (i.e., all vertices of \(K\) are accessible) if and only if \(K\) is strongly connected.

Circle packings A collection \(\mathcal {P}\) of open disks \(D_v\) is said to be a circle packing for the complex \(K=K(V,E,F)\) if it satisfies the following conditions (1)–(3):

  1. (1)

    Each vertex \(v\in V\) has an associated disk \(D_v\in \mathcal {P}\), such that \(\mathcal {P}=\{D_v: v\in V\}\).

  2. (2)

    If \(\langle u,v\rangle \in E\) is an edge of \(K\), then the disks \(D_u\) and \(D_v\) touch each other.

  3. (3)

    If \(\langle u,v,w\rangle \in F\) is a positively oriented face of \(K\), then the centers of the disks \(D_u, D_v, D_w\) form a positively oriented triangle in the plane.

A circle packing is called univalent, if its disks are non-overlapping, \(D_u \cap D_v=\emptyset \) for all \(u,v\in V\) with \(u \not = v\). In this paper all circle packings are assumed to be univalent.

Since the structure of the underlying complex \(K\) carries over to the associated packing \(\mathcal {P}\), all related attributes can be applied to the disks \(D_v\) as well—so we shall speak of boundary disks, interior disks, neighboring disks, etc.

The contact point of two neighboring disks \(D_u,D_v\) is defined by \(c(u,v):=\overline{D}_u\cap \overline{D}_v\). The contact points of a packing \(\mathcal {P}\) for the complex \(K(V,E,F)\) are the points \(c(u,v)\) with \(e(u,v)\in E\).

We denote by \(D\) the union of all disks in \(\mathcal {P}\), \(D:=\bigcup _{v\in V} D_v\). If \(\mathcal {P}\) is univalent and \(p\) and \(q\) are different points of \(\partial D\), there is at most one disk \(D_v\) whose boundary \(\partial D_v\) contains \(p\) and \(q\). If such a disk exists, we define \(\delta (p,q)\) as the positively oriented open subarc of \(\partial D_v\) from \(p\) to \(q\), and \(\delta [p,q]:=\overline{\delta (p,q)}\). In addition we set \(\delta (p,p):=\emptyset \) and \(\delta [p,p]:=\{p\}\). Note that \(\delta (p,q)\) and \(\delta [q,p]\) are complementary subarcs of \(\partial D_v\), provided that \(p \not =q\).

If \(\langle u,v,w\rangle \) is a face of \(K\), the interstice \(I(u,v,w)\) of \(\mathcal {P}\) is the Jordan domain bounded by the arcs \(\delta _u:=\delta \big (c(u,v),c(u,w)\big )\), \(\delta _v:=\delta \big (c(v,w),c(v,u)\big )\) and \(\delta _w:=\delta \big (c(w,u),c(w,v)\big )\) (see Fig. 4, left).

Fig. 4
figure 4

Definition of the interstice \(I:=I(u,v,w)\) and the carrier \(D^*\) of two packings

Besides the union \(D\) of all disks in a packing \(\mathcal {P}\) we need the carrier of \(\mathcal {P}\), which is the compact set

$$\begin{aligned} D^*:=\overline{D} \cup \bigcup _{f(u,v,w)\in F} I(u,v,w) \end{aligned}$$

(see Fig. 4, middle and right). Note that this definition is somewhat different from Stephenson’s (cp. Stephenson 2005, p. 58). The carrier is essential in the next definition.

Definition 2

Let \(G\) be a bounded, simply connected domain. We say that a (univalent) circle packing \(\mathcal {P}\) is contained in \(G\) (or lies in \(G\)) if the interior of \(D^*\) is a subset of \(G\). A packing \(\mathcal {P}\) contained in \(G\) is said to fill \(G\) if every boundary disk of \(\mathcal {P}\) touches \(\partial G\).

If \(G\) is a Jordan domain, \(\mathcal {P}\) is contained in \(G\) if and only if any disk of \(\mathcal {P}\) is a subset of \(G\). For general domains the latter condition alone would be too week, since then it could happen that “spikes” of \(\partial G\) (think of \(G\) as a slit disk) penetrate into the packing, sneaking through between two boundary disks at their contact point. This is prevented by our definition; in particular it guarantees that \(\partial G\cap I=\emptyset \) for every interstice \(I\) of \(\mathcal {P}\).

What happens when \(\partial G\) meets a contact point of two boundary disks is explored in the following lemma (an explanation of associated prime ends is given at the beginning of this section).

Lemma 2

Let \(G\) be a bounded, simply connected domain, and let \(\mathcal {P}\) be a circle packing contained in \(G\). Then every contact point \(c(u,v)\in \partial G\) is associated with the same prime end by both \(D_u\) and \(D_v\).

Proof

Let \(c=c(u,v)\) be a contact point of \(\mathcal {P}\) which lies on the boundary of \(G\). Then there exists a vertex \(w\in V\) such that \(f(u,v,w)\) is a face in the complex of \(\mathcal {P}\), and we denote by \(I=I(u,v,w)\) the corresponding interstice.

For \(\varepsilon >0\), let \(B_\varepsilon \) be an open disk centered at \(c\) with radius \(\varepsilon \) and define

$$\begin{aligned} \widetilde{B}_\varepsilon := B_\varepsilon \cap \big (D_u\cup D_v\cup \overline{I}\big ). \end{aligned}$$

If \(\varepsilon \) is sufficiently small, \(\widetilde{B}{_\varepsilon }{\setminus }\{c\}\) is a Jordan domain contained in \(G\), and we have \(D_u \cap B_\varepsilon \subset \widetilde{B}_\varepsilon \), \(D_v \cap B_\varepsilon \subset \widetilde{B}_\varepsilon \) (see Fig. 5, left). As a Jordan domain \(\widetilde{B}{_\varepsilon }{\setminus }\{c\}\) has a unique prime end \(c^*\) corresponding to its boundary point \(c\), so the prime ends of \(G\) associated with \(c\) by the disks \(D_u\) and \(D_v\), respectively, must coincide. \(\square \)

Fig. 5
figure 5

Definitions of \(\widetilde{B}_\varepsilon \), boundary arcs and boundary interstices

A packing which fills the unit disk \(\mathbb {D}\) is called maximal. A celebrated result, the Koebe-Andreev-Thurston-Theorem (which can be traced back to Koebe’s paper 1936), tells us that any complex \(K\) has an associated maximal packing, which is unique up to conformal automorphisms of \(\mathbb {D}\). A far reaching generalization is the Uniformization Theorem of Beardon and Stephenson (1990, see also Chap. II in Stephenson 2005). Recall that the boundary disks of a packing form a chain \(D_1,\ldots ,D_m\). Since this is a cyclic structure, we label it modulo \(m\), in particular \(D_0:=D_m\) and \(D_{m+1}:=D_1\). For \(k\in \{1,\ldots ,m\}\), we denote by \(\eta _k\) the closed segment which connects the centers of \(D_{k}\) and \(D_{k+1}\). These boundary segments form a (polygonal) Jordan curve \(\eta \).

If \(D_{k-1},D_k\) and \(D_{k+1}\) are three consecutive boundary disks, the contact points \(c_k^-:=\overline{D}_{k-1}\cap \overline{D}_k\) and \(c_k^+:=\overline{D}_{k}\cap \overline{D}_{k+1}\) split \(\partial D_k\) into two arcs. We call \(\delta (c_k^-,c_k^+)\) the exterior boundary arc and \(\delta (c_k^+,c_k^-)\) the interior boundary arc of \(D_k\), respectively (see Fig. 5, middle).

Lemma 3

Let \(D_k\) be a boundary disk of a circle packing \(\mathcal {P}\). Then the exterior boundary arc of \(D_k\) contains no contact points of disks in \(\mathcal {P}\).

Note that, by definition, the contact points of the disks in a packing are prescribed by the combinatorics of \(\mathcal {P}\).

Proof

The polygonal line \(\eta \) which connects consecutive centers of the boundary disks is a Jordan curve which separates the exterior boundary arcs from the interior boundary arcs. The interior of \(\eta \) contains the closures \(\overline{D}_v\) of all interior disks. Any contact point \(c\) of \(\mathcal {P}\) is either a contact point of two boundary disks, or it lies on the boundary of an interior disk. In both cases \(c\) does not belong to any exterior boundary arc. \(\square \)

To provide some more notation, let \(\mathcal {P}\) be a circle packing which fills a bounded, simply connected domain \(G\). By definition, every boundary disk \(D_k\) touches \(\partial G\) in a non-void (possibly uncountable) set \(G_k\) of points, and \(G_k\) must be contained in the closure \(\delta [c_k^-,c_k^+]\) of the exterior boundary arc \(\delta (c_k^-,c_k^+)\) of \(D_k\). Let \(\delta _k:=\delta [g_k^-,g_k^+]\) be the smallest subarc (we allow the possibility that this ‘arc’ degenerates to a point) of \(\delta [c_k^-,c_k^+]\) which contains \(G_k\). Since \(G_k\) is a closed set, we have \(g_k^-,g_k^+\in G_k\).

In order to define the boundary interstice \(I_k\) between two consecutive boundary disks \(D_k\) and \(D_{k+1}\) (see Fig. 5, right) we distinguish two cases. If \(g_k^+=c_k^+\), we set \(I_k:=\emptyset \). Otherwise we let \(\delta \) be the union of the arcs \(\delta (g_k^+,c_k^+]\) (a subarc of \(\partial D_k\)) and \(\delta [c_k^+,g_{k+1}^-)\) (a subarc of \(\partial D_{k+1}\)). The open Jordan arc \(\delta \) is contained in \(G\) with different endpoints on \(\partial G\), hence it is a crosscut. The set \(G{\setminus }\delta \) consists of two simply connected components \(G_1\) and \(G_2\). One of these components contains all disks of \(\mathcal {P}\), the other one is (by definition) the boundary interstice \(I_k\).

Lemma 4

\(I_k\cap D=\emptyset \) for all \(k=1,\ldots ,m\).

Proof

Let \(k\in \{1, \ldots ,m\}\) be fixed. If \(I_k=\emptyset \) the assertion is trivially fulfilled. Let \(I_k\ne \emptyset \) and let \(\delta \) be the crosscut defined above, so that \(G{\setminus }\delta \) consists of exactly two simply connected domains \(G_1=I_k\) and \(G_2\).

Clearly every disk of \(\mathcal {P}\) is contained either in \(G_1\) or \(G_2\). We assume that there is a disk \(D_u\) in \(G_1\) (remember \(D_k\subset G_2\)). Because \(K\) is connected there is a chain \(C\) of vertices \(\{u, \ldots ,v\}\), where \(v\) is the vertex associated with \(D_k\). Because \(D_u\subset G_1\) and \(D_k\subset G_2\) there have to be two consecutive vertices \(w_1,w_2\) in \(C\), so that \(D_{w_1}\) is contained in \(G_1\) and \(D_{w_2}\) in \(G_2\). The contact point \(c(w_1,w_2)\) must lie on \(\partial G_1{\setminus }\delta \), because there are no contact points of \(\mathcal {P}\) on \(\delta \) according to Lemma 3.

Let \(w_3\) be a vertex, so that \(f(w_1,w_2,w_3)\) is a face of \(K\). The interstice \(I:=I(w_1,w_2,w_3)\) is contained either in \(G_1\) or \(G_2\), because it is disjoint from \(\partial G\). Moreover both arcs \(\partial D_{w_1}\cap \partial I\) and \(\partial D_{w_2}\cap \partial I\) (up to their endpoints) lie in the same domain as \(I\), without being contained in the boundary of \(G\). This implies, that both disks \(D_{w_1}\) and \(D_{w_2}\) are contained either in \(G_1\) or \(G_2\), a contradiction. Hence, \(I_k\cap {D}=\emptyset \) for all \(k=1,\ldots ,m\). \(\square \)

Last but not least we state a result about glueing simply connected domains along a common boundary arc. The proof is left as an exercise (see Pommerenke 1992).

Lemma 5

Let \(G_1\) and \(G_2\) be simply connected domains with locally connected boundaries. If \(G_1\) and \(G_2\) touch each other along a Jordan arc \(J\) with endpoints \(a,b\), i.e., \(G_1 \cap G_2=\emptyset \) and \(\overline{G}_1\cap \overline{G}_2=J\), then \(\big (G_1\cup J \cup G_2\big ) {\setminus } \{a,b\}\) is a simply connected domain and its boundary is locally connected.

3 Crosscuts

Before we introduce crosscuts of a (univalent) circle packing which fills a domain \(G\), we define crosscuts of its complex.

Definition 3

A (combinatoric) crosscut of a complex \(K\) is a sequence \(L=(e_0,e_1,\ldots ,e_l)\) of edges in \(K\) with the following properties (1)–(4):

  1. (1)

    The edges are pairwise different, if \(0\le j < k\le l\) then \(e_j\not =e_k\).

  2. (2)

    For \(1\le j\le l\) the edges \(e_{j-1}\) and \(e_j\) are adjacent to a common face of \(K\).

  3. (3)

    Three consecutive edges are not adjacent to the same face of \(K\).

  4. (4)

    The edges \(e_0\) and \(e_l\) are boundary edges.

Fig. 6
figure 6

A crosscut \(L\) of \(K\), the vertex sets \(V_L^-\), \(V_L^+\), \(U_L^+\), and a corresponding packing

It is easy to see that only the first and the last edge of a crosscut can be boundary edges of \(K\). Because \(e_0\ne e_l\) we have \(l\ge 1\). When one edge of a face \(f\) belongs to \(L\), then \(L\) must contain exactly two edges of \(f\), and these are subsequent members of \(L\). So a crosscut can also be represented by a sequence \((f_1,\ldots ,f_l)\) of faces, where \(e_{j-1}\) and \(e_j\) are adjacent to \(f_j\). Since the three edges of a face are not allowed to be consecutive members of \(L\), all faces \(f_j\) must be pairwise different.

After removing the edges of a crosscut \(L\) from \(K\), the remaining graph consists of two edge-connected components \(K_L^-\) and \(K_L^+\). We assume that \(K_L^-\) ‘lies to the right’ and \(K_L^+\) ‘lies to the left’, respectively, when we move along the edges \(e_0,e_1,\ldots ,e_l\) in this order. The vertex sets of \(K_L^-\) and \(K_L^+\) are denoted by \(V_L^-\) and \(V_L^+\), respectively, and we call them the lower and the upper vertices of \(K\) with respect to \(L\). The set \(U_L^+\) is constituted by all vertices \(v\) in \(V_L^+\) which are adjacent to an edge in \(L\). These vertices and the corresponding disks are said to be the upper neighbors of \(L\). An analogous definition is made for the set \(U_L^-\) of lower neighbors of \(L\) (see Fig. 6).

Given a (combinatoric) crosscut \(L\) of a complex \(K\) and a circle packing \(\mathcal {P}\) for \(K\) which fills a domain \(G\), we define several related (geometric) crosscuts \(J\) of \(\mathcal {P}\) in \(G\). To begin with, we associate with every edge \(e_j=e(u,v)\) in \(L\) the contact point \(x_j:= \overline{D}_u \cap \overline{D}_v\) of the disks \(D_u,D_v\in \mathcal {P}\). The common tangent to \(D_u\) and \(D_v\) at \(x_j\) is denoted \(\tau _j\). The set \(X:=\{x_0,\ldots ,x_l\}\) of all contact points associated with edges of \(L\) has a natural ordering, induced by the ordering of edges in the crosscut. Since the indexing of the elements fits with this ordering, we write \(x_j<x_k\) if \(j<k\).

The polygonal crosscut \(J_L^0\) is built from the common tangents \(\tau _i\) of disks at their contact points \(x_i\) as follows. Let \(i\in \{1,\ldots ,l\}\) and assume that \(x_{i-1}\) and \(x_i\) are consecutive contact points of the pairs \(D_u,D_v\) and \(D_v,D_w\), respectively. Then the three circles \(\partial D_u, \partial D_v, \partial D_w\) bound an interstice \(I:=I(u,v,w)\). The tangents \(\tau _{i-1}\) and \(\tau _{i}\) intersect each other at a point \(s_i\) in \(I\), and the union of the closed segments \([s_{i},s_{i+1}]\) for \(i=1,\ldots ,l-1\) is a Jordan arc in \(G\) (see Fig. 7).

Fig. 7
figure 7

Local construction and global view of a polygonal crosscut

In order to complete this arc to a crosscut in \(G\) we look at the boundary disks \(D_k\) and \(D_{k+1}\) which touch each other at \(x_0\). If \(x_0\) is not a boundary point of \(G\) we define \(s_0\) as the endpoint of the largest segment \((x_0,s_0)\) on the tangent \(\tau _0\) which is contained in \(I_k\). Since there is no disk of \(\mathcal {P}\) intersecting \(I_k\) (Lemma 4) we see that \([x_0,s_0)\subset G\) is disjoint from \(\mathcal {P}\) and \(s_0\in \partial G\). If \(x_0\) is a boundary point of \(G\) we set \(s_0:=x_0\).

A similar construction is made for the point \(s_{l+1}\) as (“the first”) intersection point of the tangent \(\tau _l\) with \(\partial G\). Here \(x_0\ne x_{l}\) ensures that \([s_0,s_1]\) and \([s_l,s_{l+1}]\) live in two different boundary interstices. Although this does not exclude \(s_0=s_{l+1}\), it guarantees that \(s_0\) and \(s_{l+1}\) are endpoints of the segments \([s_1,s_0]\) and \([s_l,s_{l+1}]\), belonging to different prime ends \(s_0^*\) and \(s_{l+1}^*\), respectively.

Finally, the union of the closed segments \([s_{k},s_{k+1}]\) for \(k=0,\ldots ,l\) forms the desired polygonal crosscut \(J_L^0:=\bigcup _{k=0}^l [s_{k},s_{k+1}]\) in \(G\). It can easily be verified that \(J_L^0\) is a (topologically closed) Jordan arc which meets \(\overline{D}\) at the contact points \(x_k\) – more precisely we have \(X\subset J_L^0\cap \overline{D}\subset X\cup \{s_0,s_{l+1}\}\). The open set \(G{\setminus } J_L^0\) has two simply connected components \(G^+_0\) and \(G^-_0\), containing the disks associated with \(V^+_L\) and \(V^-_L\), respectively.

It is clear that, for a fixed combinatorial crosscut \(L\) of \(K\), the statement of Theorem 2 depends on the choice of the geometric crosscut \(J\): the assertion becomes the stronger, the larger the domain \(G_J^-\) is. Unfortunately, there exists (in general) no crosscut \(J\) which maximizes \(G_J^-\), since the boundary of the largest domain \(G_J^-\) need not be a Jordan curve. We therefore extend the concept of crosscuts somewhat, defining the maximal crosscut \(J_L^+\) in \(\mathcal {P}\) as follows.

Fig. 8
figure 8

Construction of a maximal crosscut (which is not a Jordan arc)

Recall that \(U_L^+\) is the vertex set of upper neighbors of \(L\). If \(x_{k-1}\) and \(x_k\) are contact points of the disks \(D_u,D_v\) and \(D_v,D_w\), respectively, then either \(v\in U_L^+\) or \(u,w\in U_L^+\). The interstice \(I(u,v,w)\) is bounded by three (topologically closed) circular arcs \(\alpha _u\), \(\alpha _v\) and \(\alpha _w\), respectively. If \(v\in U_L^+\) we connect \(x_{k-1}\) with \(x_{k}\) by the arc \(a_k:=\alpha _v\), in the second case we connect these points by the concatenation \(a_k:=\alpha _u\cup \alpha _w\) (see Fig. 8). In addition we connect \(x_0\) and \(x_l\) with \(\partial G\) by arcs \(a_0:=\delta (g_j^+,x_0)\) and \(a_{l+1}:=\delta (x_l,g_k^-)\) of those circles \(\partial D_j\) and \(\partial D_k\) which are upper neighbors of \(L\) and contain \(x_0\) and \(x_l\), respectively. The union \(J_L^+:=\bigcup _{k=0}^{l+1} a_k\) of these arcs is a curve which we call the maximal crosscut in \(\mathcal {P}\) with respect to \(L\).

The maximal crosscut \(J_L^+\) is composed from a finite number of circular (topologically closed) arcs \(\omega _i\) which are linked at the turning points \(t_i\) of \(J_L^+\), and every contact point \(x_k\) lies exactly on one arc \(\omega _i\) (see Fig. 8). If \(J_L^+\) is not a Jordan arc, \(G {\setminus } J_L^+\) may consist of several connected components (see Fig. 8, right), one of them containing all disks associated with vertices \(v\) in \(V_L^-\). We call this component \(G_L^-\) the maximal lower domain for \(L\) with respect to \(\mathcal {P}\), and we set \(G_L^+:=G {\setminus } \overline{G_L^-}\). For the sake of brevity we define \(\omega :=J_L^+\) and \(\Omega :=G_L^-\).

Since the curve \(\omega \) can have multiple points (see Fig. 8, right) there is no natural ordering of the points on \(\omega \). However, considering \(\omega \) as part of the boundary of \(\Omega \), we can introduce an ordering of the terminal points \(q\in \omega \) of open Jordan arcs \(\gamma (p,q)\) in \(\Omega \). In order to describe this procedure we need the following result.

Lemma 6

For any combinatorial crosscut \(L\) the maximal lower domain \(\Omega =G_L^-\) is simply connected and has a locally connected boundary.

Proof

Let \(G_0^-\) be the lower domain with respect to the polygonal crosscut \(J_0\) in \(\mathcal {P}\). Then \(G {\setminus } J_L^0\) consists of two simply connected domains \(G_0^-\) and \(G_0^+\), respectively. The maximal lower domain \(G_L^-\) is constructed by glueing a finite number of simply connected domains along straight line segments to \(G_0^-\). Hence the assertion follows from Lemma 5. \(\square \)

The assertion of Lemma 6 guarantees that any (fixed) conformal mapping \(g: \mathbb {D}\rightarrow \Omega \) has a continuous extension to \(\overline{\mathbb {D}}\), which we again denote by \(g\) (see Pommerenke 1992, Theorem 2.1). With respect to this mapping, we let \(\sigma _i\subset \mathbb {T}\) denote the preimage of the circular arcs \(\omega _i\) with \(i=1,\ldots ,n\). Then \(\sigma :=\bigcup _{i=1}^n\sigma _i\) is the preimage of the maximal crosscut \(\omega \).

By the Prime End Theorem, the mapping \(g\) induces a bijection \(g^*\) between \(\mathbb {T}\) the set of prime ends of \(\Omega \). We denote by \(\omega ^*:=g^*(\sigma )\) the set of prime ends associated with \(\Omega \), and, for \(i=1,\ldots ,n\), we let \(\omega _i^*:=g^*(\sigma _i)\) be the subsets of \(\omega ^*\) corresponding to the arcs \(\sigma _i\).

Note that the preimages \(\sigma _i\) of the circular arcs \(\omega _i\) are topologically closed subarcs of \(\mathbb {T}\), and that the preimage \(\mathbb {T}{\setminus }\sigma \) of \(\partial \Omega {\setminus }\omega \) is not empty. Therefore \(\sigma _i\) and \(\sigma _j\), and thus \(\omega _i^*\) and \(\omega _j^*\), are disjoint if \(|i-j|>1\), while their intersection contains exactly one element if \(|i-j|=1\).

Further we see that the arcs \(\sigma _1, \sigma _2, \ldots , \sigma _n\) (in this order) are arranged in clockwise direction on \(\mathbb {T}\). It is therefore just natural to order the points on the arc \(\sigma \) (and hence on each subarc \(\sigma _i\)) also in clockwise direction. The mapping \(g^*\) transplants this ordering from \(\sigma \) to the set \(\omega ^*\) of prime ends. If \(\gamma _1^*=g^*(s_1)\) and \(\gamma _2^*=g^*(s_2)\) are two prime ends of \(\omega ^*\), the notion \(\gamma _1^*\le \gamma _2^*\) refers to the ordering \(s_1\le s_2\) of the associated points on \(\sigma \).

Remark Every \(\omega _i\) without its endpoints is an open Jordan arc, so there is a one-to-one correspondence between the interior points of \(\omega _i\) and \(\sigma _i\). Let \(\gamma \) in \(\Omega \) be an open Jordan arc with terminal point \(q\) on \(\omega \), then the associated unique prime end \(\gamma ^*\) in \(\omega ^*\) must lie in \(\omega ^*_i\), whenever \(q\) is an interior point of \(\omega _i\). Only if \(q\) is an endpoint of \(\omega _i\) there is a chance that the prime end \(\gamma ^*\) is not contained in \(\omega ^*_i\), because now \(\gamma ^*\) depends on how \(\gamma \) approaches \(q\).

4 Loners

So far we have studied properties of a single circle packing \(\mathcal {P}\) which fills \(G\). In the next step we consider pairs \((\mathcal {P},\mathcal {P}')\) of packings which are subject to the assumptions of Theorem 2.

Definition 4

A pair \((\mathcal {P},\mathcal {P}')\) of univalent circle packings for the complex \(K\) is said to be admissible (for the crosscut \(L\) of \(K\) in \(G\) with alpha-vertex \(v_\alpha \)) if it satisfies the following conditions:

  1. (1)

    The packing \(\mathcal {P}\) fills the bounded, simply connected domain \(G\), and the packing \(\mathcal {P}'\) is contained in \(G\) (see Definition 2).

  2. (2)

    For all vertices \(v\in U_L^-\) (the lower neighbors of \(L\)) the disks \(D'_v\) are contained in \(G_L^-\) (the maximal lower domain of \(G\) for \(L\) with respect to \(\mathcal {P}\)).

  3. (3)

    The centers of the alpha-disks of \(\mathcal {P}\) and \(\mathcal {P}'\) coincide and lie in \(G_L^+:=G {\setminus } G_L^-\).

Though it would be more precise to speak of an admissible sixtuple \((K,L,G,\mathcal {P},\mathcal {P}',v_\alpha )\), we shall use the term “admissible” generously, for instance saying that “\(L\) is an admissible crosscut for \((\mathcal {P},\mathcal {P}')\)”.

Recall that \(U_L^+\) denotes the vertex set of those disks in \(\mathcal {P}\) which lie in \(G_L^+\) and touch the crosscut (“upper neighbors of \(L\)”). In the next step we are going to explore the interplay of the disks \(D_v'\) in \(\mathcal {P}'\) and \(D_w\) in \(\mathcal {P}\) for \(v,w\in U_L^+\).

Definition 5

Let \((\mathcal {P},\mathcal {P}')\) be an admissible pair of circle packings for the complex \(K\) with crosscut \(L\). A vertex \(v\) in \(U_L^+\) is called a loner, if \(D_v'\cap D_w= \emptyset \) for all \(w\in U_L^+\) with \(w \not =v\).

The concept of loners was introduced by  Schramm (1991) in a similar but somewhat different context. The main characteristic of a loner is the following.

Lemma 7

Let \(v\) in \(U_L^+\) be a loner of the admissible pair \((\mathcal {P},\mathcal {P}')\) with complex \(K\) and crosscut \(L\). Then \(D'_v\cap (G^+_L{\setminus } D_v)=\emptyset \).

Proof

Let \(u\in U^-_L\) and \(w\in U^+_L\) be neighbors of \(v\), and let \(p\) and \(q\) be the contact points of the disks \(D'_v\) with \(D'_u\) and \(D_v\) with \(D_w\), respectively. Clearly \(p\ne q\), otherwise \(D'_u\) had to intersect \(D_v\) or \(D_w\), a contradiction to condition (2) of the admissible pair \((\mathcal {P},\mathcal {P}')\).

Assume that \(p\) is a boundary point of \(D_v\). Then \(\partial D_v\) and \(\partial D'_v\) have a common tangent at \(p\), otherwise \(D'_u\) had to intersect \(D_v\), a contradiction to condition (2) of the admissible pair \((\mathcal {P},\mathcal {P}')\). It follows that either \(\overline{D'_v}{\setminus }\{p\}\subset D_v\) or \(D'_v=D_v\) or \(\overline{D_v}{\setminus }\{p\}\subset D'_v\). The latter implies that \(q\in D'_v\), hence \(D'_v\cap D_w\ne \emptyset \), which is impossible since \(v\) is a loner. The other two cases imply the statement we want to prove.

Assume that \(p\) is not a boundary point of \(D_v\). Suppose that the assertion of Lemma 7 were false, i.e., there is some point \(r\) in \(D'_v\) which is also contained in \(G^+_L{\setminus } D_v\). Because \(p\) lies in the maximal lower domain \(G^-_L\), and \(r\) lies in the upper domain \(G^+_L\), the boundary of \(D'_v\) must intersect the maximal crosscut \(J^+_L\). Since the vertex \(v\) is a loner, every such intersection point must lie in \(\partial D_v\). If \(\partial D'_v\cap \partial D_v\) consists of exactly one point \(r_1\), then the boundary of \(D'_v\) is the union of \(\delta [p,r_1]\) and \(\delta [r_2,p]\), hence \(D'_v\cap G^+_L=\emptyset \), a contradiction to \(r\in D'_v\). If there is a second point \(r_2\in \partial D'_v\cap \partial D_v\) with \(r_1\ne r_2\), then we have \(\partial D'_v\cap D_v=\delta (r_2,r_1)\), hence \(r\) must be contained in \(D_v\), a contradiction to \(r\in G^+_L{\setminus } D_v\). \(\square \)

In Sect. 6 the property of loners described in Lemma 7 will allow us to move the crosscut \(L\) through the packing, reducing in every step the number of disks in \(G^+_L\). The next result is crucial for the applicability of this procedure.

Lemma 8

(Existence of loners) Every admissible pair \((\mathcal {P},\mathcal {P}')\) of circle packings with crosscut \(L\) has a loner.

The proof is divided into several steps; the first part uses the geometry of disks, then we employ some topology, and finally everything is reduced to pure combinatorics. We start with some preparations.

Recall the definition of the contact points \(x_k\): If \(L=(e_0,\ldots ,e_l)\) and \(e_k=\langle u,v\rangle \), for some \(k\in \{0,\ldots ,l\}\), then \(x_k:= \overline{D_u}\cap \overline{D_v}\). Using the same notation, the corresponding contact points of disks in \(\mathcal {P}'\) are given by \(y_k:= \overline{D'_u} \cap \overline{D'_v}\), where \(Y:=\{y_0,\ldots ,y_l\}\) is the set of all such contact points.

The contact points \(x_k\) form an ordered set on the maximal crosscut \(\omega :=J_L^+\), which is the upper boundary of the maximal lower domain \(\Omega :=G_L^-\). Since every \(x_k\) lies on exactly one arc \(\omega _i\), the set \(X\) of contact points splits into classes \(X_i:=\{x_k\in X: x_k\in \omega _i\}\), \(i=1,\ldots ,n\). The set \(Y\) of the contact points of \(\mathcal {P}'\) is divided accordingly, \(Y_i:=\{y_k\in Y: x_k\in \omega _i\}\) (the \(x_k\) is no typo here). Like \(X\), the set \(Y\) is endowed with a natural ordering, we write \(y_j<y_k\) if \(j<k\).

Our next aim is to construct a Jordan arc \(\alpha \) which is contained in \(\overline{\Omega }\) and carries the contact points \(y_k\) in their natural order.

Fig. 9
figure 9

Construction of the Jordan arc \(\alpha \) in Case 1 (left) and Case 2 (middle, right)

Lemma 9

If \((\mathcal {P},\mathcal {P}')\) is an admissible pair, then there exist oriented Jordan arcs \(\alpha _k\) from \(y_{k-1}\) to \(y_k\) such that \(\alpha := \cup _{k=1,\ldots l} \alpha _k\) is a Jordan arc in \(\overline{\Omega }\) and \(\alpha \cap \omega \subset Y\).

Proof

Let \(k\in \{1,\ldots ,l\}\). In order to determine the arc \(\alpha _k\) of \(\alpha \) which connects \(y_{k-1}\) with \(y_k\) we remark that both points lie on the boundary of one and the same disk \(D_v'\in \mathcal {P}'\). We distinguish two cases:

Case 1 If \(v\in V_L^-\), then the disk \(D_v'\) is contained in \(\Omega \), and we choose the segment \(\alpha _k:=[y_{k-1},y_k]\) (see Fig. 9, left).

Case 2 If \(v\in V_L^+\), then \(e_{k-1},e_k\) and a third edge \(\langle u,w\rangle \) of \(K\) form a face of \(K\), and the (neighboring) disks \(D_u'\) and \(D_w'\) are both contained in \({\Omega }\). So we let \(z_k:=\overline{D_u'}\cap \overline{D_w'}\) and connect \(y_{k-1}\) with \(y_k\) by \([y_{k-1},z_k]\cup [z_k,y_k]\subset \overline{\Omega }\) (see Fig. 9, middle).

It is clear that all open segments \((y_{k-1},y_k)\), \((y_{k-1},z_k)\), \((z_k,y_k)\) for \(k=1,\ldots ,l\) are pairwise disjoint, and that \(y_k \not =z_j\). However, it is possible that two endpoints \(z_k\) and \(z_j\) coincide for \(j\not =k\), in which case the concatenation of the arcs \(\alpha _k\) is not a Jordan arc.

If this happens, the point \(z:=z_j=z_k\) is the contact point of two disks \(D_u'\) and \(D_w'\) with \(u,w\in V_L^-\). A little thought shows that then \(z\) can neither lie on the boundary of \(G\) nor on \(\omega \), and hence it must be an interior point of \(\Omega \). This allows one to resolve the double point of \(\alpha \) at \(z\) without destroying its other properties (see Fig. 9, right).\(\square \)

In the next step we transform the existence of loners to a topological problem. Technically this is much simpler when \(\alpha \) and \(\omega \) are disjoint. We consider this ‘regular case’ in Sect. 4.1. The ‘critical case’, where intersections of \(\alpha \) and \(\omega \) are admitted, will be treated in Sect. 4.2.

4.1 The regular case

Here we assume that \(\alpha \cap \omega =\emptyset \), which implies that all contact points \(y_k\) (\(k=0,\ldots ,l)\) lie in the lower domain \(\Omega \).

We fix \(i\in \{1,\ldots ,n\}\) and denote by \(y_i^-\) and \(y_i^+\) the smallest and the largest member of \(Y_i\) with respect to the natural ordering of \(Y\), respectively. Both points (which may coincide), as well as all elements of \(Y_i\), lie on the same circle \(\partial D_v'\), associated with a vertex \(v=v(i)\in V\).

Let \(\delta _i'\) be the negatively oriented topologically closed subarc of \(\partial D_v'\) from \(y_i^-\) to \(y_i^+\). We consider the largest subarcs \(\nu _i\) and \(\pi _i\) of \(\delta _i'\) which are contained in \(\overline{\Omega }{\setminus } \omega \) and have initial points \(y_i^-\) (for \(\eta _i\)) and \(y_i^+\) (for \(\pi _i\)), respectively (see Fig. 10).

Fig. 10
figure 10

The arcs \(\nu _i\) and \(\pi _i\) and their intersection with the boundary of \(G_L^+\)

Lemma 10

If there exists no loner, then the terminal points \(\nu _i^+\) and \(\pi _i^+\) of \(\nu _i\) and \(\pi _i\), respectively, lie on \(\omega \) for \(i=1,\ldots ,n\).

Proof

If one of the arcs \(\nu _i\) or \(\pi _i\) does not intersect \(\omega \), then both coincide with \(\delta _i'\). In this case, the disk \(D_{v(i)}'\) is separated from \(G_L^+\) by the union of the arcs \(\alpha \) and \(\delta _i'\), which implies that \(D_{v(i)}'\) cannot intersect any disk \(D_w\) with \(w\in U_L^+\), so that \(v(i)\) is a loner. \(\square \)

Since (with the exception of their endpoints) the circular arcs \(\nu _i\) (\(i=2,\ldots , n\)) and \(\pi _i\) (\(i=1,\ldots , n-1\)) lie in \(\Omega \) and have terminal points \(\nu _i^+\) and \(\pi _i^+\) on \(\omega \), they define prime ends \(\nu _i^*\) and \(\pi _i^*\) in \(\omega ^*\).

Because the arcs \(\nu _1\) and \(\pi _n\) need not lie in \(\Omega \), a modified definition is needed for the prime ends \(\nu _1^*\) and \(\pi _n^*\). To do so, we replace \(\nu _1\) and \(\pi _n\) by slightly perturbed circular arcs \(\nu _1^\varepsilon \) and \(\pi _n^\varepsilon \), respectively, which have the same endpoints as \(\nu _1\) and \(\pi _n\), respectively, and lie in \(\Omega \) (with the exception of their endpoints). Then \(\nu _1^*\) and \(\pi _n^*\) are defined as the prime ends associated with the terminal points of \(\nu _1^\varepsilon \) and \(\pi _n^\varepsilon \), respectively. Clearly such arcs \(\nu _1^\varepsilon \) and \(\pi _n^\varepsilon \) exist, and for all sufficiently small \(\varepsilon \) they define the same prime ends \(\nu _1^*,\pi _n^*\in \omega ^*\), respectively.

Since the set of prime ends \(\omega ^*\) is endowed with a natural ordering, we can compare the prime ends \(\nu _i^*\) and \(\pi _i^*\).

Lemma 11

If \((\mathcal {P},\mathcal {P}')\) has no loner, the prime ends \(\nu _i^*\) and \(\pi _i^*\) form an interlacing sequence with respect to the prime end ordering of \(\omega ^*\),

$$\begin{aligned} \nu _1^* \le \pi _1^* \le \nu _2^* \le \pi _2^* \le \cdots \le \nu _n^* \le \pi _n^*. \end{aligned}$$

Proof

Let \(y_-:=y_0\) and \(z_-\) be the initial and terminal points of \(\nu _1\), while \(y_+:=y_l\) and \(z_+\) are the initial and terminal points of \(\pi _n\), respectively. We have \(z_-,z_+\in \omega \) due to Lemma 10. Further, let \(\omega ^*_0\) be the set of all prime ends \(\gamma ^*\) of \(\omega ^*\) with \(\nu _1^*\le \gamma ^*\le \pi _n^*\), and denote the set of all corresponding points on \(\omega \) by \(\omega _0\). The set \(\omega _0\) is a curve or a single point. Together with the Jordan arcs \(\nu _1\), \(\alpha \) and \(\pi _n\) it forms the boundary of a simply connected domain \(\Omega _0\subset \Omega \) with locally connected boundary. Let \(\Omega _0^*\) be the set of all prime ends associated with points on \(\partial \Omega _0\). Because \(\Omega _0{\setminus }\omega _0\) is an open Jordan arc, the points \(y_-,y_+\) are associated with uniquely determined prime ends \(y_-^*,y_+^*\) of \(\Omega _0\).

Contrary to this, the points \(z_-,z_+\) may be associated with several prime ends of \(\Omega _0\). In order to explain which one we choose, let again \(\nu _1^\varepsilon ,\pi _n^\varepsilon \) be small perturbations (as explained above) of \(\nu _1,\pi _n\), respectively, so that both arcs are crosscuts in \(\Omega _0\). We define \(z_-^*\) and \(z_+^*\) as the prime ends in \(\omega ^*\) associated with the terminal points \(z_-\) and \(z_+\) of \(\nu _1^\varepsilon ,\pi _n^\varepsilon \), respectively.

We have \(n>1\), because otherwise a loner would exist. It follows that \(y_-\ne y_+\), so \(y_-^*\ne y_+^*\). From \(\alpha \cap \omega =\emptyset \) we get \(z_-,z_+\notin \{y_-,y_+\}\), hence \(z_-^*,z_+^*\notin \{y_-^*,y_+^*\}\).

If \(z_-^*=z_+^*=:z^*\), we directly get \(\omega ^*\cap \Omega _0^*=z^*\). This implies \(\nu _1^*=\pi _1^*=\nu _2^*=\cdots =\pi _n^*=z^*\), so the lemma holds. (We consider this case here, though Lemma 12 shows, that it cannot occur).

If \(z_-^*\ne z_+^*\), the prime ends \(y_-^*\), \(y_+^*\), \(z_-^*\) and \(z_+^*\) are pairwise distinct and with respect to the (cyclic) ordering of \(\Omega _0\) we have \(y_-^*<y_+^*<z_-^*<z_+^*<y_-^*\). Therefore \(\Omega _0\) can be mapped conformally onto a rectangle \(Q\) (with appropriately chosen aspect ratio) such that \(y_-^*,y_+^*,z_-^*\) and \(z_+^*\) correspond to the four corners of \(Q\) (see Pommerenke 1992), which is depicted in Fig. 11.

Any of the arcs \(\nu _i\) (\(i=2,\ldots ,n)\) and \(\pi _i\) (\(i=1,\ldots ,n-1\)) is mapped onto a crosscut of \(Q\) which connects two opposite sides of this rectangle. Since these Jordan arcs cannot cross each other in the interior of \(Q\), the ordering of their initial points on one side of \(Q\) is transplanted to the ordering of their terminal points on the opposite side of \(Q\). Translated back to \(\Omega _0\), this implies that the ordering of the prime ends \(\nu _i^*\) and \(\pi _i^*\) is the same as the ordering of the initial points \(y_i^-\) and \(y_i^+\) of \(\nu _i\) and \(\pi _i\), respectively, along the Jordan curve \(\alpha \). By construction, the latter points form an interlacing sequence. \(\square \)

Fig. 11
figure 11

Construction of \(\Omega _0\) and \(Q\) from \(\omega ,\alpha \) and \(\nu _1,\pi _n\)

Lemma 12

If both prime ends \(\nu _i^*\) and \(\pi _i^*\) belong to \(\omega _i^*\), then the corresponding vertex \(v(i)\) is a loner.

Proof

Let \(v:=v(i)\). It follows from \(\nu _i^*,\pi _i^*\in \omega _i^*\) that \(\nu _i^+,\pi _i^+\in \omega _i \subset \partial D_v\). If \(\pi _i^+\ne \nu _i^+\), the positively oriented open subarc \(\delta _i''\) of \(P'_v\) from \(\pi _i^+\) to \(\nu _i^+\) lies in \(D_v\). If \(\pi _i^+=\nu _i^+\), we set \(\delta _i'':=\emptyset \). In both cases the union of \(\alpha _i,\pi _i,\delta _i''\) and \(\nu _i\) is a Jordan curve which does not intersect the disks \(D_u\) with \(u\in U_L^+\) and \(u \not =v\). So either \(D_v'\) is disjoint to all such disks \(D_u\), or one of the disks \(D_u\) is contained in \(D_v'\). In the latter case the prime ends \(\nu _i^*\) and \(\pi _i^*\) cannot both belong to the same set \(\omega _i^*\). \(\square \)

Proof of Lemma 8

After these preparations we are ready to harvest the fruits: Assume that \((\mathcal {P},\mathcal {P}')\) has no loner. Then, by Lemma 10, the endpoint \(\nu _i^+\) of the arc \(\nu _i\) must lie on \(\omega \) and hence \(\nu _i\) is associated with a prime end \(\nu _i^*\in \omega ^*\). If \(\nu _i^*\in \omega _k^*\), we choose the smallest such \(k\) and set \(l(i):=k\). Similarly, we denote by \(r(i)\) the smallest number \(k\) for which \(\pi _i^*\in \omega _k^*\).

Lemma 11 tells us that \(l(i)\le r(i)\le l(i+1)\). In conjunction with Lemma 12 we conclude that the first condition implies \(r(i)\ge l(i)+1\). Starting with \(l(1)\ge 1\), we get inductively that \(r(i)\ge i+1\) for \(i=1,\ldots ,n\), ending up with the contradiction \(r(n)\ge n+1\). This proves Lemma 8 in the regular case. \(\square \)

4.2 The critical case

The second case, where we admit that \(\alpha \cap \omega \not =\emptyset \), will be reduced to the regular case by an appropriate deformation of the Jordan arc \(\alpha \).

Definition 6

A contact point \(y\in Y\) is called regular if \(y\notin \omega \), otherwise it is said to be critical.

If \(y\in Y\) is a critical contact point, then \(y\in \alpha \cap \omega \not =\emptyset \), and hence \(y\in \omega _j\) for some \(j\). Since \(y=\partial {D_u'} \cap \partial {D_v'}\) with some \(u\in U_L^-\) and \(v=v(i)\in U_L^+\), we see that \(y\) cannot be an endpoint of \(\omega _j\) (turning point of \(\omega \)) – otherwise \(D_u'\) would not be contained in \(\Omega \). Moreover, the circles \(\partial D_u'\), \(\partial D_v'\), and \(\omega _j\) must be mutually tangent at \(y\). The arc \(\omega _j\) is a subset of the circle \(\partial D_w\) (with \(w=v(j)\in U_L^+\)). Hence either \(D'_v\subset D_w\) (with \(D'_v=D_w\) admitted) or \(D_w\) is a proper subset of \(D'_v\).

In the next step we modify the Jordan arc \(\alpha \) in a neighborhood of \(y\) and redefine the arcs \(\nu _i\) and \(\pi _i\) (connecting \(y\) with \(\omega \)) introduced in the regular case.

Fig. 12
figure 12

Modification of \(\alpha \) and definition of the arcs \(\nu _i\) and \(\pi _i\) for critical contact points \(y\)

Let \(\varepsilon \) be a sufficiently small positive number. Denote by \(z\) the \(\varepsilon \)-shift of \(y\) in the direction of the center of \(D_u'\). Append to \(D_v'\) an equilateral open triangular domain \(T\) with one vertex at \(z\), two vertices on \(\partial D_v'\), and symmetry axis through \(y\) and \(z\) (see Fig. 12).

For \(y\notin \{y_0,y_l\}\) let \(\nu _i\) (and \(\pi _i\)) be the largest negatively (positively) oriented subarc of \(\partial (D_v'\cup T)\) which has initial point \(z\) and is contained in \(\Omega \). For \(y\in \{y_0,y_l\}\) (and only then) it can happen that \(y\) is a boundary point of \(G\). Therefore we define \(\nu _i:=[z,y]\) in the case \(y=y_0\), and \(\pi _i:=[y,z]\) in the case \(y=y_l\). The case \(y_0=y_l\) can never occur, because \(l\ge 1\).

Denote by \(\nu _i^+\) and \(\pi _i^+\) the terminal points of \(\nu _i\) and \(\pi _i\). Clearly, \(\nu _i^+,\pi _i^+\in \omega \), so let \(\nu _i^*, \pi _i^*\in \omega ^*\) be their associated prime ends.

We see, that the statement of Lemma 10 holds in the critical case, too. Moreover, for the critical case, Lemma 11 can be proved in exactly the same way as for the regular case, we just have to apply the adapted definitions of \(\nu _i^*\) and \(\pi _i^*\). All what is missing is the following “critical” version of Lemma 12.

Lemma 13

Assume that \(\partial D_v'\) with \(v=v(i)\in U_L^+\) contains a critical contact point \(y\in Y\cap \omega \). Then \(v\) is a loner if and only if \(\nu _i^*\) and \(\pi _i^*\) belong to \(\omega _i^*\).

Proof

We use the notations introduced above, with \(\varepsilon >0\) fixed and sufficiently small. We distinguish two cases.

Case 1 Let \(D_v'\subset D_w\) (see Fig. 12, left). Then \(v\) is a loner if and only if \(w=v\), and this holds, if and only if \(j=i\) and \(\nu _i^*,\pi _i^*\in \omega _i^*\).

Case 2 Let \(D_w\subset D_v'\) and \(D_w \not = D_v'\) (see Fig. 12, right). Then \(D_v'\) intersects at least two “upper” disks (namely \(D_w\) and one of its neighbors), so that \(v\) is not a loner. According to our construction, we have \(\nu _i^*\le y^*\le \pi _i^*\) (where \(y^*\in \omega _j^*\) is the prime end corresponding to \(y\) and \(w=v(j)\)), but both equalities are never fulfilled at the same time, and \(\nu _i^*,\pi _i^*\notin \omega _j^*\) for \(w=v(j)\). Therefore \(\nu _i^*\in \omega _m^*\) and \(\pi _i^*\in \omega _n^*\) with \(m\le j\le n\), but \(m<n\), so the prime ends \(\nu _i^*\) and \(\pi _i^*\) cannot both belong to the same class \(\omega _i^*\). \(\square \)

Remark If \(D_v'\) has several critical contact points \(y\in Y\cap \omega _j\) with the same arc \(\omega _j\), then \(D_v'\) must be tangent to \(D_w\) with \(w=v(j)\) at two different points. This implies that \(D_v'=D_w\), which explains why the criterion is independent of the choice of \(y\).

After replacing all critical contact points \(y_k\) by the shifted points \(z_k\), and modifying the construction of the curve \(\alpha \) accordingly, Lemma 8 can be proved completely the same way as in the regular case.

In Sect. 5 we need the following generalization of Lemma 8. We point out that \(v(i)=v(j)\) is allowed in assertion (i).

Lemma 14

Let \(D_{v(i)}=D_{v(i)}'\) and \(D_{v(j)}=D_{v(j)}'\) with \(1\le i\le j\le n\). Then, in each of the following cases (1)–(3), there exists a loner \(v(k)\) which is different from \(v(i)\) and \(v(j)\), such that \(k\) satisfies the following conditions:

  1. (1)

    if \(1\le i<j-1\le n-1\), then \(i<k<j\),

  2. (2)

    if \(i>1\), then \(1\le k<i\),

  3. (3)

    if \(j<n\), then \(j<k\le n\).

Proof

The proof differs only slightly from the proof of Lemma 8. For example, in order to prove (1) we need only replace the first inequality \(l(1)\ge 1\) by \(l(i+1)\ge i+1\) (which follows from \(D_{v(i)}=D_{v(i)}'\)) and, assuming that no loner \(v(k)\) with \(i<k<j\) exists, proceed inductively for \(k=i+1,\ldots ,j\) until we arrive at \(r(j)\ge j+1\). The last condition contradicts \(D_{v(j)}=D_{v(j)}'\).

If \(v(k)=v(i)\) or \(v(k)=v(j)\), we repeat the procedure, replacing \(i\) (in the first case) or \(j\) (in the second case) by \(k\), respectively. Iterating this a number of times, if necessary, we eventually find a loner \(v(k)\) which is different from \(v(i)\) and \(v(j)\), because for all \(m=2,3, \ldots ,n-1\) we have \(v(m-1)\ne v(m)\) and \(v(m)\ne v(m+1)\). \(\square \)

5 Structure of upper neighbors

In this section we analyze the structure of the set of upper neighbors \(U_L^+\) and its subset of loners in more detail.

Two consecutive (non-oriented) edges \(e_{j-1}\) and \(e_j\) of \(L=(e_0,\ldots ,e_l)\) can be represented as \(e_{j-1}=e(u,v)\) and \(e_j=e(v,w)\). The third edge of the face \(f(u,v,w)\) is considered as oriented from \(u\) to \(w\), and we set \(e_j^0:=\langle u,w\rangle \). The set of edges \(e_j^0\) splits into two classes. We define \(E_L^-\) as the set of those \(e_j^0\) where the face \(\langle u,v,w\rangle \) is oriented clockwise, whereas \(E_L^+\) consists of those edges with counter-clockwise orientation of \(\langle u,v,w\rangle \), respectively. After renumbering the elements of \(E_L^-\) and \(E_L^+\), without changing their order, we get two sequences of oriented edges \(E_L^-=\{e_1^-,\ldots ,e_p^-\}\) and \(E_L^+=\{e_1^+,\ldots ,e_q^+\}\) (with \(p+q=l\)), which are called the sequences of lower and upper accompanying edges of the crosscut \(L\), respectively.

Here are some basic properties of \(E^-_L,E^+_L\), which follow quite easy from the definition of \(L\) (proofs are left as exercises). The oriented edges in \(E_L^-\cup E_L^+\) are pairwise disjoint; the corresponding non-oriented edges can appear at most twice, and either both in \(E^-_L\) or both in \(E^+_L\). Two consecutive edges \(e_{j-1}^\pm \) and \(e_j^\pm \) are linked at a common vertex. The vertex set of all edges in \(E_L^+\) is precisely the set \(U_L^+\) of upper neighbors of \(L\).

Figure 13 shows two examples. The involved crosscut on the right models the fourth generation of the Hilbert curve. With the exception of boundary edges, all edges in \(E_L^-\) (lighter color) and in \(E_L^+\) (darker color) appear with both orientations (not shown in the picture).

Fig. 13
figure 13

The upper and the lower accompanying edges of a crosscut

When we arrange the elements of \(U_L^+\) in the order they are met along the edge path \(E_L^+\) we get the sequence \(S_L^+\) of upper accompanying vertices. A similar definition is made for the sequence \(S_L^-\) of lower accompanying vertices. The geometry of circle packings causes some combinatorial obstructions for these sequences.

Lemma 15

The sequence \(S_L^+\) of upper accompanying vertices cannot contain the pattern \((\ldots ,u,\ldots ,v,\ldots ,u,\ldots , v, \ldots )\) with \(u \not =v\).

Proof

If the sequence \(S_L^+\) contains the pattern \((\ldots ,u,\ldots ,v,\ldots ,u,\ldots )\), the oriented curve \(\omega \) has three subarcs \(\omega _i,\omega _j,\omega _k\) with \(i<j<k\) such that \(\omega _i, \omega _k \subset \partial D_u\) and \(\omega _j \subset \partial D_v\). But then \(\omega \) cannot contain a subarc of \(\partial D_v {\setminus } \omega _j\) (see Fig. 14, left), which would be necessary to append another \(v\) to the sequence. \(\square \)

Fig. 14
figure 14

Illustrations to Lemmas 15 and  17

Definition 7

A vertex \(v\in U_L^+\) which appears only once in the sequence \(S_L^+\) is called simple, the other elements in \(U_L^+\) are said to be multiple vertices.

If \(v\) is a multiple vertex in \(U_L^+\), there are sequences \(M:=\{e_i^+,e_{i+1}^+,\ldots ,e_j^+\}\subset E_L^+\) of accompanying edges such that \(v\) is the initial vertex of \(e_i^+\), as well as the terminal vertex of \(e_j ^+\) with \(i<j\). Any such sequence is called a loop for \(v\). We say that a loop \(M\) meets a vertex \(u\in U_L^+\), if \(u\) is adjacent to an edge in \(M\) and \(u\ne v\). The set of vertices met by \(M\) is denoted by \(V_M\). A loop \(M\) also generates a sequence of vertices \(U_M=\{v,v_1,\ldots ,v_m,v\}\) when we arrange the elements of \(V_M\) in the order they are met along the edge path \(M\).

Lemma 16

Every loop \(M\) of a multiple vertex \(v\) meets a simple vertex \(u\).

Proof

We consider the sequence \(U_M=\{v,v_1,\ldots ,v_m,v\}\) of vertices in \(V_M\), arranged in the order as they are met by the edge path \(M\). Let \(w\) denote the element of this sequence with the earliest second appearance (this does not mean the first element which appears twice). Since \(w\) cannot appear twice in direct succession, there exists a vertex \(u\) in between the first two symbols \(w\).

In order to show that \(u\) is a simple vertex, we remark that \(U_M\) is a subsequence of the sequence \(S_L^+\) of upper accompanying vertices. By definition of \(w\), there cannot be a second \(u\) in \(S_L^+\) between the two symbols \(w\) next to \(u\), and by Lemma 15, the sequence \(S_L^+\) cannot contain a second \(u\) outside these two \(w\)’s. \(\square \)

Since loners are vertices in \(U_L^+\), it makes sense to speak of simple and multiple loners.

Lemma 17

Let \(v\) be a multiple loner with \(D_v'\not =D_v\). If \(u\ne v\) is a vertex which is met by a loop of \(v\), then \(u\) is a loner and \(D_u'\cap D_u=\emptyset \).

Proof

Let \(M\) be a loop of \(v\) with \(U_M=\{v,v_1, \ldots ,v_m,v\}\). Let \(i\) be the smallest index, so that \(y_i\) is a contact point of \(v_1\), and let \(j\) be the largest index, so that \(y_j\) is a contact point of \(v_m\). According to the ordering of \(Y\) and \(U_M\) (as subsequences of \(S^+_L\)), \(y_{i-1}\) and \(y_{j+1}\) are contact points of \(D'_v\). Let \(u\in \{v_1,\ldots ,v_m\}\) with \(u\ne v\).

The disk \(D'_u\) is enclosed by the union of the subarc \(\delta ':=\delta [y_{i-1},y_{j+1}]\) of \(D'_v\) and the subarc \(\alpha '\subset \alpha \) which connects the points \(y_{i-1}\) and \(y_{j+1}\) on \(\alpha \) (see Fig. 14). Since \(v\) is a loner with \(D'_v\ne D_v\), it is clear that \(y_{i-1},y_{j+1}\notin D_v\), and hence either \(D_v'\cap D_v=\emptyset \) or \(\partial D_v' \cap \partial D_v\) consists of one or two points. In both cases \(\delta '\) does not intersect \(D_v\). Therefore the union \(\alpha '\cup \delta '\) is contained in \(\overline{\Omega }\), hence \(u\) is a loner. In particular \(D_u'\cap D_u=\emptyset \), which proves the last assertion. \(\square \)

Combining Lemma 8, Lemma 14 (applied recursively), Lemma 16 and Lemma 17 (applied recursively), the essence of this section can be summarized in the following lemma.

Lemma 18

Let \((\mathcal {P},\mathcal {P}')\) be an admissible pair of circle packings with crosscut \(L\).

  1. 1.

    The pair \((\mathcal {P},\mathcal {P}')\) contains a simple loner \(v\in U_L^+\).

  2. 2.

    Every loop of a multiple loner \(v\) meets a simple loner \(u\), and if \(D_v'\not =D_v\) then \(D_u'\not =D_u\).

6 Proof of the main theorem

After all these preparations we are eventually in a position to prove Theorem 2. To begin with, we use the concept of loners and combinatorial surgery to modify the crosscut \(L\). In every step of this procedure the number of vertices in \(V_L^+\) is reduced. At the end we get a special combinatorial structure which is called a slit. Roughly speaking, this is a chain of vertices connecting the alpha-vertex with a boundary vertex. We shall prove that the disks of both packings coincide along a slit.

Then a subdivision procedure generates a sequence of slits, such that any accessible boundary vertex appears among their end points. So we get \(D_v'=D_v\) for all accessible \(v\in \partial V\), and finally a well-known theorem tells us that \(D_v'=D_v\) for all accessible \(v\in V\).

6.1 Combinatoric reduction

Let \(L\) be a combinatoric crosscut of the complex \(K\). In this section we describe how a simple vertex \(v\in U_L^+\) can be “shifted” from \(V_L^+\) to \(V_L^-\) such that we get a new crosscut \(L'\) with \(\big |V_{L'}^+\big |<\big |V_L^+\big |\). Depending on the properties of \(v\) we distinguish three cases.

  1. Case 1

    Let \(v\in U_L^+\) be a simple interior vertex.

  2. Case 2

    Let \(v\in U_L^+\) be a simple boundary vertex, and assume that neither the initial nor the terminal edge of \(L\) are adjacent to \(v\).

  3. Case 3

    Let \(v\in U_L^+\) be a simple boundary vertex, and assume that either the initial or the terminal edge of \(L\) are adjacent to \(v\).

Remark The case where the initial and the terminal edge of \(L\) are adjacent to \(v\) cannot appear. Indeed, otherwise either \(v\) is a multiple vertex (which is not considered) or all edges adjacent to \(v\) must belong to \(L\). The latter implies that \(v\) is the only vertex in \(V_L^+\), which is not allowed.

Reduction of Type 1 In order to modify the crosscut \(L=(e_0,e_1,\ldots ,e_l)\) in Case 1, we consider the flower \(B=B(v)\) of \(v\). Since \(v\) is simple, the set of edges adjacent to \(v\) consists of a subsequence \(S=(e_i,\ldots ,e_j)\) (with \(0\le i\le j\le l\)) of \(L\) and a complementary sequence, which we denote by \(S'=(e_1',\ldots ,e_k')\) (with \(k\ge 1\)). Replacing in \(L\) the sequence \(S\) by \(S'\), we get a new edge sequence

$$\begin{aligned} L' = \left( e_0, \ldots , e_{i-1},e_1',\ldots ,e_k',e_{j+1},\ldots ,e_l\right) . \end{aligned}$$

The reader can easily convince herself (see Fig. 15, left), that the sequence \(L'\) is a crosscut for \(K\) with \(\big |V_{L'}^+\big |<\big |V_L^+\big |\).

Fig. 15
figure 15

Modification of the crosscut \(L\) in Case 1 (left), Case 2 (middle) and Case 3(right)

Reduction of Type 2 In Case 2 the flower of \(v\) is incomplete. Nevertheless, the edges in \(L\) which are adjacent to \(v\) form again a sequence of consecutive edges in this incomplete flower, because \(v\) is simple. However, the local modification of \(L\) in a neighborhood of \(v\) described above does not result in a crosscut \(L'\), since the complementary sequence \(S'=S'_1\cup S'_2\) consists of exactly two connected components \(S'_1=(e_1',\ldots ,e_k')\) and \(S'_2=(e_1'',\ldots ,e_m'')\) (see Fig. 15, middle). Replacing in \(L\) the sequence \(S\) by \(S'_1\) or \(S'_2\), we get a new edge sequence \(L'\) or \(L''\), respectively, with

$$\begin{aligned} L' = \left( e_0, \ldots , e_{i-1},e_1',\ldots ,e_k'\right) , \qquad L'' = \left( e_1'',\ldots ,e_m'',e_{j+1},\ldots ,e_l\right) . \end{aligned}$$

Both sequences \(L'\) and \(L''\) are new crosscuts of \(K\), but only one (\(L'\), say) contains \(v_\alpha \) among its upper vertices, so we choose this one as the new crosscut. Clearly \(\big |V_{L'}^+\big |<\big |V_L^+\big |\).

Reduction of Type 3 If either the initial or the terminal edge of \(L\) are adjacent to \(v\), then the Type 1 reduction applied to the incomplete flower of \(v\) results in an admissible crosscut \(L'\), which has one vertex (namely \(v\)) less in \(V_{L'}^+\) than in \(V_{L}^+\) (see Fig. 15, right).

Remark No matter which type of reduction we used, the sets \(U^-_L\) and \(U^-_{L'}\) of lower neighbors before and after the reduction, respectively, always fulfill \(U^-_{L'}{\setminus } U^-_L=\{v\}\).

In order to not lose the normalization, we will only reduce vertices different from \(v_\alpha \). This leads to a situation where none of the above reductions can be applied, namely when \(v_\alpha \) is the only simple vertex in \(U_L^+\). This special case will be explored in Sect. 6.2.

6.2 Slits

The next definition and the following lemma describe the situation when all but exactly one vertex of \(U_{L}^{+}\) are multiple.

Definition 8

A combinatoric slit of the complex \(K(V,E,F)\) is a sequence \(S=(v_1,v_2,\ldots ,v_{s})\) of vertices in \(V\) which satisfies the following conditions (1)–(4):

  1. (1)

    The vertices of \(S\) are pairwise different, \(v_j\not =v_k\) if \(1\le j<k\le s\).

  2. (2)

    For \(j=1,\ldots ,s-1\), the edges \(e_j:=e(v_{j},v_{j+1})\) belong to \(E\).

  3. (3)

    For \(j=1,\ldots ,s\), the vertices \(v_{j-1}\) and \(v_{j+1}\) are the only neighbors of \(v_j\) in \(K\) which belong to \(S\) (where \(v_{0}:=\emptyset \) and \(v_{s+1}:=\emptyset \)).

  4. (4)

    The vertex \(v_1\) is a boundary vertex, and \(v_j\) are interior vertices for \(j=2,\ldots ,s\).

The vertices \(v_1\) and \(v_s\) are referred to as the initial vertex and the terminal vertex of \(S\), respectively. The sequence \(E_S:=(e_1,\ldots ,e_{s-1})\) [see (2)] is said to be the edge sequence of \(S\). Note that all \(e_j\) are interior edges.

Lemma 19

Assume that the interior vertex \(v\) is the only simple vertex in \(U_L^+\). Then the sequence of upper accompanying vertices \(S_L^+\) has the symmetric form \((v_1,\ldots ,v_{s-1},v,v_{s-1},\ldots ,v_1)\) and \(S=(v_1,\ldots ,v_{s-1},v)\) is a slit.

Proof

By definition of a multiple vertex, any vertex in \(U_L^+\) except \(v\) must appear at least twice in the sequence \(S_L^+\). If there are vertices which show up twice at a position left of \(v\), we choose one, say \(u\), whose appearances have minimal distance in the sequence \(S_L^+ = (\ldots , u,\ldots ,u,\ldots ,v,\ldots )\). Since neighboring vertices of \(S_L^+\) must be different, there exists \(w\not =u\) such that \(S_L^+ = (\ldots , u,\ldots ,w,\ldots ,u,\ldots ,v,\ldots )\). Because \(v\) is assumed to be simple and \(w\) is a multiple vertex, we have \(w\not =v\) and \(w\) must appear again at another place in \(S_L^+\). By Lemma 15 this can only happen in between the two occurrences of \(u\), which is in conflict with the minimal distance property of \(u\).

Similarly, the assumption that there exists a vertex which appears in \(S_L^+\) twice at a position right of \(v\) leads to a contradiction. Hence, with the only exception of \(v\), any vertex of \(U_L\) appears in \(S_L^+\) exactly once on either side of \(v\). Applying Lemma 15 again, we see that the ordering of the vertices left of \(v\) must be reverse to the ordering on the right of \(v\), so that \(S_L^+\) has the symmetric form claimed in the lemma.

Moreover we have shown that \(v_1,\ldots ,v_{s-1},v\) are pairwise different, which is condition (1) of Definition 8. The second condition (2) is trivial.

In order to verify condition (4), it remains to show that \(v_j\) is an interior vertex for \(j=2,\ldots ,{s-1}\), because \(v_1\) is obviously a boundary vertex, while \(v_s:=v\) is an interior vertex, by assumption. Assume \(v_j\) is a boundary vertex. The flower of \(v_j\) is incomplete and it is clear that \(v_{j-1}\) and \(v_{j+1}\) are neighbors of \(v_j\). On the one hand, since \((v_{j-1},v_j,v_{j+1})\) is a subsequence of \(S^+_L\), the crosscut \(L\) must look locally like shown in Fig. 16 left. On the other hand, the subsequence \((v_{j+1},v_j,v_{j-1})\) of \(S^+_L\) forces \(L\) to look locally like depicted in the middle of Fig. 16, a contradiction. Hence \(v_j\) must be an interior vertex and its flower must look qualitatively like shown in Fig. 16 right.

Fig. 16
figure 16

Illustrations for the proof of Lemma 19

To verify condition (3) let \(j\in \{2,\ldots ,s-1\}\) be fixed. Looking at the behavior of the crosscut \(L\) in the flower of \(v_j\), it becomes clear that any edge \(e(v_{j-1},v_{j+1})\) (with the convention \(v_s:=v\)) belonging to \(E\) must be contained in \(L\) twice, a contradiction. Furthermore, all other neighbors of \(v_j\) belong to \(V_L^-\) and hence not to \(V_L^+ \supset S^+_L\). A similar result can be derived by looking at the local behavior of \(L\) in the flower of \(v\) and the incomplete flower of \(v_1\), now using the subsequences \((v_{s-1},v_s,v_{s-1})\) and \((v_1,v_2,\ldots ,v_2,v_1)\) of \(S^+_L\), respectively. \(\square \)

The following lemma explains why we are interested in slits.

Lemma 20

Let \((\mathcal {P},\mathcal {P}')\) be an admissible pair of circle packings for the complex \(K\) with crosscut \(L\) and alpha-vertex \(v_\alpha \). Then there exists a slit \(S=(v_1,\ldots ,v_{s},v_\alpha )\subset V_L^+\) with terminal vertex \(v_\alpha \) such that \(D_v'=D_v\) for all \(v\in S\).

Proof

To begin with, we invoke Lemma 18, which tells us that the pair \((\mathcal {P},\mathcal {P}')\) has a simple loner \(v_\lambda \). The idea is to use the reduction procedures of the last section to shift \(v_\lambda \) from \(V_L^+\) to \(V_L^-\) which results in a new crosscut \(L'\).

As we remarked earlier, the one and only lower neighbor of \(L'\) which has not already been a lower neighbor of \(L\) is the simple loner \(v_\lambda \). Therefore Lemma 7 guarantees that \(L'\) is admissible for \((\mathcal {P},\mathcal {P}')\). In order to find the appropriate type of reduction we distinguish the following cases:

  1. Case 1

    There exists a simple interior loner \(v_\lambda \) different from the alpha-vertex \(v_\alpha \).

  2. Case 2

    There exists a simple boundary loner \(v_\lambda \).

  3. Case 3

    The only simple loner \(v_\lambda \) is the alpha-vertex \(v_\alpha \).

In Case 1 we apply the reduction of Type 1, while in Case 2 either the reduction of Type 2 or Type 3 can be applied, respectively, depending on whether \(v_\lambda \) is adjacent to the initial or the terminal edge of \(L\), or not. In any case we get a new combinatoric crosscut \(L'\) of \(K\). Applying the reduction in Cases 1 and 2 recursively as long as possible, the number of vertices in \(V_L^+\) decays in every step at least by one, so that we eventually arrive at Case 3.

Since the disks \(D_\alpha '\) and \(D_\alpha \) have the same centers, we either have one of the strict inclusions \(D_\alpha '\subset D_\alpha \), \(D_\alpha \subset D_\alpha '\) or \(D_\alpha '=D_\alpha \). The first case cannot occur, since otherwise all neighboring disks of \(D_\alpha '\) would intersect \(D_\alpha \), a contradiction for those disks associated with a vertex in \(U^-_L\). The second case clearly implies that \(v_\alpha \) is an intruder. So the alpha-vertex \(v_\alpha \) is a loner if and only if \(D_\alpha '=D_\alpha \). This implies, by Lemma 14, that there exists another loner \(v_\mu \). Since \(v_\alpha \) is the only simple loner, \(v_\mu \) must be a multiple loner. If \(D_\mu '\not =D_\mu \), then according to Lemma 18 (1), the vertex set \(V_M\) of any loop \(M\) of \(v_\mu \) contains a simple loner, i.e., \(M\) meets \(v_\alpha \). Because \(D_\alpha '=D_\alpha \), assertion (2) of this lemma tells us that \(D_\mu '=D_\mu \).

Applying Lemmas 14 and 18 repeatedly in this manner, we see that all vertices in \(U_L^+{\setminus } \{v_\alpha \}\) must be multiple loners and hence that \(D_v'=D_v\) for all \(v\in U_L^+\). Furthermore \(v_\alpha \) is the only simple vertex in \(U^+_L\), so, by Lemma 19, we just constructed a slit \(S\subset V_L^+\) with terminal vertex \(v_\alpha \). \(\square \)

In the next step we are going to construct crosscuts from slits. To begin with, we introduce some more notations.

Fig. 17
figure 17

Left and right neighboring edges of vertices \(v=v_1,v_j,v_s\) in a slit \(S\)

Let \(S=(v_1,\ldots ,v_s)\) be a slit. For any vertex \(v\) in \(S\) we define the subsets \(E_S^-(v)\) and \(E_S^+(v)\) of \(E(v)\) as follows. For \(v=v_1\), the (boundary) vertex \(v_1\) has two adjacent boundary edges \(e_1^-\) and \(e_1^+\) in \(E(v_1)\), such that \(e_1^-\) is the predecessor of \(e_1^+\) in the chain of boundary edges. The meaning of the inequalities in the following definitions is explained in the second paragraph on complexes in Section 2. For the initial vertex \(v_1\) we set (see Fig. 17, left)

$$\begin{aligned} E_S^-(v_1)&:=\left\{ e\in E(v_1): e(v_1,v_2)< e\le e_1^-\right\} ,\\ E_S^+(v_1)&:=\left\{ e\in E(v_1): e_1^+\le e < e(v_1,v_2)\right\} . \end{aligned}$$

If \(v=v_j\), with \(j=2,\ldots s-1\), we define (Fig. 17, middle)

$$\begin{aligned} E_S^-(v_j)&:=\left\{ e\in E(v_j): e(v_{j},v_{j+1}) < e < e(v_{j-1},v_j)\right\} ,\\ E_S^+(v_j)&:=\left\{ e\in E(v_j): e(v_{j-1},v_j) < e < e(v_j,v_{j+1})\right\} , \end{aligned}$$

and for the terminal vertex \(v_s\) of \(S\) we let (see Fig. 17, right)

$$\begin{aligned} E_S^-(v_s)=E_S^+(v_s):=\left\{ e\in E(v_s): e(v_{s-1},v_{s}) < e < e(v_{s-1},v_s)\right\} . \end{aligned}$$

The edges in

$$\begin{aligned} E_S^-:=\mathbin {\cup }_{j=1}^{s-1} E_S^-(v_j)\quad \text {and}\quad E_S^+:=\mathbin {\cup }_{j=1}^{s-1} E_S^+(v_j) \end{aligned}$$

are called the left and the right neighbors of \(S\), respectively. Note that condition (3) in Definition 8 guarantees that every edge \(e\) which is a neighbor of a slit \(S\) has exactly one adjacent vertex in \(S\).

Lemma 21

If \(S=(v_1,\ldots ,v_s,v)\) is a slit in \(K\), then there exists a combinatoric crosscut \(L\) such that \(v\in S_L^+\), and \(S_L^-=(v_1,\ldots ,v_{s-1},v_s,v_{s-1},\ldots ,v_1)\) is the sequence of lower accompanying vertices of \(L\).

Proof

Walking along the slit \(S\) from \(v_1\) to \(v_{s}\) and back to \(v_1\), we build the crosscut \(L\) from the concatenation of the edge sequences

$$\begin{aligned} E_{S}^-(v_{1}), \ldots , E_{S}^-(v_{s}), e(v_s,v), E_{S}^+(v_{s}), \ldots , E_{S}^+(v_{1}). \end{aligned}$$

It is easy to see that all edges in \(L\) are pairwise different, so that \(L\) satisfies condition (1) of Definition 3. Condition (2) can easily be verified and (4) is obvious. In order to prove (3) we assume that three edges of \(L\) would form a face of \(K\). Since these edges are neighbors of \(S\), exactly one vertex of every edge must belong to \(S\), which is impossible.

The construction also guarantees that the sequence \(S_L^-\) of lower accompanying edges of \(L\) has the desired form and that \(v\) belongs to \(S_L^+\) (see, for example, Fig. 18, left). \(\square \)

Fig. 18
figure 18

Constructing crosscuts from one slit (left) and two slits (middle, right)

A crosscut \(L\) can also be constructed from joining two slits \(S_1\) and \(S_2\) with a common terminal vertex \(v\). This procedure is somewhat more complicated, in particular when the “right side” of \(S_1\) is close to the “left side” of \(S_2\). In those cases we cannot join the cuts at their common terminal vertex \(v\), since then the resulting edge sequence \(L\) would contain some edges more than once. Instead we modify the procedure by linking \(S_1\) and \(S_2\) at some appropriately chosen vertex \(u\) in \(S_2\) or \(S_1\) which has a neighbor in \(S_1\) or \(S_2\), respectively. Figure 18 (middle, right) illustrates the result, showing an associated circle packing and the related maximal crosscuts.

Lemma 22

Let \(S_1=(v_{1},\ldots ,v_{t},v)\) and \(S_2=(w_{1},\ldots ,w_{s},v)\) be slits in \(K\) with \(S_1\cap S_2=\{v\}\). Assume further that \(E^+_{S_1}(v_1)\cap E^-_{S_2}(w_1)=\emptyset \). Then there exists a combinatoric crosscut \(L\) and a vertex \(u\in (S_1\cup S_2)\cap U_L^+\) such that

$$\begin{aligned} S_L^-=\big (w_{1},w_{2},\ldots ,w_{\sigma },u_1,\ldots ,u_k, v_\tau ,v_{\tau -1},\ldots ,v_1\big ), \qquad 1\le \tau \le t,\ 1\le \sigma \le s, \end{aligned}$$
(1)

where \(\big (w_{\sigma },u_1,\ldots ,u_k,v_\tau \big )\) is a (positively oriented) chain of neighbors of \(u\).

Note that the condition \(E^+_{S_1}(v_1)\cap E^-_{S_2}(w_1)=\emptyset \) does not exclude that \(v_1\) and \(w_1\) share an edge. Loosely speaking, it means that there is no such edge connecting the “plus side” of \(S_1\) and the “minus side” of \(S_2\);

Proof

We set \(v_{t+1}:=v\) and \(w_{s+1}:=v\). Let \(i\) be the smallest number in \(\{1,...,t+1\}\) for which \(E_{S_1}^+(v_i)\) contains an edge \(e(v_i,w)\) with \(w\in S_2\). Then let \(j\) be the smallest number in \(\{1,...,s+1\}\) for which \(E_{S_2}^-(w_j)\) contains an edge \(e(w_j,v_i)\). If \(i\ne 1\) and \(j\ne s+1\) we set \(\tau :=i-1\), \(\sigma :=j\) and \(u:=v_i\). If \(i\ne 1\) but \(j=s+1\), then \(i=t\) must hold (otherwise \(v\) would have more then one neighbor in \(S_1\)), and we set \(\tau :=t\), \(\sigma :=s\) and \(u:=v\). If \(i=1\) we set \(\tau :=1\), \(\sigma :=j-1\) and \(u:=w_j\). In the last case we have \(j>1\), since otherwise \(i=j=1\) would contradict the assumption \(E^+_{S_1}(v_1)\cap E^-_{S_2}(w_1)=\emptyset \).

In every case \(1\le \tau \le t\) and \(1\le \sigma \le s\) hold, and \(u\) is well defined. We now build \(L\) as the concatenation of the edge sequences

$$\begin{aligned} E_{S_2}^-(w_1), \ldots , E_{S_2}^-(w_{\sigma }), E^*(u), E_{S_1}^+(v_{\tau }), \ldots , E_{S_1}^+(v_1), \end{aligned}$$

where \(E^*(u)=\big (e(u,w_{\sigma }),e(u,u_1),\ldots ,e(u,u_k),e(u,v_{\tau })\big )\) is the negatively oriented chain of edges in the set \(\{e'\in E(v): e(u,w_{\sigma })\le e'\le e(u,v_\tau )\}\).

Because \(S_1,S_2\) are slits, all edges in the “\(E_{S_1}^+\)-part” and in the “\(E_{S_2}^-\)-part” of \(L\) are pairwise different. Furthermore, it cannot happen that such an edge is contained in both parts (according to the definition of \(u\)), or that it belongs to \(E^*(u)\) (by definition of \(E^*(u)\)). Hence, \(L\) satisfies condition (1) of the crosscut definition.

Condition (2) can easily be verified and (4) is trivial. In order to prove (3) we assume that three edges of \(L\) form a face of \(K\). By definition of \(u\), the sequence \((w_1,w_2, \ldots ,w_{\sigma },u,v_{\tau }, \ldots ,v_2,v_1)\) divides \(K\) into two parts \(K_1,K_2\). All edges of the “\(E_{S_1}^+\)-part” and of the “\(E_{S_2}^-\)-part” have exactly one vertex lying in \(S_1^0\cup S_2^0\) and one in \(K_1\), so three of them can never form a face of \(K\). All edges of \(E^*(u){\setminus }\{e(u,v_{\tau }),e(u,w_{\sigma })\}\) have exactly one vertex lying in \(S_1^0\cup S_2^0\) and one in \(K_2\), so again three of them can never form a face of \(K\). The only remaining edges are \(e(u,v_{\tau }),e(u,w_{\sigma })\), but two edges cannot form a face, and a combination of edges from more than one of the three distinguished edge types can clearly never form a face. Hence, \(L\) is a crosscut with \(u\in (S_1\cup S_2)\cap U_L^+\), and \(S_L^-\) has the form (1). \(\square \)

The operation described in the proof is well defined by the slits \(S_1\) and \(S_2\), and will be referred to as reflected concatenation \(S_1\circleddash S_2\) of \(S_1\) with \(S_2\). It delivers a crosscut \(L\), a vertex \(u\), and the reduced slits \(S_1^0,S_2^0\). Note that the reflected concatenation is not commutative.

6.3 Subdivision by disk chains

Let \(v_\beta \) be an arbitrary accessible boundary vertex. In this section we describe an approach which allows us to apply Lemma 20 recursively, until we find a slit \(S\) with initial vertex \(v_\beta \) such that \(D_v'=D_v\) for all \(v\in S\), so especially \(D'_{v_\beta }=D_{v_\beta }\). During this procedure we construct a sequence of crosscuts \(L_j\) such that \(V_{L_j}^+\) contains \(v_\beta \) and the number of elements in \(V_{L_j}^+\) is strictly decreasing for increasing \(j\). This procedure will be crucial for proving the following lemma, and finally Theorem 2.

Lemma 23

Let \((\mathcal {P},\mathcal {P}')\) be an admissible pair with complex \(K\), interior alpha vertex \(v_\alpha \) and crosscut \(L\). Then \(D'_v=D_v\) for all accessible boundary vertices \(v\in \partial V^*\).

Proof

To begin with, let \(S_0=(v_1,\ldots ,v_{s},v_\alpha )\) be a slit according to Lemma 20. Let \(v_\beta \) be an accessible boundary vertex. If \(v_1=v_\beta \) then \(D_\beta '=D_\beta \) and we are done. So let us assume that \(v_\beta \notin S_0\).

By Lemma 21 there exists a crosscut \(L_1\) such that \(S_{L_1}^-=(v_1,\ldots ,v_{s-1},v_s,v_{s-1},\ldots ,v_1)\) and \(v_\alpha \in S_{L_1}^+\). Applying Lemma 20 again, but now with respect to the crosscut \(L_1\), we get another slit \(S_1=(w_{1},\ldots ,w_{t},v_\alpha )\subset V_{L_1}^+\), such that \(D_v'=D_v\) for all \(v\in S_1\). If \(w_{1}=v_\beta \) then \(D_\beta '=D_\beta \) and we are done. So suppose that \(v_\beta \notin S_1\).

The three boundary vertices \(v_1\), \(w_1\) and \(v_\beta \) are pairwise different, and we assume, without loss of generality, that they are oriented such that \(w_1<v_\beta <v_1\). This ensures that \(E^+_{S_1}(v_1)\cap E^-_{S_0}(w_1)=\emptyset \), because otherwise \(v_\beta \) could be either accessible or a boundary vertex, but not both. Since, except \(v_\alpha \), all vertices of \(S_0\) belong to \(V_{L_1}^-\), we have \(S_0\cap S_1 = \{v_\alpha \}\). Consequently, by Lemma 22, the reflected concatenation \(S_0\circleddash S_1\) of \(S_0\) with \(S_1\) is well defined. It delivers a crosscut \(L_2\), a vertex \(v_{\alpha _2}\), and reduced slits \(S_2^- \subset S_0\), \(S_2^+ \subset S_1\) with common terminal vertex \(v_{\alpha _2}\). Since \(E^+_{S_1}(v_1)\cap E^-_{S_2}(w_1)=\emptyset \) (see above), by Lemma 22 the vertex \(v_{\alpha _2}\) belongs to \(S_1\) or \(S_2\) and the set \(U_{L_2}^-\) of lower neighbors of \(L_2\) consists solely of elements of \(S_0 \cup S_1\) and of (lower) neighbors of \(v_{\alpha _2}\). Since \(D_v'=D_v\) for all \(v\in S_0 \cup S_1\), this implies that \(L_2\) is an admissible crosscut for \((\mathcal {P},\mathcal {P}')\). Moreover, the order of \(S_0\) and \(S_1\) in the reflected concatenation has been chosen such that \(v_\beta \) belongs to \(V_{L_2}^+\).

The general step of the procedure is as follows. Assume that we already have an admissible crosscut \(L_j\), the alpha vertex \(v_{\alpha _j}\), and the reduced slits \(S_j^-\) and \(S_j^+\), such that \(v_\beta \in V_{L_j}^+\) (see Fig. 19, left). Denoting by \(v_j^-\) and \(v_j^+\) the initial vertices of \(S_j^-\) and \(S_j^+\), respectively, we may assume that \(v_j^-<v_\beta <v_j^+\), which will again be essential to ensure the special condition of Lemma 22.

Applying Lemma 20, we get a new slit \(S_j\subset V_{L_j}^+\), such that \(S_j^-\), \(S_j\) and \(S_j^+\) are pairwise disjoint, except at their common terminal vertex \(v_{\alpha _j}\), and \(D_v'=D_v\) for all \(v\in S_j\) (see Fig. 19, middle).

Fig. 19
figure 19

Construction of the crosscut \(L_{j+1}\) from \(L_j\)

If \(v_\beta \in S_j\) we are done. Otherwise we either have \(v_j^-<v_\beta <v_j\) or \(v_j<v_\beta <v_j^+\). In the first case we build the reflected concatenation \(S_j^-\circleddash S_j\), in the second case we form \(S_j\circleddash S_j^+\). The result is a new crosscut \(L_{j+1}\), a corresponding alpha-vertex \(v_{\alpha _{j+1}}\), and reduced slits \(S_{j+1}^-\), \(S_{j+1}^+\) (see Fig. 19, right).

It follows directly from the construction of the reflected concatenation that \(v_{\alpha _{j+1}},v_\beta \in V_{L_{j+1}}^+\). Moreover, \(v_{\alpha _{j+1}}\in S_j^-\), and hence \(D_{\alpha _{j+1}}'=D_{\alpha _{j+1}}\). To see that \(L_{j+1}\) is admissible for the pair \((\mathcal {P},\mathcal {P}')\) it remains to prove that \(D_v'\subset G_{L_{j+1}}^-\) for all \(v\in U_{L_{j+1}}^-\).

By Lemma 22 the set \(U_{L_{j+1}}^-\) of lower neighbors of \(L_{j+1}\) consists solely of elements of \(S_j^- \cup S_j^+\) and of (lower) neighbors of \(v_{\alpha _{j+1}}\). Since \(D_v'=D_v\) for all \(v\in S_j^- \cup S_j^+ \cup \{v_{\alpha _{j+1}}\}\), and \(D_v\subset G_{L_{j+1}}^-\) for all \(v\in U_{L_{j+1}}^-\), the assertion follows.

The number of elements in \(V_{L_j}^+\) is strictly decreasing in every step, and hence the procedure must come to end. This can only happen if \(v_\beta \in S_{j^*}\) for some \(j^*\in \mathbb {N}\). Because \(D_v'=D_v\) for all \(v\in S_j\) with \(j\le j^*\), we have shown \(D'_{v_\beta }=D_{v_\beta }\). \(\square \)

Now the proof of Theroem 2 is near to its end. By Lemma 1 the kernel \(K^*\) is a strongly connected complex with vertex set \(V^*\). Since we have shown that \(D_v'=D_v\) for all boundary vertices \(v\in \partial V^*\) of \(K^*\), and every boundary vertex of \(K^*\) is also a boundary vertex of \(K\) (that is \(\partial V^*=V^*\cap \partial V\)), Theorem 11.6 in  Stephenson (2005) (on the uniqueness of a locally univalent packing with presribed combinatorics and given radii of boundary disks) tells us that \(D_v'=D_v\) for all \(v\in V^*\), which is the assertion of Theorem 2.

7 Concluding remarks

All proofs in this paper work with (simple) geometric or combinatoric arguments, alone in the very last step we had recourse to a theorem established in the literature. For purists we mention that even this could have been avoided, at the expense of adding a few pages to this rather longish text.

Theorem 2 can be interpreted as a uniqueness result for (the range packing of) discrete conformal mappings. Here is a simple version:

Theorem 3

Suppose that two univalent packings \(\mathcal {P}\) and \(\mathcal {P}'\) for \(K\) fill \(G\). If \(D'_\alpha \) and \(D_\alpha \) have the same center, and if \(D'_\beta \subset D_\beta \) for some boundary vertex \(v_\beta \), then \(D_v'=D_v\) for all vertices \(v\in V^*\).

The proof follows immediately from Theorem 2 applied to the maximal crosscut which separates the disk \(D_\beta \) from the rest of the packing \(\mathcal {P}\) (see the leftmost image of Fig. 20). The condition \(D_\beta '\subset D_\beta \) can even be relaxed, it suffices to require that \(D'_\beta \) lies in the lower domain \(G_-\) with respect to this crosscut (see the second image of Fig. 20). Note that both figures show the packing \(\mathcal {P}\) and a single disk \(D'_\beta \) of \(\mathcal {P}'\) in \(G_-\).

Fig. 20
figure 20

Applications of Theorem 2 to discrete conformal mapping

We point out that the condition \(D_\beta '\subset G_-\) is always satisfied (possibly after exchanging the roles of \(\mathcal {P}\) and \(\mathcal {P}'\)), if the packings are normalized so that \(D'_\beta \) and \(D_\beta \) touch the boundary \(\partial G\) in a generalized sense at the same regular point (or, more generally, at the same regular prime end). Without explaining these concepts here (see Wegert and Krieg 2014), we mention that a point which lies on a smooth subarc of \(\partial G\) is always regular, while a point at a re-entrant corner fails to be regular. The two pictures on the right of Fig. 20 illustrate that uniqueness of domain-filling circle packings may be violated in that case. Both displayed packings \(\mathcal {P}\) and \(\mathcal {P}'\) fill a Jordan domain \(G\), \(D_\alpha \) and \(D_\alpha '\) have the same center, and \(D_\beta \) and \(D'_\beta \) touch \(\partial G\) at the same point. While this type of normalization implies uniqueness of classical conformal mappings, the corresponding circle packings \(\mathcal {P}\) and \(\mathcal {P}'\) are completely different.

We further mention that for domain-filling circle packings \(\mathcal {P}\) and \(\mathcal {P}'\) the assertions of Theorems 2 and 3 can be strengthened to \(D'_v=D_v\) for all \(v\in V\), using the results of our forthcoming paper (Krieg and Wegert 2015).

In the general setting of Theorem 2, a complete description of which disks are uniquely determined by a crosscut seems not to be known. Figure 21 shows some examples. The accessible disks are depicted in darker colors, the alpha-disk is the darkest one. By Theorem 2 these disks are uniquely determined (rigid) by the crosscut, but the rigid part also comprises the non-accessible disks shown in brighter color.

Fig. 21
figure 21

Rigid configurations of disks in a packing with crosscut

The example on the right is of special interest: a short crosscut separates only one non-accessible disk \(D_\beta \) from the alpha-disk. Here the theorem yields rigidity for the dark (blue) disks, while it says nothing about the disks depicted in lighter colors. This is somewhat counterintuitive, since the bright disks separate the dark disks from the crosscut, so that the latter seem to have no relation to the cut at all. However, a little thought shows that in fact all colored disks in the upper domain are rigid. It is a challenging problem to precisely describe the set of all rigid disks in a circle packing with crosscut.

Isn’t it wonderful that simple circles can form such fascinating structures?

Glossary

\(\circleddash \), \(S_1\circleddash S_2\) :

reflected concatenation of slit \(S_1\) with slit \(S_2\)

\(\langle u,v,w \rangle \) :

oriented face of \(K\) with vertices \(u\),\(v\) and \(w\)

\(\langle u,v \rangle \) :

oriented edge of \(K\) from vertex \(u\) to vertex \(v\)

\(\alpha _i, \alpha \) :

special Jordan arcs connecting \(y_i^-\) and \(y_i^+\), and their concatenation

\(B(v)\) :

the flower of the vertex \(v\), a subcomplex of \(K\)

\(c(u,v)\) :

contact point of the disks \(D_u\) and \(D_v\), \(c(u,v)=\overline{D}_u \cap \overline{D}_v\)

\(c_k^-, c_k^+\) :

contact points of boundary disk \(D_k\) with \(D_{k-1}\) and \(D_{k+1}\), respectively

\(D\) :

union of all disks in \(\mathcal {P}\)

\(D^*\) :

carrier of \(\mathcal {P}\)

\(D_k, D_k'\) :

boundary disks in \(\mathcal {P}\) and \(\mathcal {P}'\), respectively

\(D_v, D_v'\) :

disks in \(\mathcal {P}\) and \(\mathcal {P}'\), respectively

\(\partial \) :

boundary operator, applied to various objects

\(\delta (p,q)\) :

positively oriented open circular arc from \(p\) to \(q\) on \(\partial D\)

\(\delta [p,q]\) :

positively oriented closed circular arc from \(p\) to \(q\) on \(\partial D\)

\(\delta (c^-_k,c^+_k)\) :

exterior boundary arc of \(D_k\)

\(\delta (c^+_k,c^-_k)\) :

interior boundary arc of \(D_k\)

\(\delta _k\) :

smallest subarc of \(\delta [c^-_k,c^+_k]\) which contains \(G_k\)

\(E_S\) :

the edge sequence of the slit \(S\)

\(E\) :

the set of edges of the complex \(K\)

\(\partial E\) :

boundary edges of the complex \(K\)

\(E(v)\) :

the (cyclically ordered) sequence of edges adjacent to \(v\in V\)

\(E_L^\pm (v)\) :

sequences of upper and lower accompanying edges of the crosscut \(L\)

\(E_S^\pm (v)\) :

sequences of edges adjacent to a vertex \(v\) in a slit \(S\)

\(E_S^\pm \) :

sequences of left and right neighbor edges of slit \(S\), respectively

\(e(u,v)\) :

non-oriented edge between vertices \(u\) and \(v\)

\(e_j\) :

edges in a crosscut, \(L=(e_0,e_1,\ldots ,e_l)\)

\(e_j^-\), \(e_j^+\) :

lower and upper accompanying edges of the crosscut \(L\), respectively

\(\eta _k, \eta \) :

segments connecting the centers of \(D_k\) and \(D_{k+1}\) and their concatenation

\(F\) :

set of faces of the complex \(K\)

\(f(u,v,w)\) :

non-oriented face with vertices \(u,v\) and \(w\)

\(G\) :

a bounded simply connected domain in \(\mathbb {C}\)

\(G_L^-, G_L^+\) :

lower and upper domains of \(G\) with maximal crosscut \(J_L^+\), \(G_L^-=\Omega \)

\(G_k\) :

set of contact points of \(D_k\) with \(\partial G\), \(G_k:=\overline{D}_k \cap \partial G\)

\(g_k^-,g_k^+\) :

first and the last contact point of \(D_k\) with \(\partial G\)

\(I_k\) :

boundary interstice between \(D_k\) and \(D_{k+1}\)

\(I(u,v,w)\) :

interstice between the disks \(D_u,D_v\) and \(D_w\)

\(J_L^0\) :

polygonal (geometric) crosscut in \(G\) for (combinatoric) crosscut \(L\) in \(K\)

\(J_L^+\) :

maximal ‘crosscut’, the upper boundary of the lower domain \(G_L^-\), \(J_L^+=\omega \)

\(K\) :

simplicial 2-complex, combinatorial disk, finite triangulation, \(K(V,E,F)\)

\(K^*\) :

kernel of \(K\), largest sub-complex of \(K\) with vertex set \(V^*\)

\(L\) :

combinatorial crosscut, sequence of edges in \(K\)

\(l(i)\) :

smallest label \(k\) of prime end set \(\omega _k^*\) associated with \(\nu _i\)

\(M\), \(M(\mu )\) :

loop of a multiple loner \(v_\mu \), a sequence of edges

\(\nu _i, \pi _i\) :

negatively and positively oriented arcs on \(\partial D\) from \(y_i^-,y_i^+\) to \(\omega \), respectively

\(\nu _i^+, \pi _i^+\) :

terminal points of the arcs \(\nu _i, \pi _i\), respectively

\(\nu _i^*, \pi _i^*\) :

prime ends of \(\Omega \) associated with \(\nu _i, \pi _i\), respectively

\(\Omega \) :

lower subdomain of \(G\) with respect to a maximal crosscut, \(\Omega =G_L^-\)

\(\omega \) :

upper boundary of lower domain \(\Omega \), concatenation of the \(\omega _i\), maximal crosscut

\(\omega ^*\) :

prime ends of \(\Omega \) associated with \(\omega \)

\(\omega _i\) :

circular subarcs of \(\omega \) in between its turning points

\(\omega _i^*\) :

classes of prime ends associated with the arcs \(\omega _i\)

\(\mathcal {P}\) :

a univalent circle packing for \(K\) filling \(G\)

\(\mathcal {P}'\) :

a univalent circle packing for \(K\) in \(G\)

\(r(i)\) :

largest label \(k\) of prime end set \(\omega _k^*\) associated with \(\pi _i\)

\(S\) :

combinatoric slit, a sequence of vertices

\(S_L^-\), \(S_L^+\) :

sequences of lower and upper accompanying vertices of \(L\), respectively

\(t_i\) :

turning points of the upper boundary \(\omega \), cusps of \(\Omega \)

\(U_L^-\), \(U_L^+\) :

sets of lower and upper neighbors of \(L\), respectively, \(U_L^-\subset V_L^-, U_L^+\subset V_L^+\)

\(U_M\) :

sequence of the vertices in \(V_M\) for a loop \(M\)

\(V\) :

vertex set of the complex \(K\)

\(V^*\) :

the set of all accessible vertices of \(K\)

\(\partial V\) :

boundary vertices of the complex \(K\)

\(V_L^-,V_L^+\) :

lower and upper vertices of \(K\) with crosscut \(L\), respectively, subsets of \(V\)

\(V_M\) :

set of all vertices met by a loop \(M\)

\(v_\alpha \) :

alpha vertex of \(K\), a distinguished interior vertex

\(v(i)\) :

vertex of the disk which contains the circular arc \(\omega _i\), \(v(i)\in U_L^+\)

\(x_k, X\) :

contact points of upper with lower disks in \(\mathcal {P}\), the set of all \(x_k\)

\(X_i\) :

sets of contact points \(x_k\) on \(\omega _i\), \(X_i \subset X\)

\(y_-,y_+\) :

initial point and terminal point of \(\alpha \), respectively

\(y_k, Y\) :

contact points of upper with lower disks in \(\mathcal {P}'\), the set of all \(y_k\)

\(y_i^-,y_i^+\) :

minimal and maximal element of \(Y_i\), respectively

\(Y_i\) :

sets of contact points \(y_k\) with \(x_k\in \omega _i\), \(Y_i \subset Y\)

\(z_-,z_+\) :

terminal points of \(\nu _1\) and \(\pi _n\), respectively

\(z_k\) :

shifted contact points when \(y_k\) is critical