Skip to main content
Wiley Open Access Collection logoLink to Wiley Open Access Collection
. 2022 Feb 21;105(2):1262–1313. doi: 10.1112/jlms.12539

Self‐adjoint and Markovian extensions of infinite quantum graphs

Aleksey Kostenko 1,2,3,, Delio Mugnolo 4, Noema Nicolussi 3
PMCID: PMC9303478  PMID: 35912286

Abstract

We investigate the relationship between one of the classical notions of boundaries for infinite graphs, graph ends, and self‐adjoint extensions of the minimal Kirchhoff Laplacian on a metric graph. We introduce the notion of finite volume for ends of a metric graph and show that finite volume graph ends is the proper notion of a boundary for Markovian extensions of the Kirchhoff Laplacian. In contrast to manifolds and weighted graphs, this provides a transparent geometric characterization of the uniqueness of Markovian extensions, as well as of the self‐adjointness of the Gaffney Laplacian — the underlying metric graph does not have finite volume ends. If, however, finitely many finite volume ends occur (as is the case of edge graphs of normal, locally finite tessellations or Cayley graphs of amenable finitely generated groups), we provide a complete description of Markovian extensions upon introducing a suitable notion of traces of functions and normal derivatives on the set of graph ends.

1. INTRODUCTION

This paper is concerned with developing extension theory for infinite quantum graphs. Quantum graphs are Schrödinger operators on metric graphs, that is, combinatorial graphs where edges are considered as intervals with certain lengths. Motivated by a vast amount of applications in chemistry and physics, they have become a popular subject in the last decades (we refer to [8, 9, 26, 67] for an overview and further references). From the perspective of Dirichlet forms, quantum graphs play an important role as an intermediate setting between Laplacians on Riemannian manifolds and difference Laplacians on weighted graphs. On the one hand, being locally one‐dimensional, quantum graphs allow to simplify considerations of complicated geometries. On the other hand, there is a close relationship between random walks on graphs and Brownian motion on metric graphs, however, in contrast to the discrete case, the corresponding quadratic form in the metric case is a strongly local Dirichlet form and in this situation more tools are available (see [7, 28, 58, 59] for various manifestations of this point of view). Let us also mention that metric graphs can be seen as non‐Archimedian analogs of Riemann surfaces, which finds numerous applications in algebraic geometry (see [2, 5, 6, 70] for further references).

The most studied quantum graph operator is the Kirchhoff Laplacian, which provides the analog of the Laplace–Beltrami operator in the setting of metric graphs. Its spectral properties are crucial in connection with the heat equation and the Schrödinger equation and any further analysis usually relies on the self‐adjointness of the Laplacian. Whereas on finite metric graphs the Kirchhoff Laplacian is always self‐adjoint, the question is more subtle for graphs with infinitely many edges. For instance, a uniform lower bound for the edge lengths guarantees self‐adjointness (see [9, 67]), but this commonly used condition is independent of the combinatorial graph structure and clearly excludes a number of interesting cases (the so‐called fractal metric graphs). Moreover, most of the results on strongly local Dirichlet forms require completeness of a given metric space with respect to the ‘intrinsic’ metric (cf., for example, [74]), which coincides with the natural path (geodesic) metric in the case of metric graphs. Geodesic completeness (with respect to the natural path metric) guarantees self‐adjointness of the (minimal) Kirchhoff Laplacian, however, this result is far from being optimal (see [27, Section 4] and also Section 2.4). The search for self‐adjointness criteria for infinite quantum graphs is an open and — in our opinion — rather difficult problem.

If the (minimal) Kirchhoff Laplacian is not self‐adjoint, the natural next step is to ask for a description of its self‐adjoint extensions, which corresponds to possible descriptions of the system in quantum mechanics or, if we speak about Markovian extensions, possible descriptions of Brownian motions. Naturally, this question is tightly related to finding appropriate boundary notions for infinite graphs. Our goal in this paper is to investigate the connection between extension theory and one particular notion, namely graph ends, a concept which goes back to the work of Freudenthal [30] and Halin [38] and provides a rather refined way of compactifying graphs. However, the definition of graph ends is purely combinatorial and naturally must be modified to capture the additional metric structure of our setting. Based on the correspondence between graph ends and topological ends of metric graphs, we introduce the concept of ends of finite volume. First of all, it turns out that finite volume ends play a crucial role in describing the Sobolev spaces H1 and H01 on metric graphs. More specifically, we show that the presence of finite volume ends is the only reason for the strict inclusion H01H1 to hold. This in particular provides a surprisingly transparent geometric characterization of the uniqueness of Markovian extensions of the minimal Kirchhoff Laplacian as well as the self‐adjointness of the so‐called Gaffney Laplacian (we are not aware of its analogs either in the manifold setting or in the context of weighted graph Laplacians, cf. [35, 37, 45, 52, 61, 62]). As yet the other manifestation of the fact that finite volume graph ends represent the proper boundary for Markovian extensions of the Kirchhoff Laplacian, we provide a complete description of all finite energy extensions (that is, self‐adjoint extensions with domains contained in H1, and all Markovian extensions clearly satisfy this condition), however, under the additional assumption that there are only finitely many finite volume ends. Let us stress that this class of graphs includes a wide range of interesting models (Cayley graphs of a large class of finitely generated groups, tessellating graphs, rooted antitrees, etc. have exactly one end and in this case there are no finite volume ends exactly when the total volume of the corresponding metric graph is infinite). Moreover, we emphasize that in all those cases the dimension of the space of finite energy extensions is equal to the number of finite volume ends, however, for deficiency indices, that is, the dimension of the space of self‐adjoint extensions, this only gives a lower bound (for example, for Cayley graphs the dimension of the space of finite energy extensions is independent of the choice of a generating set, although deficiency indices do depend on this choice in a rather nontrivial way). On the other hand, it may happen that these dimensions coincide. The latter holds only if the maximal domain is contained in H1, that is, if every self‐adjoint extension is a finite energy extension. This is further equivalent to the validity of a certain nontrivial Sobolev‐type inequality (see (1.1)). The appearance of this condition demonstrates the mixed dimensional behavior of infinite metric graphs since the analogous estimate holds true in the one‐dimensional situation, but usually fails in the PDE setting.

Let us now sketch the structure of the article and describe its content and our results in greater details.

In Section 2, we collect basic notions and facts about graphs and metric graphs (Subsection 2.1); graph ends (Subsection 2.2); the minimal and maximal Kirchhoff Laplacians (Subsection 2.3); deficiency indices and their connection with the spaces of L2 harmonic and λ‐harmonic functions (Subsection 2.4).

The core of the paper is Section 3, where we discuss the Sobolev spaces H1(G) and H01(G) and introduce the set of finite volume ends C0(G) (Definition 3.8). We show that C0(G) is the proper boundary for H1 functions, which can also be seen as an ideal boundary by applying C*‐algebra techniques (see Remark 3.14). The central result of this section is Theorem 3.12, which shows that H1(G)=H01(G) if and only if there are no finite volume ends. The latter also leads to a surprisingly transparent geometric characterization of the uniqueness of Markovian extensions of the Kirchhoff Laplacian (Corollary 5.5) as well as the self‐adjointness of the Gaffney Laplacian HG (see Remark 5.6(ii) for details and the definition of HG).

Section 4 contains further applications of the above considerations. Namely, Theorem 4.1 demonstrates that deficiency indices of the minimal Kirchhoff Laplacian can be estimated from below by the number of finite volume ends. This estimate is sharp (for example, if there are infinitely many finite volume ends) and we also find necessary and sufficient conditions for the equality to hold. In particular, if there are only finitely many ends of finite volume, #C0(G)<, the latter is equivalent to the validity of the following Sobolev‐type inequality (see Remark 4.2)

fL2(G)C(fL2(G)+fL2(G)) (1.1)

for all f in the maximal domain of the Kirchhoff Laplacian. Metric graphs are locally one‐dimensional and the corresponding inequality is trivially satisfied in the one‐dimensional case, however, globally infinite metric graphs are more complex and hence (1.1) rather resembles the multi‐dimensional setting of PDEs (in particular, (1.1) does not hold true if G has a non‐free finite volume end, see Proposition 4.9).

In the next sections, we focus on a particular class of self‐adjoint extensions whose domains are contained in H1 (we call them finite energy extensions). These extensions have good properties and their importance stems from the fact that they contain the class of Markovian extensions (they also arise as self‐adjoint restrictions of the Gaffney Laplacian). In Section 5, we show that (under some additional mild assumptions) their resolvents and heat semigroups are integral operators with continuous, bounded kernels and they belong to the trace class if G has finite total volume (Theorems 5.1 and 5.2).

In Section 6, we proceed further and show that finite volume ends is the proper boundary for this class of extensions. Namely, under the additional and rather restrictive assumption of finitely many ends with finite volume, in Subsections 6.1 and 6.2, we introduce a suitable notion of a normal derivative at graph ends (as a by‐product, this also gives an explicit description of the domain of the Neumann extension, see Corollary 6.7). Section 6.3 contains a complete description of finite energy extensions and also of Markovian extensions (Theorem 6.11). Let us stress that the case of infinitely many ends is incomparably more complicated and will be the subject of future work.

In general, the inequality in (1.1) is difficult to verify/contradict and even simple examples can exhibit rather complicated behavior (see Appendix B). The only reason for which (1.1) fails to hold is the presence of L2 harmonic functions having infinite energy, that is, not belonging to H1. Moreover, in order to compute deficiency indices of the Kirchhoff Laplacian one, roughly speaking, needs to find the dimension of the space of L2 harmonic functions and description of self‐adjoint extensions requires a thorough understanding of the behavior of L2 harmonic functions at ‘infinity’. Dictated by a distinguished role of harmonic functions in analysis, there is an enormous amount of literature dedicated to various classes of harmonic functions (positive, bounded, etc.), which is further related to different notions of boundaries (metric completion, Poisson and Martin boundaries, Royden and Kuramochi boundaries, etc.) and search for a suitable notion in this context (namely, L2 harmonic functions) is a highly nontrivial problem, which seems not to be very well‐studied either in the context of incomplete manifolds (cf. [61, 62]) or in the case of weighted graphs (see [39, 45]). We further illustrate this by considering the case of rooted antitrees, a special class of infinite graphs with a particularly high degree of symmetry (see Section 7). Infinite rooted antitrees have exactly one graph end, which makes them a good toy model for our purposes. The above considerations show that the space of finite energy L2 harmonic functions is nontrivial only if a given metric antitree has finite total volume and in this case the only such functions are constants. However, adjusting lengths in a suitable way for a concrete polynomially growing antitree (Figure 1) we can make the space of L2 harmonic functions as large as we please (even infinite dimensional!).

FIGURE 1.

FIGURE 1

Antitree with sphere numbers sn=n+1

Notation

Z, R, C have their usual meaning; Za:=Z[a,).

z* denotes the complex conjugate of zC.

For a given set S, #S denotes its cardinality if S is finite; otherwise we set #S=.

If it is not explicitly stated otherwise, we shall denote by (xn) a sequence (xn)n=0.

Cb(X) is the space of bounded, continuous functions on a locally compact space X.

C0(X) is the space of continuous functions vanishing at infinity.

For a finite or countable set X, C(X) is the set of complex‐valued functions on X.

Gd=(V,E) is a discrete graph (satisfying Hypothesis 2.1).

G=(Gd,|·|) is a metric graph (see p. 6).

ϱ is the natural (geodesic) path metric on G (see p. 6).

ϱm is the star metric on V corresponding to the star weight m (see (2.13)).

Ω(Gd) denotes the graph ends of Gd (see Definition 2.1).

C(G) denotes the topological ends of a metric graph G (see Definition 2.2).

C0(G) stays for the finite volume topological ends of G (see Definition 3.8).

G^ is the end (Freudenthal) compactification of G (see p. 7).

H00 is the pre‐minimal Kirchhoff Laplacian on G (see (2.9)).

H0 is the minimal Kirchhoff Laplacian, the closure of H00 in L2(G) (see (2.9)).

n±(H0) are the deficiency indices of H0 (see (2.15)).

HF and HN are the Friedrichs and Neumann extensions of H0 (see p. 12 and, respectively, p. 24).

H is the maximal Kirchhoff Laplacian on G (see (2.8)).

2. QUANTUM GRAPHS

2.1. Combinatorial and metric graphs

In what follows, Gd=(V,E) will be an unoriented graph with countably infinite sets of vertices V and edges E. For two vertices u, vV we shall write uv if there is an edge eu,vE connecting u with v. For every vV, we denote the set of edges incident to the vertex v by Ev and

degG(v):=#{e|eEv} (2.1)

is called the degree (valency or combinatorial degree) of a vertex vV. When there is no risk of confusion about which graph is involved, we shall simplify and write deg instead of degG. A path P of length nZ0{} is a sequence of vertices (v0,v1,,vn) such that vk1vk holds true for all k{1,,n}. The following assumption is imposed throughout the paper.

Hypothesis 2.1

Gd is infinite, locally finite (deg(v)< for every vV), connected (for any two vertices u,vV there is a path connecting u and v), and simple (there are no loops or multiple edges).

Next, let us assign each edge eE a finite length |e|(0,). We can then naturally associate with (Gd,|·|)=(V,E,|·|) a metric space G: first, we identify each edge eE with a copy of the interval Ie:=[0,|e|]. The topological space G is then obtained by ‘gluing together’ the ends of edges corresponding to the same vertex v (in the sense of a topological quotient, see, for example, [13, Chapter 3.2.2]). The topology on G is metrizable by the natural path metric ϱ — the distance between two points x,yG is defined as the arc length of the ‘shortest path’ connecting them (if x or y are not vertices, then we need to allow also paths which start or end in the middle of edges; the length of such paths is naturally defined by taking the corresponding portion of the interval). The metric space G arising from the above construction is called a metric graph (associated to (Gd,|·|)=(V,E,|·|)).

Note that, by definition, (G,ϱ) is a length space (see [13, Chapter 2.1] for definitions and further details). Moreover (see, for example, [40, Chapter 1.1]), a metric graph G is a Hausdorff topological space with countable base and each xG has a neighborhood isometric to a star‐shaped set S(deg(x),rx) of degree deg(x)Z1,

S(deg(x),rx):={z=re2πik/deg(x)|r[0,rx),k=1,,deg(x)}C. (2.2)

Note that deg(x) in (2.2) coincides with the combinatorial degree if x belongs to the vertex set, and deg(x)=2 for every non‐vertex point x of G.

Sometimes, we will consider Gd as a rooted graph with a fixed root oV. In this case, we denote by Sn, nZ0 the nth combinatorial sphere with respect to the order induced by o (note that S0={o}).

2.2. Graph ends

One possible definition of a boundary for an infinite graph is the notion of the so‐called graph ends (see [30; 38; 76, Section 21]).

Definition 2.1

A sequence of distinct vertices (vn)nZ0 (respectively, (vn)nZ) which satisfies vnvn+1 for all nZ0 (respectively, for all nZ) is called a ray (respectively, double ray). A subsequence of such a sequence is called a tail.

Two rays R1,R2 are called equivalent — and we write R1R2 — if there is a third ray containing infinitely many vertices of both R1 and R2. An equivalence class of rays is called a graph end of Gd and the set of graph ends will be denoted by Ω(Gd). Moreover, we will write Rω whenever R is a ray belonging to the end ωΩ(Gd).

An important feature of graph ends is their relation to topological ends of a metric graph G.

Definition 2.2

Consider sequences U=(Un)n=0 of non‐empty open connected subsets of G with compact boundaries and such that Un+1Un for all n0 and n0Un¯=. Two such sequences U and U are called equivalent if for all n0 there exist j and k such that UnUj and UnUk. An equivalence class γ of sequences is called a topological end of G and C(G) denotes the set of topological ends of G.

For locally finite graphs, there is a bijection between topological ends of a metric graph C(G) and graph ends Ω(Gd) of the underlying combinatorial graph Gd (see [23, Section 8.6 and also pp. 277–278; 76, Section 21]; for the case of graphs which are not locally finite see [18, 24]).

Theorem 2.3

For every topological end γC(G) of a locally finite metric graph G=(Gd,|·|) there exists a unique graph end ωγΩ(Gd) such that for every sequence U representing γ, each Un contains a ray from ωγ. Moreover, the map γωγ is a bijection between C(G) and Ω(Gd).

Therefore, we may identify topological ends of a metric graph G and graph ends of the underlying graph Gd. We will simply speak of the ends of G. One obvious advantage of this identification is the fact that the definition of Ω(Gd) is purely combinatorial and does not depend on edge lengths.

Definition 2.4

An end ω of a graph Gd is called free if there is a finite set X of vertices which separates ω from all other ends of the graph (that is, the rays of all ends ωω end up in different connected components of VX than the rays of ω).

Remark 2.5

Let us mention several examples.

  • (i)

    Z has two ends both of which are free.

  • (ii)

    ZN has one end for all N2.

  • (iii)

    A k‐regular tree, k3, has uncountably many ends, none of which is free.

  • (iv)

    If Gd is a Cayley graph of a finitely generated infinite group G, then the number of ends of Gd is independent of the generating set and Gd has either one, two, or infinitely many ends. Moreover, Gd has exactly two ends only if G is virtually infinite cyclic (it has a finite normal subgroup N such that the quotient group G/N is isomorphic either to Z or Z2*Z2). These results are due to Freudenthal [30] and Hopf [42] (see also [75]). The classification of finitely generated groups with infinitely many ends is due to Stallings [73]. Let us mention that if G has infinitely many ends, then the result of Stallings implies that it contains a non‐Abelian free subgroup and hence is non‐amenable. For further details we refer to, for example, [32, Chapter 13].

  • (v)

    Let us also mention that by Halin's theorem [38] every locally finite graph Gd with infinitely many ends contains at least one end which is not free.

One of the main features of graph ends is that they provide a rather refined way of compactifying graphs (see [23, Section 8.6; 29; 76]). Namely, we introduce a topology on G^:=GC(G) as follows. For an open subset UG, denote its extension U^ to G^ by

U^:=U{γC(G)|U=(Un)γsuchthatU0U}. (2.3)

Now we can introduce a neighborhood basis of γC(G) as follows

{U^|UGisopen,γU^}. (2.4)

This turns G^ into a compact topological space, called the end (or Freudenthal) compactification of G.

Remark 2.6

Note that an end γC(G) is free exactly when {γ} is open as a subset of C(G) (here C(G) carries the induced topology from G^). This is further equivalent to the existence of a connected subgraph Gγ with compact boundary Gγ such that UnGγ eventually for any sequence U=(Un) representing γ and UnGγ= eventually for all sequences U=(Un) representing an end γγ.

Let us mention that topological ends can be obtained in a constructive way by means of compact exhaustions. Namely, a sequence of connected subgraphs (Fn) of G such that each Fn has finitely many vertices and edges, FnFn+1 for all n0 and nFn=G is called a compact exhaustion of G. Clearly, each Fn may be identified with a compact subset of G. Now iteratively construct a sequence (Un) by choosing in each step a non‐compact, connected component Un of GFn satisfying UnUn1. It is easy to check that each such sequence (Un) defines a topological end γC(G) and in fact all ends γC(G) are obtained by this construction. Note also that the open subsets Un of such representations γ(Un) (actually, their topological closures, since we need to add endpoints of edges which also belong to V(Fn)) can again be identified with connected subgraphs Gn(γ):=Un¯ and we will frequently use this fact.

Let us finish this section with a few more notations. Suppose R is a ray or a finite path without self‐intersections in Gd. We may identify R with a subgraph of Gd and hence with a subset of G, that is, we can consider it as the union of all edges of R. The latter can further be identified with the interval IR:=[0,|R|) of length |R|, where

|R|:=eR|e|.

Also, we need to consider paths — and in particular rays — in G starting or ending at a non‐vertex point. In particular, given a path (v0,v1,,vN) and a point x in the interior of some edge e attached to v0, eev0,v1, we add the interval [x,v0]e to (v0,v1,,vN). For the resulting set, we shall write (x,v0,v1,,vN) and call it a non‐vertex path; and likewise for rays. The set of all non‐vertex rays will be denoted by R(G).

2.3. Kirchhoff Laplacian

Let G be a metric graph satisfying Hypothesis 2.1. Upon identifying every eE with a copy of the interval Ie=[0,|e|], we denote by

L2(e):=L2(Ie;dxe)

the L2‐space for the (unweighted) Lebesgue measure dxe on Ie and introduce the Hilbert space L2(G) of functions f:GC such that

L2(G):=eEL2(e)={f={fe}eE|feL2(e),eEfeL2(e)2<}.

The subspace of compactly supported L2(G) functions will be denoted by

Lc2(G):={fL2(G)|f0onlyonfinitelymanyedgeseE}.

For every eE, consider the maximal operator He,max acting on functions fH2(e) as a negative second derivative. Here and below Hs(e) for s0 denotes the usual Sobolev space on e (see, for example, [12, Chapter 8]). In particular, H0(e)=L2(e) and

H1(e)={fAC(e)|fL2(e)},H2(e)={fH1(e)|fH1(e)}.

This defines the maximal operator on L2(G) by

Hmax:=eEHe,max,He,max=d2dxe2,dom(He,max)=H2(e). (2.5)

If v is a vertex of the edge eE, then for every fH2(e) the following quantities

fe(v):=limxevf(xe),fe(v):=limxevf(xe)f(v)|xev| (2.6)

are well‐defined. Considering G as the union of all edges glued together at certain endpoints, let us equip a metric graph with the Laplace operator. The Kirchhoff (also called standard or Kirchhoff–Neumann) boundary conditions at every vertex vV are then given by

fiscontinuousatv,eEvfe(v)=0. (2.7)

Imposing these boundary conditions on the maximal domain dom(Hmax) yields the maximal Kirchhoff Laplacian

H:=Hmaxdom(H),dom(H)={fdom(Hmax)|fsatisfies(2.7)foranyvV}. (2.8)

Restricting further to compactly supported functions we end up with the pre‐minimal operator

H00:=Hmaxdom(H00),dom(H00)={fdom(Hmax)Lc2(G)|fsatisfies(2.7)foranyvV}. (2.9)

Integrating by parts one obtains

H00f,fL2(G)=G|f(x)|2dx,fdom(H00), (2.10)

and hence H00 is a non‐negative symmetric operator. We call its closure H0:=H00¯ in L2(G) the minimal Kirchhoff Laplacian. The following result is well‐known (see, for example, [16, Lemma 3.9]).

Lemma 2.7

Let G be a metric graph. Then

H0*=H. (2.11)

2.4. Deficiency indices

In the following, we are interested in the question whether H0 is self‐adjoint, or equivalently whether the equality H0=H holds true. Let us recall one sufficient condition. Define the star weight m(v) of a vertex vV by

m(v):=eEv|e|=vol(Ev), (2.12)

and also introduce the star path metric on V by

ϱm(u,v):=infP=(v0,,vn)u=v0,v=vnvkPm(vk). (2.13)

Theorem 2.8

( [27]) If (V,ϱm) is complete as a metric space, then H00 is essentially self‐adjoint and H00¯=H0=H.

If a symmetric operator is not (essentially) self‐adjoint, then the degree of its non‐self‐adjointness is determined by its deficiency indices. Recall that the deficiency subspace Nz(H0) of H0 is defined by

Nz(H0):=ker(H0*z)=ker(Hz),zC. (2.14)

The numbers

n±(H0):=dimN±i(H0)=dimker(Hi) (2.15)

are called the deficiency indices of H0. Note that n+(H0)=n(H0) since H0 is non‐negative. Moreover, H0 is self‐adjoint exactly when n+(H0)=n(H0)=0.

Lemma 2.9

If 0 is a point of regular type for H0, then

n±(H0)=dimker(H). (2.16)

The claim immediately follows from [1, Section 78] or [69, Proposition 3.3]. Indeed, the set of regular points of H0 is an open subset of C. Moreover, by the Krasnoselskii–Krein theorem (see, for example, [1, Section 78] or [69, Proposition 2.4]), dimNz(H0) is constant on each connected component of the set of regular type points of H0. Since H0 is symmetric, each zCR is a point of regular type for H0. Therefore, if 0 is a point of regular type for H0, we immediately get dimker(H)=dimN0(H0)=n+(H0)=n(H0).

Using the Rayleigh quotient, define

λ0(G):=inffdom(H0)f=1H0f,fL2(G)=inffdom(H0)f=1G|f|2dx. (2.17)

Noting that the operator H0 is non‐negative, 0 is a point of regular type for H0 if λ0(G)>0. Thus, we arrive at the following result.

Corollary 2.10

If λ0(G)>0, then (2.16) holds true.

The positivity of λ0(G) is known in the following simple situation.

Corollary 2.11

If G has finite total volume,

vol(G):=eE|e|<, (2.18)

then H0 is not self‐adjoint and (2.16) holds true.

Indeed, by the Cheeger‐type estimate [55, Corollary 3.5(iv)], we have

λ0(G)14vol(G)2>0, (2.19)

and hence (2.16) holds true by Corollary 2.10. Moreover, 1Gker(H), where 1G denotes the constant function on G, and hence

n±(H0)=dim(kerH)1.

It remains to note that H0 is self‐adjoint exactly when n±(H0)=0.

Remark 2.12

By [55, Theorem 3.4], λ0(G)>0 holds true if the isoperimetric constant α(G) of the metric graph G is positive. For antitrees, the isoperimetric constant is tightly related to the structure of its combinatorial spheres (see [56, Theorem 7.1]). If Gd is the edge graph of a tessellation of R2, then positivity of α(G) can be deduced from certain curvature‐type quantities [65].

On the other hand, by [55, Corollary 4.5(i)], λ0(G)>0 holds true if the combinatorial isoperimetric constant of Gd is positive and *(G):=supeE|e|<. For example, this holds true if Gd is an infinite tree without leaves [55, Lemma 8.1] or if Gd is a Cayley graph of a non‐amenable finitely generated group [55, Lemma 8.12(i)].

Finally, let us remark that ker(H)=H(G)L2(G), where H(G) denotes the space of harmonic functions on G, that is, the set of all ‘edgewise’ affine functions satisfying Kirchhoff conditions (2.7) at each vertex vV. Note that every function fH(G) is uniquely determined by its vertex values f:=f|V=(f(v))vV. Recall also the following result (see, for example, [55, Equation (2.32)]).

Lemma 2.13

Let G be a metric graph satisfying the assumptions in Hypothesis 2.1. If fH(G), then fL2(G) if and only if f2(V;m), that is,

vV|f(v)|2m(v)<. (2.20)

Remark 2.14

The above considerations indicate that in order to understand the deficiency indices of the Kirchhoff Laplacian one needs to find the dimension of the space of L2 harmonic (or, more carefully, λ‐harmonic) functions. Moreover, in order to describe self‐adjoint extensions one has to understand the behavior of L2 harmonic functions at ‘infinity’, that is, near a ‘boundary’ of a given metric graph. However, graphs admit a lot of different notions of boundary (ends, Poisson and Martin boundaries, Royden and Kuramochi boundary, etc.) and search for a suitable notion in this context (namely, L2 harmonic functions) is a highly nontrivial problem, which seems to be not very well‐studied neither in the context of incomplete manifolds nor in the case of weighted graphs.

Let us also mention that recently there has been a tremendous amount of work devoted to the study of harmonic functions and self‐adjoint extensions of Laplacians on weighted graph (we only refer to a brief selection of articles [19, 35, 39, 43, 44, 45, 46, 51]).

3. GRAPH ENDS AND H1(G)

This section deals with the Sobolev space H1 on metric graphs. Its importance stems, in particular, from the fact that it serves as a form domain for a large class of self‐adjoint extensions of H0.

3.1. H1(G) and boundary values

First recall that

H1(G)={fL2(G)C(G)|feH1(e)foralleE,fL2(G)2<}, (3.1)

where C(G) is the space of continuous complex‐valued functions on G and

fL2(G)2:=eEfeL2(e)2.

Note that (H1(G),·H1) is a Hilbert space when equipped with the standard norm

fH1(G)2:=fL2(G)2+fL2(G)2=eEfeH1(e)2,fH1(G).

Moreover, dom(H00)H1(G) and we define H01(G) as the closure of dom(H00) with respect to the norm ·H1(G).

Remark 3.1

If H00 is essentially self‐adjoint, then H1(G)=H01(G). However, the converse is not true in general. In fact this equality is tightly connected to the uniqueness of Markovian extensions of H0 and, as we shall see, it is possible to characterize it in terms of topological ends of G (see Corollary 5.5).

Note also that H01(G) is the form domain of the Friedrichs extension HF of H00 and λ0(G) defined by (2.17) is the bottom of the spectrum of HF.

By definition, H1(G) is densely and continuously embedded in L2(G).

Lemma 3.2

H1(G) is continuously embedded in Cb(G)=C(G)L(G) and

f:=supxG|f(x)|CGfH1(G) (3.2)

holds for all fH1(G) with CG:=coth(12supR|R|), where the supremum is taken over all non‐vertex paths without self‐intersections.

For every interval IR the embedding of H1(I) into L(I) is bounded and

supxI|f(x)|C|I|fH1(I) (3.3)

holds for all fH1(I) with C|I|=coth(|I|) (see [60]). Note that we may identify the restriction f|R of fH1(G) to a (non‐vertex) path without self‐intersections R with a function defined on IR=[0,|R|). It is easy to check that upon this identification f|RH1(IR) and (f|R)=f|R.

Suppose now that R is a fixed non‐vertex path without self‐intersections in G. Then for every xG, connecting x and R by some finite non‐vertex path R0, we see that there exists a non‐vertex path without self‐intersections Rx such that xRx and |Rx||R|/2 (if x lies on R already, then the connecting argument is superfluous and we can simply take a portion of R). Applying (3.3) to Rx, we easily deduce the estimate (3.2).

Remark 3.3

The diameter of G (as a metric space (G,ϱ)) is defined by

diam(G):=supx,yGϱ(x,y). (3.4)

Therefore, diam(G)supR|R| and hence CGcoth(12diam(G)).

The above considerations, in particular, imply the following crucial property of H1‐functions: if R=(vn) is a ray, then

f(γR):=limnf(vn)

exists. Indeed, upon the identification of R with the interval IR=[0,|R|), the latter is an immediate corollary of the description of a Sobolev space H1 in one dimension — for a bounded interval this follows from [12, Theorem 8.2] and in the unbounded case see [12, Corollary 8.9]. Moreover, this limit only depends on the equivalence class of R (indeed, for any two equivalent rays R and R there exists a third ray R containing infinitely many vertices of both R and R, which immediately implies that f(γR)=f(γR)=f(γR)). This enables us to introduce the following notion.

Definition 3.4

For every fH1(G) and a (topological) end γC(G), we define

f(γ):=f(γR), (3.5)

where Rωγ is any ray belonging to the corresponding graph end ωγ (see Theorem 2.3). Sometimes we shall also write f(ωγ):=f(γ).

It turns out that (3.5) enables us to obtain an extension by continuity of every function fH1(G) to the end compactification G^ of G (see Subsection 2.2).

Lemma 3.5

Let G be a metric graph and γC(G). If fH1(G), then

limnsupxUn|f(x)f(γ)|=0 (3.6)

for every sequence U=(Un) representing γ.

Let γC(G) and let U=(Un) be a sequence representing γ. Let also

Rn(γ):={RR(G)|RUn}

be the set of all non‐vertex rays contained in Un, n0.

We proceed by case distinction. First, assume that for n sufficiently large, all rays in Rn(γ) have length at most one. If xUn, then there exists a (non‐vertex) ray RxRn(γ) such that Rx=(x,v0,) and its tail R:=(v0,v1,) (see Definition 2.1) belong to ωγ.

By our assumption, |Rx|1 and hence

|f(γ)f(x)|=|f(γRx)f(x)|=|Rxf(y)dy|fL2(Rx)fL2(Un).

Since xUn is arbitrary, this implies

supxUn|f(γ)f(x)|fL2(Un).

Since U=(Un) represents γ, nUn¯= and hence limnfL2(Un)=0. This implies (3.6).

Assume now that for every nZ0 there is a ray RRn(γ) with |R|>1. Take n0 and choose an xUn. We can find a finite (non‐vertex) path without self‐intersections RxUn such that xRx and |Rx|=1/2 (take into account that Un contains at least one ray of length greater than 1). Hence, we get

|f(x)|supyRx|f(y)|C1/2fH1(Rx)C1/2fH1(Un),

where C1/2=coth(1/2) is the constant from (3.3). Since xUn is arbitrary,

supxUn|f(x)|C1/2fH1(Un).

However, nUn¯= and hence supxUn|f(x)|=o(1) as n. It remains to note that f(γ)=0. Indeed, by Theorem 2.3, for every n0 there is a ray Rnωγ such that RnUn and hence

|f(γ)|=|f(γRn)|supxUn|f(x)|=o(1)

as n. This finishes the proof.

Taking into account the topology on G^=GC(G), the next result is a direct consequence of Lemmas 3.2 and 3.5.

Proposition 3.6

Each fH1(G) has a unique continuous extension to the end compactification G^ of G and this extension is given by (3.5). Moreover,

f=supxG^|f(x)|CGfH1(G).

3.2. Nontrivial and finite volume ends

Observe that some ends lead to trivial boundary values for H1 functions. For example, f(γ)=0 for all fH1(G) if ωγΩ(Gd) contains a ray R with infinite length |R|=. On the other hand, it might happen that all rays have finite length, however, f(γ)=0 for all fH1(G) (see, for example, the second step in the proof of Lemma 3.5).

Definition 3.7

A topological end γC(G) is called nontrivial if f(γ)0 for some fH1(G).

We also need the following notion.

Definition 3.8

A topological end γC(G) has finite volume (or, more precisely, finite volume neighborhood) if there is a sequence U=(Un) representing γ such that vol(Un)< for some n. Otherwise, γ has infinite volume. The set of all finite volume ends is denoted by C0(G). Here and below, vol(A) is the Lebesgue measure of a measurable set AG.

Remark 3.9

If C(G) contains only one end, then this end has finite volume exactly when vol(G)<. Analogously, if γC(G) is a free end, then there is a finite set of vertices X separating ωγ from all other ends and hence this end has finite volume exactly when the corresponding connected component Gγ has finite total volume.

If γ is not free, then the situation is more complicated. For example, for a rooted tree G=To the ends are in one‐to‐one correspondence with the rays from the root o and hence one may possibly confuse the notion of a finite/infinite volume of an end with the finite/infinite length of the corresponding ray. More specifically, let γ be an end of To and let Rγ=(o,v1,v2,) be the corresponding ray. For each n1, let Tn be the subtree of To having its root at vn and containing all the ‘descendant’ vertices of vn. Then by definition γ has finite volume (neighborhood) if and only if there is n1 such that the corresponding subtree Tn has finite total volume. In particular, this implies that G would have uncountably many finite volume ends in this case (here we assume for simplicity that all vertices are essential, that is, deg(v)>2 for all vV). In particular, |Rγ|< is a necessary but not sufficient condition for γ to have finite volume.

It turns out that nontrivial and finite volume ends are closely connected.

Theorem 3.10

Let G be a metric graph. Then γC(G) is nontrivial if and only if γ has finite volume. Moreover, for any finite collection of distinct nontrivial ends {γj}j=1N there exists fH1(G)dom(H) such that f(γ1)=1 and f(γ2)==f(γN)=0.

It is not difficult to see that f(γ)=0 for all fH1(G) if γ has infinite volume. Indeed, assuming that there is fH1(G) such that f(γ)0, Lemma 3.5 would imply that there exists U=(Un) representing γ such that

|f(x)||f(γ)|/2>0

for all xUN and some NZ0. However, since vol(UN)=, we conclude that f is not in L2(G) and this gives a contradiction.

Suppose now that γC(G) has finite volume. Take a sequence U=(Un) representing γ with vol(U0)<. Pick a function ϕH2(0,1) such that ϕ(0)=ϕ(0)=ϕ(1)=0 and ϕ(1)=1 and then define f:GC by

f(xe)=1,xeeandbothverticesofeareinU0,0,xeeandbothverticesofearenotinU0,ϕ|xeu||e|,xee=eu,vanduVU0,vU0.

Clearly, fH2(e) for every eE. Moreover, it is straightforward to check that f satisfies Kirchhoff conditions (2.7) at every vV. By assumption, U0 is compact and hence it is contained in finitely many edges. Thus, there are only finitely many edges eE such that one of its vertices belongs to U0 and the other one does not belong to U0. This implies that fL2(G) and, moreover, f0 only on finitely many edges, which proves the inclusion fdom(H)H1(G). Taking into account that f1 on Un for large enough n, we conclude that f(γ)=1 and hence γ is nontrivial.

It remains to prove the second claim. Suppose that γ1,,γNC(G) are distinct nontrivial ends. Then we can find Uj=(Unj), sequences representing γj, j{1,,N}, such that vol(U01)< and U01U0j= for all j=2,,N (see [29, Satz 3] or [24, Lemma 3.1]). Using the above procedure, we can construct a function fdom(H)H1(G) such that supp(f)U0 and f(γ)=1. The latter also implies that f(γ2)==f(γN)=0.

Remark 3.11

If vol(G)=eE|e|<, then all ends have finite volume and the end compactification G^ of G coincides with several other spaces, among them the metric completion of G and the Royden compactification of a related discrete graph (see [35, Corollary 4.22] and also [34, p. 1526]). Note that the natural path metric ϱ can be extended to G^=GC(G) (see [34]). That is, the distance ϱ(x,γ) between a point xG and an end γC(G) is the infimum over all lengths of rays starting at x and belonging to γ. Similarly, the distance ϱ(γ,γ) between two ends is the infimum over the lengths of all double rays with one tail part in γ and the other one in γ. Then (G^,ϱ) is a metric completion of G and G^ is compact and homeomorphic to the end compactification of G (see [34] for further details).

The metric completion was considered in connection with quantum graphs in [16, 17]; however, it can have a rather complicated structure if vol(G)= and a further analysis usually requires additional assumptions. Moreover, there are clear indications that metric completion is not a good candidate for these purposes.

3.3. Description of H01(G)

Recall that the space H01(G) is defined as the closure of dom(H00)H1(G) with respect to ·H1(G). One can naturally conjecture that H01(G) consists of those H1‐functions which vanish on C(G). In fact, the results of the previous two sections enable us to show that this is indeed the case.

Theorem 3.12

Let G be a metric graph and C(G) be its ends. Then

H01(G)={fH1(G)|f(γ)=0forallγC(G)}. (3.7)

First of all, it immediately follows from Proposition 3.6 that fH01(G) vanishes at every end γC(G) (since this holds for each fdom(H00)).

To prove the converse inclusion, we will follow the arguments of the proof of [35, Theorem 4.14]. Namely, suppose that fH1(G) and f(γ)=0 for all γC(G). Without loss of generality, we may assume that f is real‐valued and f0. To prove that fH01(G), it suffices to construct a sequence of compactly supported functions fnH1(G) which converges to f in H1(G). Define ϕn:[0,)[0,) by

ϕn(s):=s1n,ifs1n,0,ifs<1n, (3.8)

and then let fn:G[0,) be the composition fn:=ϕnf, n0. Since ϕn(s)s for all s0 and |ϕn(s)ϕn(t)||st| for all s,t0, |fn(x)||f(x)| and |fn(x)||f(x)| for almost every xG. Hence fnH1(G) and

fnH1(G)fH1(G) (3.9)

for all n. Let us now show that fn has compact support. Indeed, assuming the converse, there exist infinitely many distinct edges ek in E such that fn is non‐zero on each ek. Taking into account (3.8), for each k we can find a non‐vertex point xk on ek such that fn(xk)>1n. Since G^ is compact, the sequence (xk) has an accumulation point xG^. By construction each edge eE contains at most one of the points xk. It follows that xG and hence xG^ is an end. On the other hand, f is continuous on G^ by Proposition 3.6 and thus f(x)1n, which contradicts our assumptions on f.

It remains to show that fn converges to f in H1(G) as n. Taking into account the above properties of fn, we get

ffnL22+ffnL222(fL22+fnL22+fL22+fnL22)4fH12,

and hence by dominated convergence it is enough to show that fnf and fnf pointwise almost everywhere (a.e.) on G. The first claim is clearly true since limnϕn(s)=s for all s[0,). To prove the second claim, suppose that f is differentiable at a non‐vertex point xG. If f(x)>0, then by continuity of f, there is a neighborhood U of x such that fn=f1n holds on U for all sufficiently large n>0. Hence, fn is differentiable at x with fn(x)=f(x) for all large enough n. Finally, if f(x)=0, then for each n there is a neighborhood Un of x such that f1n on Un. Hence fn0 on Un and, in particular, fn is differentiable at x with fn(x)=0. However, since f0 on G and f is differentiable at x, it follows that f(x)=0 as well. This finishes the proof.

Combining Theorem 3.12 with Theorem 3.10, we arrive at the following fact.

Corollary 3.13

The equality H1(G)=H01(G) holds true if and only if all topological ends of G have infinite volume.

Remark 3.14

In the related setting of (weighted) discrete graphs, an important concept is the construction of boundaries by employing C*‐algebra techniques (this includes both Royden and Kuramochi boundaries, see [35, 48, 53, 64, 71] for further details and references). Finite volume graph ends can also be constructed by using this method. Indeed, A:=H1(G)Cb(G) is a subalgebra by Lemma 3.2 and hence its ·‐closure A:=A¯· is isomorphic to C0(X), where X is the space of characters equipped with the weak*‐topology with respect to A. In general, finding X for some concrete C*‐algebra is a rather complicated task. However, it turns out that in our situation X coincides with G:=GC0(G). Indeed, G=GC0(G) equipped with the induced topology of the end compactification G^ is a locally compact Hausdorff space. Proposition 3.6 together with Theorem 3.10 shows that each function fH1(G) has a unique continuous extension to G and this extension belongs to C0(G). Moreover, by Theorem 3.10, H1(G) is point‐separating and nowhere vanishing on G and hence A=C0(G) by the Stone–Weierstrass theorem. Thus, the resulting boundary notion is precisely the space of finite volume graph ends.

Let us also mention that G is compact only if vol(G)< and in this case one can show that the Royden compactification of G as well as its Kuramochi compactification coincide with the end compactification G^ (see [35; 48, Theorem 7.11; 49, p. 215] and also [41, p. 2] for the discrete case).

4. DEFICIENCY INDICES

Intuitively, deficiency indices should be linked to boundary notions for underlying combinatorial graphs. However, spectral properties of the operator H0 also depend on the edge lengths and this suggests that it is difficult to expect a purely combinatorial formula for the deficiency indices n±(H0) of H0. Recall that throughout the paper, we always assume that G satisfies Hypothesis 2.1.

4.1. Deficiency indices and graph ends

The main result of this section provides criteria which allow to connect n±(H0) with the number of graph ends.

Theorem 4.1

Let G be a metric graph and let H0 be the corresponding minimal Kirchhoff Laplacian. Then

n±(H0)#C0(G). (4.1)

Moreover, the equality

n±(H0)=#C0(G) (4.2)

holds true if and only if either #C0(G)= or dom(H)H1(G).

Remark 4.2

Since the map

D:H1(G)L2(G)ff

is bounded, the inclusion dom(H)H1(G) holds true if and only if there is a positive constant C>0 such that

fL2(G)2C(fL2(G)2+fL2(G)2) (4.3)

holds for all fdom(H). It can be shown by examples that (4.3) may fail.

Before proving Theorem 4.1, let us first comment on some of its immediate consequences.

Corollary 4.3

If G is a metric graph with finite total volume vol(G)<, then

n±(H0)#Ω(Gd). (4.4)

Moreover,

n±(H0)=#Ω(Gd) (4.5)

if and only if either G contains a non‐free end (and hence #Ω(Gd)= in this case) or the inclusion ker(H)H1(G) holds.

In fact, we only need to mention that by Halin's theorem [38] (see Remark 2.5(v)) and the finite total volume of G, #C0(G)= only if G contains a non‐free end.

Recall that for a finitely generated group G, the number of graph ends of a Cayley graph is independent of the generating set (see, for example, [32]). Combining this fact with the above statement, we obtain the following result.

Corollary 4.4

Let Gd be a Cayley graph of a finitely generated group G with infinitely many ends. If vol(G)<, then n±(H0)=.

4.2. Proof of Theorem 4.1

The proof of Theorem 4.1 is based on the following observation. Let HF be the Friedrichs extension of H0. Then dom(H) admits the following decomposition

dom(H)=dom(HF)ker(Hz)=dom(HF)Nz(H0), (4.6)

for every z in the resolvent set ρ(HF) of HF (see, for example, [69, Proposition 14.11]). In particular, (4.6) holds for all z(,λ0(G)), where λ0(G)0 is defined by (2.17). Moreover, dom(HF)H01(G) and hence the inclusion dom(H)H1(G) depends only on the inclusion ker(Hz)H1(G) for some (and hence for all) zρ(HF). Let us stress that N0(H0)=ker(H)=H(G)L2(G) and hence in the case λ0(G)>0, one is interested in whether all L2 harmonic functions belong to H1(G) or not, which is known to depend on the geometry of the underlying metric graph.

We also need the fact that functions in Nλ(H0) with λ(,0) can be considered as subharmonic functions and hence they should satisfy a maximum principle.

Lemma 4.5

Suppose G is a metric graph and let λ(,0).

  • (i)
    If fNλ(H0)=ker(Hλ) is real‐valued and f(x0)>0 for some x0G, then
    supxGf(x)=supvVf(v). (4.7)
  • (ii)
    If additionally fH1(G), then
    supxGf(x)=supγC(G)f(γ). (4.8)
  • (iii)
    If (not necessarily real‐valued) fNλ(H0)H1(G) satisfies
    f(γ)=0 (4.9)
    for all γC(G), then f0.
  • (i)
    Let fNλ(H0) be real‐valued. If xG is such that f(x)>0 and eE is an edge with xe, then upon identifying e with the interval Ie=[0,|e|] and taking into account that f=λf on e, we get
    f(y)=f(x)coshλ(yx)+f(x)λsinhλ(yx) (4.10)
    for all ye. If f(x)0, then obviously f(ei)f(x), where ei is the vertex of e identified with the right endpoint of Ie. Similarly, f(eo)f(x) for the other vertex eo of e if f(x)<0. Hence, f attains its maximum on e at the vertices of e, which clearly implies (4.7).
  • (ii)
    Now let vV be a vertex with f(v)>0. By (2.7), there is an edge eEv such that fe(v)0. If uV is the other vertex of e, then by (4.10) we get
    f(u)=f(v)coshλ|e|+fe(v)λsinhλ|e|>f(v).
    Observe that fe(u)<0. Hence, setting v0=v and v1=u and using induction, we can construct a ray R=(vn) such that f(vn+1)>f(vn) for all n0. Since fH1(G), we get
    0<f(v)<limnf(vn)=f(γR)supγC(G)f(γ),
    which proves (4.8).
  • (iii)

    By considering ±f (and splitting into real and imaginary part, if necessary), (4.9) clearly follows from (4.8).

Remark 4.6

Note that the arguments used in the proof of Lemma 4.5(ii) in fact show that functions in Nλ(H0) with λ(,0) admitting positive values on G cannot attain global maxima in G, that is, if f attains a positive value at some xG, then for every compact subgraph GG the following holds

supxGf(x)=supxGGf(x).

Clearly, analogous statements hold true for functions admitting negative values, however, then sup must be replaced with inf.

Lemma 4.7

Suppose G is a metric graph and let λ(,0). Then

dim(Nλ(H0)H1(G))=#C0(G). (4.11)

Using (4.6) with z=λ(,0) and noting that dom(HF)H01(G), Theorems 3.10 and 3.12 imply that dim(Nλ(H0)H1(G))#C0(G). The converse inequality follows from Lemma 4.5(iii), which shows that the mapping f(f(γ))γC0(G) is injective on the subspace Nλ(H0)H1(G).

After all these preparations, we are now in position to complete the proof of Theorem 4.1.

Proof of Theorem 4.1

Observe that the inequality (4.1) immediately follows from (4.6) and (4.11) since n±(H)=dim(Nλ(H0)).

Clearly, the second claim is trivial if #C0(G)=. Hence, it remains to show that in the case #C0(G)< equality (4.2) holds exactly when dom(H)H1(G). Applying (4.6) once again, the inclusion dom(H)H1(G) holds true exactly when Nλ(H0)H1(G). Taking into account once again that n±(H)=dimNλ(H0) and using (4.11), we arrive at the conclusion.

Remark 4.8

Let us mention that one can prove the second claim of Theorem 4.1 in a different way. Namely, if #C0(G)<, then it is possible to reduce the problem to the study of a finite volume graph with a single end.

Let us stress that in the proof of Theorem 4.1 the equivalence of equality (4.2) and the inclusion dom(H)H1(G) was proved in the case when all finite volume ends are free. The next result shows that the inclusion never holds if there is a finite volume end which is not free.

Proposition 4.9

Let G be a metric graph having a finite volume end which is not free. Then there exists a function fdom(H) which does not belong to H1(G).

First observe that we can restrict our considerations to the case of a metric graph G having finite total volume. Indeed, if γ is a non‐free finite volume end of G, then there exists a sequence U=(Un) representing γ such that vol(Un)< for all n. By definition, each Un is open and has compact boundary. Choosing G0G as the subgraph with vertex set V(G0)=VU0 and edge set E(G0)={eE|eU0}, it is easy to see that G0 is a connected finite volume subgraph and γ is a non‐free end of G0 (see also the notion of graph representation of an end in Section 6.1). Moreover, by construction the set G0 of boundary points (here, G0 is seen as a closed subset of G) is finite.

Let GG be a connected, compact subgraph and consider the finitely many connected components of GG. Since G has infinitely many ends, there is a connected component U which contains at least two distinct graph ends γ,γC(G). Following the proof of Theorem 3.10, we readily construct a real‐valued function f=fUdom(H)H1(G) with f(γ)=0, f(γ)=1 and 0f1 on C(G) (in fact, it suffices to choose the corresponding function ϕ with 0ϕ1). Taking into account Theorem 3.12 and decomposition (4.6), we can assume that f belongs to H1(G)Nλ(H0) for some (fixed) λ(,0). However, Lemma 4.5(iii) implies that

f=supxG|f(x)|=supxGf(x)=1.

On the other hand, there exist two rays R, RR(Gd) representing the ends γ and, respectively, γ such that both R, R are contained in U and have the same initial vertex v0. This leads to another estimate

1=|f(γ)f(γ)|=|f(γ)f(v0)+f(v0)f(γ)|=|Rf(x)dxRf(x)dx|2vol(U)fL2(U)2vol(U)fL2(G).

Assume now that (4.3) holds for all functions gNλ(H0). Then · and ·H1 are in fact equivalent norms on Nλ(H0). Indeed, combining (4.3) and the finite volume property, we get

gH12C(gL22+HgL22)=C(1+λ2)gL22C(1+λ2)vol(G)g2

for all gNλ(H0), whereas gCGgH1 by Lemma 3.2. Choosing compact subgraphs Gε with vol(GGε)ε2 (which is possible since G has finite volume), we clearly get vol(Uε)ε2 and hence the above constructed function fε:=fUεH1(G)Nλ(H0) satisfies

fεL2(G)fεL2(Uε)12vol(Uε)12ε.

However, by construction, fε=1, which obviously contradicts to the equivalence of norms · and ·H1 on Nλ(H0) since ε>0 is arbitrary.

We conclude this section by mentioning some explicit examples.

Example 4.10

(Radially symmetric trees) Let G=T be a radially symmetric (metric) tree: that is, a rooted tree T such that for each n0, all vertices in the combinatorial sphere Sn have the same number of descendants dn2 and all edges between the combinatorial spheres Sn and Sn+1 have the same length. It is well‐known that in this case H is self‐adjoint if and only if vol(T)= and deficiency indices are infinite, n±(H0)=, otherwise (see, for example, [15, 72]). Moreover, due to the symmetry assumptions, all graph ends are of finite volume simultaneously. Hence, we arrive at the equality

n±(H0)=#C0(G)=,ifvol(T)<,0,ifvol(T)=.

Moreover, by Theorem 4.1 and Proposition 4.9, the inclusion dom(H)H1(G) holds true if and only if vol(T)=.

Example 4.11

(Radially symmetric antitrees) Consider a metric antitree G=A (see Section 7.1 for definitions) and additionally suppose that A is radially symmetric, that is, for each n0, all edges between the combinatorial spheres Sn and Sn+1 have the same length. Combining [56, Theorem 4.1] (see also Corollary 7.3) with the fact that antitrees have exactly one graph end, #C(A)=1, we conclude that

n±(H0)=#C0(G)=1,ifvol(A)<,0,ifvol(A)=.

In particular, H is self‐adjoint if and only if vol(A)=. Moreover, the inclusion dom(H)H1(G) holds true for all radially symmetric antitrees by Theorem 4.1.

Remark 4.12

Both radially symmetric trees and antitrees are particular examples of the so‐called family preserving metric graphs (see [11] and also [10]). Employing the results from [11], it is in fact possible to extend the conclusions in Examples 4.10 and 4.11 to this general setting. More precisely, for each family preserving metric graph G without horizontal edges, the Kirchhoff Laplacian H is self‐adjoint if and only if vol(G)= and moreover

n±(H0)=#C0(G)=#C(G),ifvol(G)<,0,ifvol(G)=.

If in addition G has finitely many ends, then the inclusion dom(H)H1(G) holds true. On the other hand, if G has infinitely many ends, then dom(H)H1(G) holds true if and only if vol(G)=. The last two statements are again immediate consequences of Theorem 4.1 and Proposition 4.9.

In conclusion, let us also emphasize that the example of the rope ladder graph in Appendix B shows that the assumption on horizontal edges cannot be omitted. More precisely, the rope ladder graph is a family preserving graph in the sense of [10] with exactly one graph end. However, it possesses infinitely many horizontal edges (that is, edges connecting vertices in the same combinatorial sphere) and Example B.5 shows that in general n±(H0)>#C0(G), even if the edge lengths are chosen symmetrically to the root, |en+|=|en| for all nZ0.

5. PROPERTIES OF SELF‐ADJOINT EXTENSIONS

The Sobolev space H1(G) plays a distinctive role in the study of self‐adjoint extensions of the minimal operator H0. A self‐adjoint extension H of H0 is called a finite energy extension if its domain is contained in H1(G), that is, every function fdom(H) has finite energy, fL2(G)<. The main result of this section already indicates that finite energy self‐adjoint extensions of the minimal operator (note that among those are the Friedrichs extension and, as we will see later in this section, all Markovian extensions) possess a number of important properties.

Theorem 5.1

Let H be a self‐adjoint lower semi‐bounded extension of H0. Assume that z belongs to its resolvent set ρ(H). Then the following assertions hold.

  • (i)

    If the form domain of H is contained in H1(G), then the resolvent R(z,H) of H is an integral operator whose kernel Kz is both of class L(G×G) and jointly Hölder continuous of exponent β=1/2.

  • (ii)

    If additionally, G has finite total volume, then R(z,H) is of trace class.

  • (i)
    Let H be a self‐adjoint lower semi‐bounded extension of H0, Hc for some cR. Without loss of generality, we may assume c=0. Then we can consider its positive semi‐definite square root H1/2, which is again self‐adjoint and whose domain agrees with the form domain of H. Accordingly, for all zC[0,) and λ=z we get
    H1/2λH1/2+λ=Hz,
    and hence
    R(z,H)=R(λ,H1/2)R(λ,H1/2). (5.1)
    If the form domain of H is contained in H1(G), and hence by Lemma 3.2 in Cb(G), then R(±λ,H1/2) maps L2(G) into L(G), and hence by duality also maps L1(G) into L2(G). Thus, (5.1) implies that R(z,H) maps L1(G) into L(G) and hence, by the Kantorovich–Vulikh theorem (see, for example, [4, Theorem 1.3] or [63, Theorem 1.1]), R(z,H) is an integral operator with the L‐kernel K(z;·,·).
    To prove the assertion about joint Hölder continuity, we need to take a closer look at the kernel K by adapting the proof of [3, Proposition 2.1]: as noticed before, the resolvent R(λ,H1/2) is bounded from L2(G) to L(G) by Lemma 3.2 for any λ in the resolvent set of H1/2. Applying the Kantorovich–Vulikh theorem (see, for example, [4, p. 113]) once again, we see that
    R(λ,H1/2)u(x)=Gu(y)κ(λ;x,y)dy=u,κ(λ;x,·)*L2(G)
    for all xG and some κ(λ;x,·)L2(G) such that supxGκ(λ;x,·)L2(G)<. Moreover, observe that there exists C=C(λ)>0 such that
    κ(λ;x,·)κ(λ;x,·)L2(G)Cϱ(x,x) (5.2)
    for all x,xG, where ϱ(x,x) denotes the distance in the natural path metric on G. Indeed, for any function uL2(G),
    |Gu(y)(κ(λ;x,y)κ(λ;x,y))dy|=|R(λ,H1/2)u(x)R(λ,H1/2)u(x)|ϱ(x,x)R(λ,H1/2)uH1Cϱ(x,x)uL2, (5.3)
    where we have used the Cauchy–Schwarz inequality and the fact that the resolvent R(λ,H1/2) is a bounded operator from L2 to the domain of H1/2 equipped with the graph norm, and (5.2) immediately follows. Now, taking into account R(λ,H1/2)*=R(λ*,H1/2) and the equalities (5.1), we conclude that
    R(z,H)u(x)=R(λ,H1/2)R(λ,H1/2)u(x)=R(λ,H1/2)u,κ(λ;x,·)*L2(G)=u,R(λ*,H1/2)κ(λ;x,·)*L2(G)=Gu(y)Gκ(λ;x,s)κ(λ*;y,s)*dsdy=:Gu(y)K(z;x,y)dy
    for all uL2(G). It remains to prove that the mapping
    K:G×G(x,y)Gκ(λ;x,s)κ(λ*;y,s)*dsC
    is jointly Hölder continuous. However, recalling that supxGκ(λ;x,·)L2(G)<, this immediately follows from (5.2), since
    |K(x,y)K(x,y)|κ(λ;x,·)(κ(λ*;y,·)*κ(λ*;y,·)*)L1+κ(λ*;y,·)*(κ(λ;x,·)κ(λ;x,·))L1
    for all pairs (x,y),(x,y)G×G.
  • (ii)

    If G has finite total volume, then L(G×G)L2(G×G) and hence the resolvents R(±λ,H1/2) are Hilbert–Schmidt operators. Thus, by (5.1) we conclude that R(z,H) is of trace class.

Observe that the first step in the proof of Theorem 5.1 is the factorization (5.1), which has the natural counterpart for semigroups

ezHezH=e2zH,Rez>0.

Because the semigroup generated by a self‐adjoint semi‐bounded extension H is analytic, it is a bounded operator from the Hilbert space into its generator's form domain whenever Rez>0. A careful look at the proof of Theorem 5.1 shows that this is sufficient to establish that ezH is an integral operator; all further steps in the proof of Theorem 5.1 carry over almost verbatim to the study of semigroups. We can hence easily deduce the following result.

Theorem 5.2

Let H be a self‐adjoint lower semi‐bounded extension of H0 and let zC with Rez>0. Then the following assertions hold.

  • (i)

    If the domain of H is contained in H1(G), then the semigroup ezH generated by H is an integral operator whose kernel is both of class L(G×G) and jointly Hölder continuous of exponent β=1/2.

  • (ii)

    If additionally, G has finite total volume, then ezH is of trace class.

Estimating as in (5.3) and using analyticity of ezH yields the inequality

|pt(x,y)pt(x,y)|Ctϱ(x,x),t>0,x,y,xG, (5.4)

for the heat kernel pt(x,y) of a non‐negative extension H, where in contrast to (5.3) the constant C>0 is independent of t>0. Such Hölder estimates are known to be related to Sobolev‐type inequalities and also important for upper and lower Gaussian bounds (cf., for example, [20; 66, Chapter 6]). However, we do not pursue this line of study here and this will be done elsewhere.

Remark 5.3

A few remarks are in order.

  • (i)

    If supR|R|<, where the supremum is taken over all non‐vertex paths without self‐intersections, then the path metric ϱ has a natural extension ϱ^ to the end compactification G^. Moreover, in this case (G^,ϱ^) coincides with the metric completion of (G,ϱ). Indeed, the metric completion of (G,ϱ) is obtained by adding to G equivalence classes of rays of finite length (see [34, Section 2.3] for details) and the distance of xG to a boundary point is defined as the ‘shortest length’ of a path in the corresponding equivalence class starting at x.

    Therefore, Theorems 5.1 and 5.2 imply that in this case the corresponding resolvent and semigroup kernels have a bounded and uniformly continuous extension to (G^,ϱ^). However, we stress that in contrast to the case vol(G)< (see Remark 3.11), the topology generated by ϱ^ on G^ can differ from the end compactification topology if vol(G)=.

  • (ii)

    Discreteness of the spectrum of the Friedrichs extension HF is a standard fact in the case of finite total volume (see, for example, [16, Proposition 3.11] or [56, Corollary 3.5(iv)]). However, Theorem 5.1(ii) implies the stronger assertion that the resolvent of HF belongs to the trace class if vol(G)<. Let us also stress that it is not true in general that every self‐adjoint extension of H will have a discrete spectrum if vol(G)<, since in case of infinite deficiency indices such a self‐adjoint extension could have a domain large enough to make compactness of the embedding of H1(G) into L2(G) irrelevant.

Recall that a self‐adjoint extension H of H0 is called Markovian if H is a non‐negative self‐adjoint extension and the corresponding quadratic form is a Dirichlet form (for definitions and further details, we refer to [31, Chapter 1]). Hence, the associated semigroup etH, t>0, as well as resolvents R(λ,H), λ>0, are Markovian: that is, are both positivity preserving (map non‐negative functions to non‐negative functions) and L ‐contractive (map the unit ball of L(G), and then by duality of Lp(G) for all p[1,], into itself). Let us stress that the Friedrichs extension HF of H0 is a Markovian extension. Consider also the following quadratic form in L2(G)

tN[f]=G|f(x)|2dx,dom(tN)=H1(G). (5.5)

This form is non‐negative and closed, hence we can associate in L2(G) a self‐adjoint operator with it, let us denote it by HN. We will refer to it as the Neumann extension. It is straightforward to check that tN is a Dirichlet form and HN is also a Markovian extension of H0.

It turns out that Theorems 5.1 and 5.2 apply to all Markovian extensions of H0. More specifically, the analog of the results for discrete Laplacians [39, Theorem 5.2] and Laplacians in Euclidean domains [31, Chapter 3] and Riemannian manifolds [37, Theorem 1.7] holds true for quantum graphs as well.

Theorem 5.4

If H is a Markovian extension of H0, then dom(H)H1(G) and, moreover,

HNHHF, (5.6)

where the inequalities are understood in the sense of forms.

We omit the proof of Theorem 5.4 since the proofs of either [39, Theorem 5.2] or [37, Lemma 3.6] carry over verbatim to our setting (see also the proof of [31, Theorem 3.3.1]).

Let us finish this section with the following observation.

Corollary 5.5

The following are equivalent:

  • (i)

    H0 has a unique Markovian extension,

  • (ii)

    H01(G)=H1(G),

  • (iii)

    all topological ends of G have infinite volume, C0(G)=.

The claimed equivalences follow from Theorem 5.4 and Corollary 3.13.

Remark 5.6

Let us finish this section with a few comments.

  • (i)

    The equivalence (i)(ii) in Corollary 5.5 is known for Riemannian manifolds [37, Theorem 1.7] (see also [31, Chapter 3; 62, Theorem 1]) as well as for weighted Laplacians on graphs [39, Corollary 5.6]. However, to the best of our knowledge these settings do not admit any further geometric characterization.

  • (ii)
    The list of equivalences in Corollary 5.5 can be extended by adding a claim on the self‐adjointness of the so‐called Gaffney Laplacian. Namely, since H01(G) and H1(G) are Hilbert spaces, the operators denoted by D and N and defined in L2(G) on the domains, respectively, H01(G) and H1(G) by ff are closed. Note that with this notation at hand we have HF=D*D and HN=N*N. Now we can introduce the Gaffney Laplacian HG:=D*N as the restriction of the maximal operator H onto the domain (compare with [37, p. 610] for the definition in the manifolds case)
    dom(HG):={fH1(G)|Nfdom(D*)}. (5.7)
    Clearly, dom(HF)dom(HG), dom(HN)dom(HG), and HG is not necessarily symmetric. It turns out that HG is symmetric (and hence self‐adjoint) if and only if the Kirchhoff Laplacian H0 has a unique Markovian extension. Moreover, in this case HF=HN=HG (cf. [37, Theorem 1.7(ii)] in the manifold setting). Let us mention that the Markovian/finite energy extensions of H0 are exactly the Markovian/self‐adjoint restrictions of HG and hence the deficiency indices of HG*=N*D are equal to #C0(G).

6. FINITE ENERGY SELF‐ADJOINT EXTENSIONS

It turns out that finite volume (topological) ends provide the right notion of the boundary for metric graphs to deal with finite energy and also with Markovian extensions of the minimal Kirchhoff Laplacian H0. In particular, we are going to show that this end space is well‐behaved as concerns the introduction of both traces and normal derivatives. More specifically, the goal of this section is to give a description of finite energy self‐adjoint extensions of H0 in the case when the number of finite volume ends of G is finite, that is, #C0(G)<. Note that in this case all finite volume ends are free.

6.1. Normal derivatives at graph ends

Let G=(V,E) be a (possibly infinite) connected subgraph of G. Recall that its boundary G (with respect to the natural topology on G, see Subsection 2.1) is given by

G={vV|degG(v)<degG(v)}. (6.1)

For a function fdom(H), we define its (inward) normal derivative at vG by

fnG(v):=eEvEfe(v). (6.2)

With this definition at hand, we end up with the following useful integration by parts formula.

Lemma 6.1

Let G be a compact (not necessarily connected) subgraph of the metric graph G. Then

Gf(x)g(x)dx=Gf(x)g(x)dx+vGg(v)fnG(v) (6.3)

for all fdom(H) and gH1(G). In particular,

Gf(x)dx=vGfnG(v). (6.4)

The claim follows immediately from integrating by parts, taking into account that f satisfies (2.7). Setting g1 in (6.3), we arrive at (6.4).

To simplify our considerations, we need to introduce the following notion. Let γC(G) be a (topological) end of G. Consider a sequence (Gn) of connected subgraphs of G such that GnGn+1 and #Gn< for all n. We say that the sequence (Gn) is a graph representation of the end γC(G) if there is a sequence of open sets U=(Un) representing γ such that for each n0 there exist j and k such that GnUj and UnGk. It is easily seen that all graphs Gn are infinite (they have infinitely many edges). Moreover, graph representations (Gn) of an end can be constructed with the help of compact exhaustions; in particular each graph end γC(G) has a representation by subgraphs (see Subsection 2.2).

Proposition 6.2

Let G be a metric graph and let γC(G) be a free end of finite volume. Then for every function fdom(H) and any sequence (Gk) of subgraphs representing γ, the limit

limkvGkfnGk(v) (6.5)

exists and is independent of the choice of (Gk).

First of all, note that uniqueness of the limit follows from the inclusion property in the definition of the graph representations of γ. Hence, we only need to show that the limit in (6.5) indeed exists.

Let (Gk) be a graph representation of a free finite volume end γC0(G). Since γ is free, we can assume that vol(G0)< and that G0Uk= eventually for every sequence U=(Uk) representing an end γγ. First observe that G=GkGj can again be identified with a compact subgraph of G whenever kj. Indeed, if G has infinitely many edges {en}E, choose for each n a point xn in the interior of the edge en. Since G^=GC(G) is compact, the set {xn} has an accumulation point xG^. By construction, xG and hence xG^G=C(G) is an end. However, we have that xnGj and recalling (2.3) and (2.4), this implies that x=γ for a topological end γγ. On the other hand, xnG0 for all n and using the properties of G0 and (2.3)–(2.4) once again, we arrive at a contradiction.

Now, using (6.1) it is straightforward to verify that

vGkfnGk(v)vGjfnGj(v)=vGfnG(v).

Hence by (6.4) and the Cauchy–Schwarz inequality, we get

vGkfnGk(v)vGjfnGj(v)=GkGjf(x)dxvol(Gk)HfL2(G), (6.6)

whenever kj. This implies the existence of the limit in (6.5) since vol(Gk)=o(1) as k.

Proposition 6.2 now enables us to introduce a normal derivative at graph ends.

Definition 6.3

Let γC(G) be a free end of finite volume and let (Gk) be a graph representation of γ. Then for every fdom(H)

nf(γ):=fn(γ):=limkvGkfnGk(v) (6.7)

is called the normal derivative of f at γ.

Remark 6.4

In fact, it is not difficult to extend the definitions (6.2) and (6.7) to general sequences U=(Un) of open sets representing the free end γC0(G). However, while the idea of the proof of Proposition 6.2 naturally carries over, the analysis becomes more technical and we restrict to the case of subgraphs for the sake of a clear exposition.

Let us mention that the normal derivative can also be expressed in terms of compact exhaustions.

Lemma 6.5

Let G be a metric graph having finite total volume and only one end γ, C(G)={γ}. If (Fk) is a compact exhaustion of G and fdom(H), then

nf(γ)=limkvFkfnFk(v). (6.8)

The fact that we are not approximating γ by its neighborhoods, but rather by compact subgraphs, is responsible for the different sign in (6.7) and (6.8).

First of all, note that GFk can be identified with a subgraph of G and

vFkfnFk(v)=v(GFk)fnGFk(v)

for all fdom(H). If, moreover, GFk is a connected subgraph for all k0, then it is clear that (Gk) with Gk:=GFk for all k0, is a graph representation of γ and this proves (6.8) in this case.

If GFk is not connected, then it has only one infinite connected component Gkγ and finitely many compact components (since C(G)={γ}). Adding these compact components to Fk, we obtain a compact exhaustion (Fk) with GFk=Gkγ. Arguing as in the proof of Proposition 6.2 (see (6.6)), we get

|vFkfnFk(v)vFkfnFk(v)|=|FkFkf(x)dx|=o(1)

as k. Hence, (6.8) holds true also in the general case.

6.2. Properties of the trace and normal derivatives

In this section, we collect some basic properties of the trace maps. We shall adopt the following notation. Since we shall always assume throughout this section that #C0(G)<, we set H:=2(C0(G)), which can be further identified with C#C0(G). Next, we introduce the maps Γ0:H1(G)H and Γ1:dom(H)H1(G)H by

Γ0:ff(γ)γC0(G),Γ1:fnf(γ)γC0(G), (6.9)

where the boundary values and normal derivative of f are defined by (3.4) and (6.7), respectively.

Proposition 6.6

Let G be a metric graph with #C0(G)<. Then,

  • (i)
    for every f^H, there exists fdom(H)H1(G) such that
    Γ0f=f^,Γ1f=0;
  • (ii)
    moreover, the Gauss–Green formula
    Hf,gL2(G)=f,gL2(G)Γ1f,Γ0gH (6.10)
    holds true for every fdom(H)H1(G) and gH1(G).
  • (i)
    Since #C0(G)<, each finite volume end γC0(G) is free. For every γC0(G), let Gγ be a subgraph with the properties as in Remark 2.6. We can also assume that vol(Gγ)<. Following the proof of Theorem 3.10, we can construct for each end γC0(G) a function fγdom(H)H1(G) such that fγ is non‐constant only on finitely many edges (since #Gγ<), fγ(γ)=1 and fγ(γ)=0 for all other ends γC0(G){γ}. Clearly, Γ1fγ=0 for every γC0(G). Thus, setting
    f=γC0(G)f^(γ)fγ
    for a given f^H, we clearly have Γ0f=f^ and Γ1f=0.
  • (ii)
    Let us first show that (6.10) holds true for all fdom(H)H1(G) if g=fγH1(G). Take a compact exhaustion (Fk) of G. Then by Lemma 6.1,
    Hf,fγL2(G)f,fγL2(G)=limkHf,fγL2(Fk)f,fγL2(Fk)=limkvFkfnFk(v)fγ(v)*=limkvFkVγfnFk(v),
    where Vγ is the set of vertices of Gγ. Note that the subgraph Gγ itself is a connected infinite graph having finite total volume and exactly one end, which can be identified with γ in an obvious way. Moreover, setting Fkγ:=FkGγ for all k0 and noting that Fkγ is connected for all sufficiently large k, the sequence (Fkγ) provides a compact exhaustion of Gγ. Since GγFkγ=FkVγ and
    fnFkγ(v)=fnFk(v),vGγFkγ,
    for all large enough k0, we get by applying Lemma 6.5
    Hf,fγL2(G)f,fγL2(G)=limkvFkVγfnFkγ(v)=fn(γ).
    Hence, (6.10) holds true if g=fγH1(G).

    Now observe that a simple integration by parts implies that (6.10) is valid for all compactly supported gH1(G). By continuity and Theorem 3.12, this extends further to all gH01(G). Finally, setting g:=gγC0(G)g(γ)fγ for gH1(G), it is immediate to check that, by Theorem 3.12, gH01(G). It remains to use the linearity of Γ0.

It turns out that the domain of the Neumann extension admits a simple description.

Corollary 6.7

Let G be a metric graph with #C0(G)<. Then the Neumann extension HN is given as the restriction HN=H|dom(HN) to the domain

dom(HN)={fdom(H)H1(G)|Γ1f=0}. (6.11)

By the first representation theorem [50, Chapter VI.2.1], dom(HN) consists of all functions fH1(G) such that there exists hL2(G) with

f,gL2(G)=h,gL2(G),forallgH1(G).

Moreover, in this case HNf:=h. Taking into account Proposition 6.6 and the fact that HN is a restriction of H, we immediately arrive at (6.11).

Our next goal is to prove surjectivity of the normal derivative map.

Proposition 6.8

If G is a metric graph with #C0(G)<, then the mapping Γ1 is surjective.

In fact, Proposition 6.8 will follow from the following lemma.

Lemma 6.9

Suppose G is a metric graph with vol(G)< and only one end, C(G)={γ}. Then there exists fdom(H)H1(G) such that

nf(γ)0.

We will proceed by contradiction. Suppose that ng(γ)=0 for all gdom(H)H1(G). Then, by Corollary 6.7, dom(HF)dom(HN)=dom(H)H1(G). However, both HF and HN are self‐adjoint restrictions of H and hence dom(HF)=dom(HN). Therefore, HF=HN and their quadratic forms also coincide, which implies that H01(G)=H1(G). This contradicts Corollary 3.13 and hence completes the proof.

Proof of Proposition 6.8

Let Gγ, γC0(G) be the subgraphs of G constructed in the proof of Proposition 6.6(i). Every Gγ is a connected graph with vol(Gγ)< and only one end, which can be identified with γ. Hence we can apply Lemma 6.9 to obtain a function gγdom(Hγ)H1(Gγ) such that ngγ(γ)=1. Here Hγ denotes the Kirchhoff Laplacian on Gγ.

Since #Gγ<, we can obviously extend gγ to a function gγ on G such that gγdom(H)H1(G) and gγ is identically zero on a neighborhood of each end γγ (see also the proof of Theorem 3.10). In particular, this implies that ngγ(γ)=0 for all γC0(G){γ}. Upon identification of γ with the single end of Gγ we also have that

ngγ(γ)=ngγ(γ)=1.

This immediately implies surjectivity.

6.3. Description of self‐adjoint extensions

Our next goal is a description of all finite energy self‐adjoint extensions of H0, that is, self‐adjoint extensions H satisfying the inclusion dom(H)H1(G). We will be able to do this under the additional assumption that G has finitely many finite volume ends. Recall that in this case H=2(C0(G)) is a finite‐dimensional Hilbert space.

Let C, D be two linear operators on H satisfying Rofe–Beketov conditions [68]:

CD*=DC*,rank(C|D)=dimH=#C0(G). (6.12)

Consider the quadratic form tC,D defined by

tC,D[f]:=G|f(x)|2dx+D1CΓ0f,Γ0fH (6.13)

on the domain

dom(tC,D):={fH1(G)|Γ0fran(D*)}. (6.14)

Here and in the following the mappings, Γ0 and Γ1 are given by (6.9) and D1:ran(D)ran(D*) denotes the inverse of the restriction D|ker(D):ran(D*)ran(D). In particular, (6.12) implies that tC,D[f] is well‐defined for all fdom(tC,D) (see also (A.4)).

Remark 6.10

It is straightforward to check that tI,0=tF and t0,I=tN are the quadratic forms corresponding to the Friedrichs extension HF and, respectively, Neumann extension HN (see Remark 3.1 and (5.5)).

Now we are in position to state the main result of this section.

Theorem 6.11

Let G be a metric graph with finitely many finite volume ends, #C0(G)<. Let also C, D be linear operators on H satisfying Rofe‐Beketov conditions (6.12). Then,

  • (i)

    the form tC,D given by (6.13), (6.14) is closed and lower semi‐bounded in L2(G);

  • (ii)
    the self‐adjoint operator HC,D associated with the form tC,D is a self‐adjoint extension of H0 and its domain is explicitly given by
    dom(HC,D)={fdom(H)H1(G)|CΓ0f+DΓ1f=0}; (6.15)
  • (iii)

    conversely, if H is a self‐adjoint extension of H0 such that dom(H)H1(G), then there are C,D satisfying (6.12) such that H=HC,D;

  • (iv)

    moreover, H=HC,D is a Markovian extension if and only if the corresponding quadratic form t^C,D[y]=D1Cy,yH, dom(t^)=ran(D*) is a Dirichlet form on H in the wide sense.

  • (i)

    Since H is finite dimensional, it is straightforward to see that the form tC,D is closed and lower semi‐bounded in L2(G) whenever C and D satisfy (6.12).

  • (ii)
    By the first representation theorem [50, Chapter VI.2.1], dom(HC,D) consists of all functions fdom(tC,D)H1(G) for which there exists hL2(G) such that
    f,gL2(G)+D1CΓ0f,Γ0gH=h,gL2(G) (6.16)
    for all gdom(tC,D). Moreover, in this case HC,Df:=h.
    The Gauss–Green identity (6.10) implies that for any fdom(HC,D) and gdom(tC,D),
    D1CΓ0f,Γ0gH=Γ1f,Γ0gH.
    Taking into account the surjectivity property in Proposition 6.6(i), the inclusion ‘’ in (6.15) follows. The converse inclusion is then an immediate consequence of the Gauss–Green identity (6.10).
  • (iii)
    To prove the claim, it suffices to show that
    Θ={(Γ0f,Γ1f)|fdom(H)}H×H
    is a self‐adjoint linear relation (for further details we refer to Appendix A). By definition, Θ* is given by
    Θ*={(g,h)H×H|Γ1f,gH=Γ0f,hHforallfdom(H)}.
    The inclusion ΘΘ* follows immediately from the Gauss–Green identity (6.10) and the self‐adjointness of H. Indeed, we clearly have
    0=Hf,fL2(G)f,HfL2(G)=Γ1f,Γ0fH+Γ0f,Γ1fH
    for all functions f,fdom(H). On the other hand, by Propositions 6.8 and 6.6, for any (g,h)Θ* there is a function fdom(H)H1(G) such that g=Γ0f and h=Γ1f. Employing the identity (6.10) once again, we see that
    Hf,fL2(G)=f,fL2(G)Γ1f,gH=f,fL2(G)Γ0f,hH=f,HfL2(G)
    for all fdom(H). Hence, fdom(H) and in particular (g,h)Θ. Since Θ is self‐adjoint, there are C and D in H satisfying Rofe–Beketov conditions (6.12) and such that Θ={(f,g)H×H|Cf+Dg=0}.
  • (iv)
    The first direction of the equivalence is clear: since the quadratic form tN associated with the Neumann extension HN is Markovian and
    Γ0(φf)=(φf)(γ)γC0(G)=:φ(Γ0f)
    for all functions fH1(G) and every normal contraction φ, the extension HC,D is Markovian if t^C,D is a Dirichlet form on H in the wide sense.
    To prove the converse direction, let, for simplicity, fdom(t^C,D) be real‐valued and fix some real‐valued fH1(G) with Γ0f=f (the existence of such an f follows from Proposition 6.6). For any (real‐valued) normal contraction φ:RR, we can construct a continuous and piecewise affine function ψ:RR (that is, ψ is affine on every component of R{x1,,xM} for finitely many points x1,,xM) such that ψ(0)=0, ψ(f(γ))=φ(f(γ)) for all γC0(G) and |ψ(x)|=1 for almost every xR. Note that every function ψ with the above properties is a normal contraction. Hence, if tC,D is Markovian, it follows that ψfdom(tC,D). However, its boundary values are precisely given by
    Γ0(ψf)=ψf=φf
    and we conclude that φf belongs to dom(t^C,D). Finally, the Markovian property of tC,D implies that
    tC,D[ψf]=G|(ψf)|2dx+t^C,D[φf]tC,D[f]=G|f|2dx+t^C,D[f],
    and noticing that |(ψf)|=|f| almost everywhere on G, the proof is complete.

Let us demonstrate Theorem 6.11 by applying it to Cayley graphs.

Corollary 6.12

Let Gd be a Cayley graph of a finitely generated group G with one end. Then the Kirchhoff Laplacian H0 admits a unique Markovian extension if and only if the underlying metric graph G=(Gd,|·|) has infinite total volume, vol(G)=. Moreover, if G has finite total volume, then the set of all Markovian extensions of H0 forms a one‐parameter family given explicitly by

dom(Hθ)={fdom(H)H1(G)|cos(θ)Γ0f+sin(θ)Γ1f=0}, (6.17)

where θ[0,π/2].

Taking into account that amenable groups have finitely many ends, the above result applies to amenable finitely generated groups, which are not virtually infinite cyclic (see Remark 2.5(iv)). In a similar way one can obtain a complete description of Markovian extensions in the case of virtually infinite cyclic groups, however, they have two ends and the corresponding description looks a little bit more cumbersome and we leave it to the reader (cf. [31, p. 147]). The case of groups with infinitely many ends remains an open highly nontrivial problem.

Remark 6.13

A few remarks are in order.

  • (i)

    Let us mention that in the case when the domain of the maximal operator H is contained in H1(G) and G has finitely many finite volume ends (note that by Theorem 4.1 in this case n±(H0)=#C0(G)<), Proposition 6.11 provides a complete description of all self‐adjoint extensions of H0. Let us also mention that Proposition 6.11 provides a complete description of all self‐adjoint restrictions of the Gaffney Laplacian HG, see Remark 5.6(ii).

  • (ii)

    Some of the results of this section extend (to a certain extent) to the case of infinitely many ends. Let us stress that by Proposition 4.9 in the case when G has a finite volume end which is not free the above results would lead only to some (not all!) self‐adjoint extensions of H0. In our opinion, even in the case of radially symmetric trees having finite total volume the description of all self‐adjoint extensions of H0 is a difficult problem.

  • (iii)

    Similar relations between Markovian realizations of elliptic operators on domains or finite metric graphs (with general couplings at the vertices) on one hand, and Dirichlet property of the corresponding quadratic form's boundary term on the other hand, are of course well‐known in the literature (see, for example, [14, Proposition 5.1; 47, Theorem 3.5; 57, Theorem 6.1]). However, the setting of infinite metric graphs additionally requires much more advanced considerations of combinatorial and topological nature. In particular, it seems noteworthy to us that the results of the previous sections provide the right notion of the boundary for metric graphs, namely, the set of finite volume ends, to deal with finite energy and also with Markovian extensions of the minimal Kirchhoff Laplacian. In particular, this end space is well‐behaved as concerns the introduction of traces and normal derivatives.

  • (iv)
    Taking into account certain close relationships between quantum graphs and discrete Laplacians (see [27, Section 4]), one can easily obtain the results analogous to Theorems 4.1 and 6.11 for a particular class of discrete Laplacians on Gd defined by the following expression
    (τf)(v):=1m(v)uvf(v)f(u)|eu,v|,vV, (6.18)
    where m is the star weight (2.12). Markovian extensions of weighted discrete Laplacians were considered also in [52]. On the other hand, [52] does not contain a finiteness assumption, however, the conclusion in our setting appears to be slightly stronger than in [52, Theorem 3.5], where the correspondence between Markovian extensions and Markovian forms on the boundary is in general not bijective.

7. DEFICIENCY INDICES OF ANTITREES

The main aim of this section is to construct for any NZ1{} a metric antitree such that the corresponding minimal Kirchhoff Laplacian H0 has deficiency indices n±(H0)=N. Our motivation stems from the fact that every antitree has exactly one end and hence, according to considerations in the previous sections, H0 admits at most one‐parameter family of Markovian extensions.

7.1. Antitrees

Let Gd=(V,E) be a connected, simple combinatorial graph. Fix a root vertex oV and then order the graph with respect to the combinatorial spheres Sn, n0 (note that S0={o}). Gd is called an antitree if every vertex in Sn, n1, is connected to all vertices in Sn1 and Sn+1 and no vertices in Sk for all |kn|1 (see Figure 1). Note that each antitree is uniquely determined by its sequence of sphere numbers (sn), sn:=#Sn for n0.

While antitrees first appeared in connection with random walks [25, 54, 77], they were actively studied from various different perspectives in the last years (see [11, 22 56] for quantum graphs and [21, Section 2] for further references).

Let us enumerate the vertices in every combinatorial sphere Sn by (vin)i=1sn and denote the edge connecting vin with vjn+1 by eijn, 1isn, 1jsn+1. We shall always use A to denote (metric) antitrees.

It is clear that every (infinite) antitree has exactly one end. By Theorem 4.1, the deficiency indices of the corresponding minimal Kirchhoff Laplacian are at least 1 if vol(A)<. On the other hand, under the additional symmetry assumption that A is radially symmetric (that is, for each n0, all edges connecting combinatorial spheres Sn and Sn+1 have the same length), it is known that the deficiency indices are at most 1 (see [56, Theorem 4.1] and Example 4.11). It turns out that upon removing the symmetry assumption it is possible to construct antitrees such that the corresponding minimal Kirchhoff Laplacian has arbitrary finite or infinite deficiency indices. More precisely, the main aim of this section is to prove the following result.

Theorem 7.1

Let A be the antitree with sphere numbers sn=n+1, n0 (Figure 1). Then for each NZ1{} there are edge lengths such that the corresponding minimal Kirchhoff Laplacian H0 has the deficiency indices n±(H0)=N.

7.2. Harmonic functions

As it was mentioned already, every harmonic function is uniquely determined by its values at the vertices. On the other hand, fC(V) defines a function fH(A) with f|V=f if and only if the following conditions are satisfied:

j=1sn+1f(vjn+1)f(vkn)|ekjn|+i=1sn1f(vin1)f(vkn)|eikn1|=0, (7.1)

at each vkn, 1ksn with n0. We set s1:=0 for notational simplicity and hence the second summand in (7.1) is absent when n=0. We can put the above difference equations into the more convenient matrix form. Denote fn:=f|Sn=(f(vin))i=1sn for all nZ0 and introduce matrices

Mn+1:=1|e11n|1|e12n|1|e1sn+1n|1|e21n|1|e22n|1|e2sn+1n|1|esn1n|1|esn2n|1|esnsn+1n|Rsn×sn+1, (7.2)

and

Dn:=diag(dkn)Rsn×sn,dkn:=j=1sn+11|ekjn|+i=1sn11|eikn1|, (7.3)

for all nZ0. Note the following useful identity

d10=M11s1,d1ndsnn=Dn1sn=(Mn+1Mn*)1sn+11sn1,n1, (7.4)

where 1sn:=(1,,1)Rsn. Hence, (7.1) can be written as follows

M1f1=j=1s11|e1j0|f0=d10f0, (7.5)
Mn+1fn+1=DnfnMn*fn1,n1. (7.6)

Since Dn is invertible, we get

fn=Dn1(Mn+1Mn*)fn+1fn1 (7.7)

for all n1. In particular, fnran(Dn1(Mn+1Mn*)) for all n1, which implies that the number of linearly independent solutions to the above difference equations (and hence the number of linearly independent harmonic functions) depends on the ranks of the matrices (Mn+1Mn*), n1. Let us demonstrate this by considering the following example.

Lemma 7.2

Let A be a radially symmetric antitree. Then

H(A)=span{1G}. (7.8)

Let for each n0, all edges connecting combinatorial spheres Sn and Sn+1 have the same length, say n>0. Clearly, in this case

ran(Mn+1)=ran(Mn*)=span{1sn},

for all n1. Moreover, each Dn is a scalar multiple of the identity matrix Isn and hence (7.7) implies that fn=cn1sn with some cnC for all n0. Plugging this into (7.5) and (7.6), we get

c1=c0,cn+1=cn+sn1nsn+1n1(cncn1),n1.

Hence, cn=c0=f(o) for all n0, which proves the claim.

The latter in particular implies the following statement (cf. [56, Theorem 4.1]).

Corollary 7.3

If A is a radial antitree with finite total volume, then n±(H0)=1.

By Corollary 2.11, we only need to show that n±(H0)1. However, this is clear since n±(H0)=dim(ker(H))dim(H(A))=1.

7.3. Finite deficiency indices

We restrict our further considerations to a special case of polynomially growing antitrees. Namely, for every NZ1, the antitree AN has sphere numbers s0=1 and sn:=n+N for all nZ1. To define its lengths, pick a sequence of positive numbers (n) and set

|eijn|:=2n,if1i=jN,n,otherwise, (7.9)

for all nZ0.

Lemma 7.4

If a metric antitree AN has lengths given by (7.9), then

dimH(AN)=N+1. (7.10)

Denoting

Bn,m:=111111111Rn×m,Bn:=Bn,nRn×n, (7.11)

we get the following block‐matrix form of the matrices Mn+1:

Mn+1=1nBN12INBN,n+1Bn,NBn,n+1 (7.12)

for all n1. Taking into account (7.3) and denoting

dn1:=n+N3/2n1+n+N+1/2n,dn2:=n+N1n1+n+N+1n,

we get

Dn=dn1INdn2In, (7.13)

for all n2. Since M1R1×(N+1) and

ran(Mn+1)=ran(Mn*)=spanfN1n|fNCN (7.14)

for all n2, (7.7) implies that every f solving (7.5)–(7.6) must be of the form

fn=fnNcn1nCN+n,fnNCN,cnC, (7.15)

for all n1. Plugging (7.15) into (7.6) and taking into account that

BNfnN=f¯nN1N,f¯nN:=fnN,1N=B1,NfnN,

we get after straightforward calculations

f¯n+1N+cn+1(n+1)n1N12nfn+1N=dn1fnNf¯n1N+cn1(n1)n11N+12n1fn1N, (7.16)
f¯n+1N+cn+1(n+1)n=cndn2f¯n1N+cn1(n1)n1 (7.17)

for all n2. Multiplying (7.17) with 1N and then subtracting (7.16), we end up with

fn+1N=2n(cndn21Ndn1fnN)nn1fn1N,n2. (7.18)

Next taking the inner product in (7.16) with 1N and then subtracting (7.17) multiplied by N1/2, we finally get

cn+1=nn+1(2dn1f¯nN(2N1)dn2cn)cn1(n1)n(n+1)n1,n2. (7.19)

Taking into account that the value of f at the root o is determined by f1 via

f(o)=f0=202N+1M1f1, (7.20)

and noting that f2N and c2 are also determined by f1, we conclude that (7.18) and (7.19) define f uniquely once f1CN+1 is given.

Lemma 7.4 immediately implies that n±(H0)N+1 if vol(AN)<, where H0 is the associated minimal operator. The next result shows that it can happen that n±(H0)=N+1 upon choosing lengths n with a sufficiently fast decay.

Proposition 7.5

Let AN be the antitree as in Lemma 7.4. If (n) is decreasing and

n=O1(6N)n(n+N+3)! (7.21)

as n, then n±(H0)=N+1.

It is immediate to see that vol(AN)< if (7.21) is satisfied. Next, taking into account (7.9), observe that

m(v)=vEv|e|(n+N)n1+(n+N+2)nnn1,vSn,

as n. Suppose fH(A) and set f=f|V. Then f has the form (7.15) and hence

fn2=vSn|f(v)|2=fnN2+n|cn|2,

for all n1. This implies the following estimate

vV|f(v)|2m(v)=n0vSn|f(v)|2m(v)n1n2n1(fnN2+|cn|2). (7.22)

Next, (7.18) and (7.19) can be written as follows

fn+1Ncn+1=A1,nfnNcn+A2,nfn1Ncn1, (7.23)

where the matrices A1,n,A2,nR(N+1)×(N+1) are given explicitly by

A1,n:=2ndn1IN2ndn2BN,12ndn1n+1B1,N(2N1)ndn2n+1I1,A2,n:=nn1INn1n+1I1, (7.24)

for all n2. Since n1n and

dn1<dn2=n+N1n1+n+N+1n2(n+N)n (7.25)

for all n2, it is not difficult to get the following rough bounds

A1,n6N(n+N),A2,n=nn11, (7.26)

for all n2N. Denoting

Fn:=fnNcn,n1,

the recurrence relations (7.18) and (7.19) can be written in the following matrix form

Fn+1Fn=A1,nA2,nIN+10N+1FnFn1=AnFnFn1. (7.27)

Taking into account (7.26), we get An6N(n+N+1) for all n2N, which implies the estimate

fnN2+|cn|2=FnCk=1n1Ak(6N)n(n+N)! (7.28)

for all n2. Combining this bound with (7.21), it is easy to see that the series on the right hand side in (7.22) converges and by Lemma 2.13 we conclude that H(AN)L2(A). Thus, ker(H)=H(AN) and the use of Corollary 2.11 finishes the proof.

7.4. Infinite deficiency indices

Consider the antitree A with sphere numbers sn:=n+1, n0. Next pick a sequence of positive numbers (n) and define lengths as follows

|eijn|=2n,1i=jn+1,n,otherwise, (7.29)

for all nZ0. Thus, the corresponding matrix Mn+1 given by (7.2) has the form

Mn+1=1nBn+112In+1Bn+1,1R(n+1)×(n+2) (7.30)

for all n0. Let us denote this antitree by A.

Lemma 7.6

dim(H(A))=.

Consider the difference equations (7.5) and (7.6). Clearly, the matrix Mn+1 has the maximal rank n+1 for every n0. Taking into account that

Bn+112In+11=42n+1Bn+12In+1=:Cn,n0,

(7.6) then reads

In+122n+1Bn+1,1fn+1=nCn(DnfnMn*fn1) (7.31)

for all n1. Observe that

In+122n+1Bn+1,1f1fn+10=f1fn+1

and hence for any fnCn+1 and fn1Cn there always exists a unique fn+1=(f1,,fn+1,0) satisfying (7.31). Now pick a natural number N and define fNC(A) by setting fnN:=(0,,0)Cn+1 for all n{0,,N},

fN+1N:=(1,,1,N1/2),

and

fn+1N:=nCn(DnfnNMn*fn1N)0Cn+2 (7.32)

for all nN+1. Clearly, fN satisfies (7.5) and (7.6) and hence defines a harmonic function fNH(A). Moreover, it is easy to see that span{fN}N1 is infinite dimensional, which proves the claim.

Proposition 7.7

Let H0 be the minimal Kirchhoff Laplacian associated with the antitree A. If n is decreasing and

n=O16n(n+3)! (7.33)

as n, then n±(H0)=.

Clearly, it suffices to show that every fN constructed in the proof of Lemma 7.6 belongs to L2(G) if n decays as in (7.33). To prove this, we shall proceed as in the proof of Proposition 7.5. First, taking into account (7.29), observe that

m(v)nn1,vSn,

as n. Since fnN2=vSn|fN(v)|2 for all n0, we get the estimate

vV|fN(v)|2m(v)nN+1vSn|fN(v)|2m(v)nN+1nn1fnN2. (7.34)

Denoting Fn:=fnN for all n1, we can put (7.31) into the matrix form

Fn+1Fn=A1,nA2,nIn+10n+1,nFnFn1=AnFnFn1 (7.35)

for all nN+1, where

A1,n:=nCnDn01,n+1R(n+2)×(n+1),A2,n:=nCnMn*01,nR(n+2)×n. (7.36)

Now observe that Cn=2 and nDn2(n+1) for all n1. Moreover, nMn*n+1 for all n1, which immediately implies the following estimate

AnnCnDn2+1+nCnMn*26(n+1),nN+1. (7.37)

Hence, we get

fn+1NCk=N+1nAkC6nN(n+1)!(N+1)!6n(n+1)!

for all nN+1. Combining this estimate with (7.34) and (7.33) and using Lemma 2.13, we conclude that fNL2(A) for each N1.

Remark 7.8

It is not difficult to show that fN does not belong to H1(A) for the above choices of edge lengths. In fact, it follows from the maximum principle for H(A) that if vol(A)<, then H(A)H1(A) consists only of constant functions.

7.5. Proof of Theorem 7.1

Clearly, the case of infinite deficiency indices follows from Proposition 7.7. On the other hand, since adding and/or removing finitely many edges and vertices to a graph does not change the deficiency indices of the minimal Kirchhoff Laplacian, Proposition 7.5 completes the proof of Theorem 7.1. Indeed, every antitree AN can be obtained from A by first removing all the edges between combinatorial spheres S0 and SN and then adding N+1 edges connecting the root o with the vertices in SN.

Remark 7.9

Since every infinite antitree has exactly one end, Theorem 6.11(iv) implies that the Kirchhoff Laplacian H0 in Theorem 7.1 has a unique Markovian extension exactly when vol(A)=. If vol(A)<, then Markovian extensions of H0 form a one‐parameter family explicitly given by (6.17). Note that (6.17) looks similar to the description of self‐adjoint extensions of the minimal Kirchhoff Laplacian on radially symmetric antitrees obtained recently in [56].

Let us also emphasize that the antitree constructed in Proposition 7.7 has finite total volume and H0 has infinite deficiency indices, however, the set of Markovian extensions of H0 forms a one‐parameter family.

Let us finish this section with one more comment. As it was proved, the dimension of the space of Markovian extensions depends only on the space of graph ends and, moreover, it is equal to the number of finite volume ends. However, deficiency indices (dimension of the space of self‐adjoint extensions) are in general independent of graph ends and we can only provide a lower bound. Moreover, the above example of a polynomially growing antitree shows that the space of non‐constant harmonic functions heavily depends on the choice of edge lengths (in particular, its dimension may vary between zero and infinity). In this respect, let us also emphasize that in the case of Cayley graphs of finitely generated groups the end space is independent of the choice of a generating set, however, simple examples show that the space of harmonic functions does depend on this choice.

JOURNAL INFORMATION

The Journal of the London Mathematical Society is wholly owned and managed by the London Mathematical Society, a not‐for‐profit Charity registered with the UK Charity Commission. All surplus income from its publishing programme is used to support mathematicians and mathematics research in the form of research grants, conference grants, prizes, initiatives for early career researchers and the promotion of mathematics.

ACKNOWLEDGEMENTS

We thank Matthias Keller, Daniel Lenz, Primož Moravec, Andrea Posilicano, and Wolfgang Woess for useful discussions and hints with respect to the literature. We also thank the referees for their comments which have helped us to improve the manuscript. N. Nicolussi appreciates the hospitality at the Institute of Mathematics, University of Potsdam, during a research stay funded by the OeAD (Marietta Blau‐grant, ICM‐2019‐13386), where a part of this work was done. Research was supported by the Austrian Science Fund (FWF) under Grant Numbers: P 28807 (A. Kostenko and N. Nicolussi) and W 1245 (N. Nicolussi), by the German Research Foundation (DFG) under Grant Number: 397230547 (D. Mugnolo), and by the Slovenian Research Agency (ARRS) under Grant Number: J1‐1690 (A. Kostenko). The authors would like to acknowledge that this article is based upon work from COST Action CA18232 MAT‐DYN‐NET, supported by COST (European Cooperation in Science and Technology).

APPENDIX A. LINEAR RELATIONS IN HILBERT SPACES

In this section, we collect basic notions and facts on linear relations in Hilbert spaces, a very convenient concept of multi‐valued linear operators. For simplicity, we shall assume that H is a finite‐dimensional Hilbert space, N:=dim(H)<.

A linear relation Θ in H is a linear subspace in H×H. Linear operators become special linear relations (single valued) after identifying them with their graphs in H×H. Consider linear relations in H having the form

ΘC,D={(f,g)H×H|Cf=Dg}, (A.1)

where C,D are linear operators on H. Note that different C and D may define the same linear relation. The domain and the multi‐valued part of ΘC,D are given by

dom(ΘC,D)={fH|gH,Cf=Dg}={fH|Cfran(D)},mul(ΘC,D)={gH|Dg=0}=ker(D).

In particular, ΘC,D is a graph of a linear operator only if ker(D)={0}.

The adjoint relation ΘC,D* to ΘC,D is given by

ΘC,D*={(f,g)H×H|g,fH=f,gH(f,g)ΘC,D}={(D*f,C*f)|fH}. (A.2)

Thus, a linear relation ΘC,D is self‐adjoint, ΘC,D=ΘC,D*, if and only if C and D satisfy the Rofe–Beketov conditions [68] (see also [69, Exercises 14.9.3‐4]):

CD*=DC*,0ρ(C*C+D*D). (A.3)

Taking into account that every linear relation in H admits one of the forms (A.1) or (A.2), this provides a description of self‐adjoint linear relations in H. Note also that the second condition in (A.3) is equivalent to the fact that the matrix (C|D)CN×2N has the maximal rank N.

Recall also that every self‐adjoint linear relation admits the representation Θ=ΘopΘmul, where Θmul:={0}×mul(Θ) and Θop, called the operator part of Θ, is a graph of a linear operator. In particular, for a self‐adjoint linear relation ΘC,D one has

dom(ΘC,D)=mul(ΘC,D)=ker(D)=ran(D*). (A.4)

For further details on linear relations, we refer the reader to, for example, [69, Chapter 14.1].

APPENDIX B. A ROPE LADDER GRAPH

Let us introduce a rope ladder graph depicted on Figure B.1. Let Gd=(V,E) be a simple graph with the vertex set V:={o}V+V, where o=v0 is a root, V+=(vn+)n1 and V=(vn)n1 are two disjoint countably infinite sets of vertices. The edge set E is defined as follows:

  • o is connected to v1+ and v1 by the ‘diagonal’ edges e0+ and e0, respectively;

  • for each n1, vn± is connected to vn+1± by the vertical edge en±;

  • for each n1, vn+ and vn are connected by the horizontal edge en.

FIGURE B.1.

FIGURE B.1

The rope ladder graph

By construction, deg(o)=2 and deg(vn+)=deg(vn)=3 for all n1. Moreover, an infinite rope ladder graph has exactly one end. Note also that a similar example was studied in [46, Section 7] (see also [33, Section 5]) in context with the construction of non‐constant harmonic functions of finite energy.

Equip now Gd with edge lengths |·|:E(0,) and consider the corresponding minimal Kirchhoff Laplacian H0 on the metric graph G=(Gd,|·|). The next result immediately follows from Theorem 2.8 and Corollary 2.11.

Corollary B.1

If

n1|en+|+|en|=,andn1|en|+|en|=, (B.1)

then the Kirchhoff Laplacian H0 is self‐adjoint. If

vol(G)=n1|en+|+|en|+|en|<, (B.2)

then n±(H0)1.

We omit the proof since it is easy to check that the first condition is equivalent to the geodesic completeness of (V,ϱm) (cf. Theorem 2.8). Due to the symmetry of the underlying combinatorial graph, the gap between the above two conditions is equivalent to the fact that the corresponding lengths satisfy

n1|en+|=,n1|en|+|en|<. (B.3)

Next, let us describe the space of harmonic functions H(G).

Lemma B.2

Let a,bC. Then there is exactly one fH(G) such that

f(v1+)=a,f(v1)=b. (B.4)

Moreover, this function f is recursively given by

f(o)=b|e0+|+a|e0||e0+|+|e0| (B.5)

and

f(vn+1±)=1+|en±||en1±|+|en±||en|f(vn±)|en±||en1±|f(vn1±)|en±||en|f(vn) (B.6)

for all nZ1, where we use the notation v0+:=v0:=o.

Suppose a,bC are given and fH(G) satisfies (B.4). Since f is linear on every edge and satisfies (2.7) at v=o, we get

0=fe0+(o)+fe0(o)=f(v1+)f(o)|e0+|+f(v1)f(o)|e0|=af(o)|e0+|+bf(o)|e0|,

which implies (B.5). Moreover, Kirchhoff conditions (2.7) at v=vn±, n1 read

f(vn+1±)f(vn±)|en±|+f(vn1±)f(vn±)|en1±|+f(vn)f(vn±)|en|=0.

This implies that f is given by (B.6). Hence there is at most one fH(G) satisfying (B.4) for given a,bC. However, the same calculation shows that f defined by (B.5) and (B.6) has this property. Thus, existence follows as well.

From Lemma B.2, it is clear that dim(H(G))=2, and, moreover,

H(G)=span{1G,g0},

where 1G denotes the constant function on G and g0H(G) is the function defined, for example, by the following normalization

g0(0)=0,g0(v1+)=|e0+|,g0(v1)=|e0|. (B.7)

Note that g0(vn±), n1 are then given recursively by (B.6).

Lemma B.3

If vol(G)<, then

H(G)H1(G)=span{1G}. (B.8)

The claim immediately follows from the fact that a rope ladder graph has exactly one end. However, let us present a direct proof based on the analysis of harmonic functions.

Taking into account (B.8), we only need to show that g0H1(G). First, observe that (g0(vn+))n1 and (g0(vn))n1 are strictly increasing positive, respectively, strictly decreasing negative sequences. Indeed,

|e0|=g0(v1)<0=g0(o)<g0(v1+)=|e0+|

by the very definition of g0. Let n1 and assume now that we have already shown that (g0(vk+))k=1n is strictly increasing and (g0(vk))k=1n is strictly decreasing. Since g0(o)=0, (B.6) implies

g0(vn+1+)=1+|en+||en1+|+|en+||en|g0(vn+)|en+||en1+|g0(vn1+)|en+||en|g0(vn)>1+|en+||en|g0(vn+)+|en+||en1+|(g0(vn+)g0(vn1+))>g0(vn+).

A similar argument shows that g0(vn+1)<g0(vn) and hence the claim follows by induction. Now monotonicity immediately implies

g0L2(G)2=eEe|g0(xe)|2dxen0en|g0(xe)|2dxe=n=0|g0(vn+)g0(vn)|2|en||g0(v1+)g0(v1)|2n=01|en|=,

since vol(G)<. Thus, g0H1(G).

In particular, this also leads to the following result:

Corollary B.4

If vol(G)<, then n±(H0){1,2}. Moreover, n±(H0)=1 if and only if g0L2(G).

The claim about the deficiency indices follows from Corollary 2.11 and the fact that 1GL2(G). The equivalences then follow from Lemma B.3.

As the next example shows, the inclusion g0L2(G) heavily depends on the choice of edge lengths.

Example B.5

Fix s>3 and equip the rope ladder graph with edge lengths

|en+|=|en|:=1(n+1)s,|en|:=2n(n+1)sns,nZ0.

Then |en|n2s for large n and hence vol(G)<. Moreover, for this particular choice of edge lengths we have g0(vn±)=±n for all n1. Indeed, g0(v1±)=±1 by (B.7). Assuming we have already proven that g0(vk±)=±k for kn with some n1, we have by (B.6):

g0(vn+1+)=1+ns(n+1)s+1(n+1)s|en|nns(n1)(n+1)s+n(n+1)s|en|=n+ns(n+1)s+2n(n+1)s|en|=n+ns(n+1)s+(n+1)sns(n+1)s=n+1.

Analogously, g0(vn+1)=(n+1) and hence the claim follows by induction.

Applying Lemma B.3 and using again that |en|n2s as n, we conclude that g0L2(G) exactly (see Lemma 2.13) when

n1|g0(vn±)|2(|en1±|+|en±|)=n1n2((n+1)s+ns)<

and

n1|g0(vn±)|2|en1|=n12n3(n+1)sns<.

Clearly, the latter holds only if s>5. Hence, by Lemma B.4, n±(H0)=2 for all s>5. In particular, ker(H)H1(G)s5.

Present address Aleksey Kostenko, Faculty of Mathematics and Physics, University of Ljubljana, Jadranska ul. 19, Ljubljana, Slovenia, and Faculty of Mathematics, University of Vienna, Oskar‐Morgenstern‐Platz 1, Vienna, Austria

Footnotes

Equivalently, R1R2 if and only if R1 and R2 cannot be separated by a finite vertex set, that is, for every finite subset XV the remaining tails of R1 and R2 in VX belong to the same connected component of VX.

Note that for a subgraph G of G its boundary is G={vV(G)|degG(v)<degG(v)} and hence G is compact only if #G<.

For an operator T with dense domain in a Hilbert space H, λC is called a point of regular type of T if there exists c=cλ>0 such that (Tλ)fcf for all fdom(T).

A classification of groups having infinitely many ends is given in Stallings's ends theorem [73] (see also [32, Theorem 13.5.10] and Remark 2.5(iv)).

We shall write AB for two non‐negative self‐adjoint operators A and B if their quadratic forms tA and tB satisfy dom(tB)dom(tA) and tA[f]tB[f] for every fdom(tB).

Here we do not assume that t^ is densely defined, see [31, p. 29]. We stress that in order for t^ to be a Dirichlet form even merely in the wide sense, it is necessary that dom(t^) is a sublattice of H, hence that the orthogonal projector onto ran(D*) is a positivity preserving operator.

A normal contraction is a function φ:CC such that φ(0)=0 and |φ(x)φ(y)||xy| for all x,yC.

For instance, for any s,L>0 such that sL, the function ψ0(x):=L+s2|xL+s2| satisfies ψ0(0)=0, ψ0(L)=s and |ψ0|1. The construction in the general case follows easily from this example.

Here and below to estimate norms, we use the equality A=A*A and the following simple estimate for non‐negative 2×2 block‐matrices A=A11A12A12*A22: AA11+A22. There are other estimates (for example, [36, inequality (2.3.8)]), however, they do not seem to work as good as the above approach.

REFERENCES

  • 1. Akhiezer N. I. and Glazman I. M., Theory of linear operators in Hilbert spaces, vol. II, Dover Publications, New York, 1993. [Google Scholar]
  • 2. Amini O. and Caporaso L., Riemann–Roch theory for weighted graphs and tropical curves, Adv. Math. 240 (2013), 1–23. [Google Scholar]
  • 3. Arendt W. and ter Elst A. F. M., Operators with continuous kernels, Integr. Equ. Oper. Theory 91 (2019), no. 5, Art. 45. [Google Scholar]
  • 4. Arendt W. and Bukhvalov A. V., Integral representations of resolvents and semigroups, Forum. Math. 6 (1994), 111–135. [Google Scholar]
  • 5. Baker M. and Norine S., Riemann–Roch and Abel–Jacobi theory on a finite graph, Adv. Math. 215 (2007), no. 2, 766–788. [Google Scholar]
  • 6. Baker M. and Rumely R., Potential theory and dynamics on the Berkovich projective line, Mathematical Surveys and Monographs, vol. 159, Amer. Math. Soc., Providence, RI, 2010. [Google Scholar]
  • 7. Barlow M. T. and Bass R. F., Stability of parabolic Harnack inequalities, Trans. Amer. Math. Soc. 356 (2004), no. 4, 1501–1533. [Google Scholar]
  • 8. Berkolaiko G., Carlson R., Fulling S., and Kuchment P., Quantum graphs and their applications, Contemporary Mathematics, vol. 415, Amer. Math. Soc., Providence, RI, 2006. [Google Scholar]
  • 9. Berkolaiko G. and Kuchment P., Introduction to quantum graphs, Amer. Math. Soc., Providence, RI, 2013. [Google Scholar]
  • 10. Breuer J. and Keller M., Spectral analysis of certain spherically homogeneous graphs, Oper. Matrices 7 (2013), no. 4, 825–847. [Google Scholar]
  • 11. Breuer J. and Levi N., On the decomposition of the Laplacian on metric graphs, Ann. Henri Poincaré 22 (2020), no. 2, 499–537. [Google Scholar]
  • 12. Brezis H., Functional analysis, Sobolev spaces and partial differential equations, Universitext, Springer, New York, 2011. [Google Scholar]
  • 13. Burago D., Burago Yu., and Ivanov S., A course in metric geometry, Graduate Studies in Mathematics, vol. 33, Amer. Math. Soc., Providence, RI, 2001. [Google Scholar]
  • 14. Cardanobile S. and Mugnolo D., Parabolic systems with coupled boundary conditions, J. Differential Equations 247 (2009), 1229–1248. [Google Scholar]
  • 15. Carlson R., Nonclassical Sturm–Liouville problems and Schrödinger operators on radial trees, Electr. J. Differential Equations 2000 (2000), no. 71, 1–24. [Google Scholar]
  • 16. Carlson R., Boundary value problems for infinite metric graphs , in [26].
  • 17. Carlson R., Dirichlet to Neumann maps for infinite quantum graphs, Netw. Heterog. Media 7 (2012), no. 3, 483–501. [Google Scholar]
  • 18. Cartwright D. I., Soardi P. M., and Woess W., Martin and end compactifications for nonlocally finite graphs, Trans. Amer. Math. Soc. 338, no. 2, 679–693 (1993). [Google Scholar]
  • 19. Colin de Verdière Y., Torki‐Hamza N., and Truc F., Essential self‐adjointness for combinatorial Schrödinger operators II–metrically non complete graphs, Math. Phys. Anal. Geom. 14 (2011), no. 1, 21–38. [Google Scholar]
  • 20. Coulhon T., Off‐diagonal heat kernel lower bounds without Poincaré, J. Lond. Math. Soc. (2) 68 (2003), 795–816. [Google Scholar]
  • 21. Cushing D., Liu S., Münch F., and Peyerimhoff N., Curvature calculations for antitrees, in Keller M. et al. (eds.), Analysis and geometry on graphs and manifolds, London Mathematical Society Lecture Notes Series, vol. 461, Cambridge University Press, Cambridge, 2020. [Google Scholar]
  • 22. Damanik D., Fang L., and Sukhtaiev S., Zero measure and singular continuous spectra for quantum graphs, Ann. Henri Poincaré 21 (2020), no. 7, 2167–2191. [Google Scholar]
  • 23. Diestel R., Graph theory, 5th ed., Graduate Texts in Mathematics, vol. 173, Springer, Heidelberg, 2017. [Google Scholar]
  • 24. Diestel R. and Kühn D., Graph‐theoretical versus topological ends of graphs, J. Combin. Theory, Ser. B 87 (2003), 197–206. [Google Scholar]
  • 25. Dodziuk J. and Karp L., Spectral and function theory for combinatorial Laplacians, Geometry of random motion (Ithaca, NY, 1987), Contemporary Mathematics, vol. 73, Amer. Math. Soc., Providence, RI, 1988, pp. 25–40. [Google Scholar]
  • 26. Exner P., Keating J. P., Kuchment P., Sunada T., and Teplyaev A., Analysis on graphs and its applications, Proceedings of Symposia in Pure Mathematics, vol. 77, Amer. Math. Soc., Providence, RI, 2008. [Google Scholar]
  • 27. Exner P., Kostenko A., Malamud M., and Neidhardt H., Spectral theory of infinite quantum graphs, Ann. Henri Poincaré 19 (2018), no. 11, 3457–3510. [Google Scholar]
  • 28. Folz M., Volume growth and stochastic completeness of graphs, Trans. Amer. Math. Soc. 366 (2014), 2089–2119. [Google Scholar]
  • 29. Freudenthal H., Über die Enden topologischer Räume und Gruppen, Math. Z. 33 (1931), 692–713. [Google Scholar]
  • 30. Freudenthal H., Über die Enden diskreter Räume und Gruppen, Comment. Math. Helv. 17 (1944), 1–38. [Google Scholar]
  • 31. Fukushima M., Oshima Y., and Takeda M., Dirichlet forms and symmetric Markov processes, 2nd ed., De Gruyter, Berlin, 2010. [Google Scholar]
  • 32. Geoghegan R., Topological methods in group theory, Graduate Texts in Mathematics, vol. 243, Springer, Berlin, 2008. [Google Scholar]
  • 33. Georgakopoulos A., Uniqueness of electrical currents in a network of finite total resistance, J. Lond. Math. Soc. 82 (2010), 256–272. [Google Scholar]
  • 34. Georgakopoulos A., Graph topologies induced by edge lengths, Discrete Math. 311 (2011), no. 15, 1523–1542. [Google Scholar]
  • 35. Georgakopoulos A., Haeseler S., Keller M., Lenz D., and Wojciechowski R., Graphs of finite measure, J. Math. Pures Appl. 103 (2015), S1093–S1131. [Google Scholar]
  • 36. Golub G. and Van Loan C. F., Matrix computations, 4th ed., The Johns Hopkins University Press, Baltimore, 2012. [Google Scholar]
  • 37. Grigor'yan A. and Masamune J., Parabolicity and stochastic completeness of manifolds in terms of the Green formula, J. Math. Pures Appl. 100 (2013), 607–632. [Google Scholar]
  • 38. Halin R., Über unendliche Wege in Graphen, Math. Ann. 157 (1964), 125–137. [Google Scholar]
  • 39. Haeseler S., Keller M., Lenz D., and Wojciechowski R., Laplacians on infinite graphs: Dirichlet and Neumann boundary conditions, J. Spectr. Theory 2 (2012), no. 4, 397–432. [Google Scholar]
  • 40. Haeseler S., Analysis of Dirichlet forms on graphs , Ph.D. thesis, Jena, 2014; arXiv:1705.06322, 2017.
  • 41. Hinz M. and Schwarz M., A note on Neumann problems on graphs , arXiv:1803.08559, 2018.
  • 42. Hopf H., Enden offener Räume und unendliche diskontinuierliche Gruppen, Comment. Math. Helv. 16 (1943), 81–100. [Google Scholar]
  • 43. Hua B., Liouville theorem for bounded harmonic functions on manifolds and graphs satisfying non‐negative curvature dimension condition, Calc. Var. Partial Differential Equations 58 (2019), no. 2, Art. 42. [Google Scholar]
  • 44. Hua B. and Keller M., Harmonic functions of general graph Laplacians, Calc. Var. Partial Differential Equations 51 (2014), 343–362. [Google Scholar]
  • 45. Huang X., Keller M., Masamune J., and Wojciechowski R., A note on self‐adjoint extensions of the Laplacian on weighted graphs, J. Funct. Anal. 265 (2013), 1556–1578. [Google Scholar]
  • 46. Jorgensen P. and Pearse E., Gel'fand triples and boundaries of infinite networks, New York J. Math. 17 (2011), 745–781. [Google Scholar]
  • 47. Kant U., Klauß T., Voigt J., and Weber M., Dirichlet forms for singular one‐dimensional operators and on graphs, J. Evol. Equ. 9 (2009), 637–659. [Google Scholar]
  • 48. Kasue A., Convergence of metric graphs and energy forms, Rev. Mat. Iberoam.  26 (2010), no. 2, 367–448. [Google Scholar]
  • 49. Kasue A., Convergence of Dirichlet forms induced on boundaries of transient networks, Potential Anal. 47 (2017), no. 2, 189–233. [Google Scholar]
  • 50. Kato T., Perturbation theory for linear operators, 2nd ed., Springer, Berlin, 1976. [Google Scholar]
  • 51. Keller M. and Lenz D., Dirichlet forms and stochastic completeness of graphs and subgraphs, J. Reine Angew. Math. 666 (2012), 189–223. [Google Scholar]
  • 52. Keller M., Lenz D., Schmidt M., and Schwarz M., Boundary representation of Dirichlet forms on discrete spaces, J. Math. Pures Appl. 126 (2019), 109–143. [Google Scholar]
  • 53. Keller M., Lenz D., Schmidt M., and Wojciechowski R. K., Note on uniformly transient graphs, Rev. Mat. Iberoam. 33 (2017), no. 3, 831–860. [Google Scholar]
  • 54. Keller M., Lenz D., and Wojciechowski R. K., Volume growth, spectrum and stochastic completeness of infinite graphs, Math. Z. 274 (2013), no. 3–4, 905–932. [Google Scholar]
  • 55. Kostenko A. and Nicolussi N., Spectral estimates for infinite quantum graphs, Calc. Var. Partial Differential Equations 58 (2019), no. 1, Art. 15. [Google Scholar]
  • 56. Kostenko A. and Nicolussi N., Quantum graphs on radially symmetric antitrees, J. Spectral Theory 11 (2021), no. 2, 411–460. [Google Scholar]
  • 57. Kostrykin V., Potthoff J., and Schrader R., Contraction semigroups on metric graphs , in [26].
  • 58. Lupu T., From loop clusters and random interlacements to the free field, Ann. Probab. 44 (2016), 2117–2146. [Google Scholar]
  • 59. Lupu T., Convergence of the two‐dimensional random walk loop‐soup clusters to CLE, J. Eur. Math. Soc. 21 (2019), 1201–1227. [Google Scholar]
  • 60. Marti J. T., Evaluation of the least constant in Sobolev's inequality for H1(0,s) , SIAM J. Numer. Anal. 20 (1983), 1239–1242. [Google Scholar]
  • 61. Masamune J., Essential self‐adjointness of Laplacians on Riemannian manifolds with fractal boundary, Comm. Partial Differential Equations 24, no. 3–4, 749–757 (1999). [Google Scholar]
  • 62. Masamune J., Analysis of the Laplacian of an incomplete manifold with almost polar boundary, Rend. Mat. Appl. (7) 25 (2005), no. 1, 109–126. [Google Scholar]
  • 63. Mugnolo D. and Nittka R., Properties of representations of operators acting between spaces of vector‐valued functions, Positivity 15 (2011), 135–154. [Google Scholar]
  • 64. Murakami A. and Yamasaki M., An introduction of Kuramochi boundary of an infinite network, Mem. Fac. Sci. Eng. Shimane Univ. Ser. B Math. Sci. 30 (1997), 57–89. [Google Scholar]
  • 65. Nicolussi N., Strong isoperimetric inequality for tessellating quantum graphs, in Fatihcan F. et al. (eds.), Discrete and continuous models in the theory of networks, Operator Theory: Advances and Applications, vol. 64, Birkhäuser, Basel, 2020, pp. 271–290. [Google Scholar]
  • 66. Ouhabaz E. M., Analysis of heat equations on domains, Princeton University Press, Princeton, NJ, 2005. [Google Scholar]
  • 67. Post O., Spectral Analysis on graph‐like spaces, Lecture Notes in Mathematics, vol. 2039, Springer, Berlin, 2012. [Google Scholar]
  • 68. Rofe‐Beketov F. S., Self‐adjoint extensions of differential operators in a space of vector‐valued functions, Teor. Funkcii, Funkcional. Anal. i Priložen. 8 (1969), 3–24 (in Russian). [Google Scholar]
  • 69. Schmüdgen K., Unbounded self‐adjoint operators on Hilbert space, Graduate Texts in Mathematics, vol. 265, Springer, Berlin, 2012. [Google Scholar]
  • 70. Shokrieh F. and Wu C., Canonical measures on metric graphs and a Kazhdan's theorem, Invent. Math. 215 (2019), 819–862. [Google Scholar]
  • 71. Soardi P. M., Potential theory on infinite networks, Lecture Notes in Mathematics, vol. 1590, Springer, Berlin, 1994. [Google Scholar]
  • 72. Solomyak M., On the spectrum of the Laplacian on regular metric trees, Waves Random Media 14 (2004), S155–S171. [Google Scholar]
  • 73. Stallings J. R., Group theory and three‐dimensional manifolds, Yale University Press, New Haven, CT, 1971. [Google Scholar]
  • 74. Sturm K.‐T., Analysis on local Dirichlet spaces I. Recurrence, conservativeness and Lp‐Liouville properties, J. Reine Angew. Math. 456 (1994), 173–196. [Google Scholar]
  • 75. Wall C. T. C., Poincaré complexes: I, Ann. of Math. 86 (1967), no. 2, 213–245. [Google Scholar]
  • 76. Woess W., Random walks on infinite graphs and groups, Cambridge University Press, Cambridge, 2000. [Google Scholar]
  • 77. Wojciechowski R. K., Stochastically incomplete manifolds and graphs, in Lenz D. et al. (eds.), Random walks, boundaries and spectra, Progress in Probability, vol. 64, Birkhäuser/Springer, Basel, 2011, pp. 163–179. [Google Scholar]

Articles from Journal of the London Mathematical Society are provided here courtesy of Wiley

RESOURCES