This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Spaces of random plane triangulations and the density of states

Nathan Hannon

1. Introduction

An important tool for studying tilings is the tiling space. [13] The set of all tilings of n\mathbb{R}^{n} forms a topological space under a metric defined in such a way that two tilings are close if and only if, after a small translation, they agree on a large ball around the origin. The tiling space of a tiling TT, also called the hull of TT, is the closure in this space of the set of all translates of TT. Another, perhaps more intuitive, characterization of the hull of TT is the set of all tilings TT^{\prime} such that every finite patch of TT^{\prime} can be found, up to translation, in TT. There are variations on this construction that also take rotations into account. The spectral properties, particularly the integrated density of states (IDS), of operators on tiling spaces are of interest as they are related to physical properties of solids modeled by those tilings. These properties have been determined in several cases, such as by Julien and Savinien [8].

In a more general context in which a metric space is equipped with a group action (which includes tiling spaces with the action of translation), Lenz and Veselić [11] determined that the IDS of a class of operators can be approximated uniformly by analogues constructed on finite sets, and that jumps of the IDS correspond to compactly supported eigenfunctions of those operators. In a different but related setting, Beckus and Pogorzelski [3] proved that the density of states of a random operator on a Delone dynamical system is continuous with respect to the system (under appropriately defined topologies).

Other results have dealt with groupoid structures. For example, Lenz et al. [10] constructed a von Neumann algebra, trace, and density of states in a setting involving random operators on a groupoid. Additionally, Beckus et al. [2] proved that the spectra of certain operators on subsets of a groupoid are continuous, with respect to suitable topologies, as a function of the subset. Gap-labeling conjectures for some cases have been proven by Benameur and Mathai [4, 5] and by Kaminker and Putnam [9].

We want to generalize the tiling space construction to tiling-like structures that do not live in n\mathbb{R}^{n} - in this case, random triangulations of 2-manifolds, although the construction could easily work with any sufficiently well-behaved cell complex. Although most of the triangulations that we will consider are homeomorphic to 2\mathbb{R}^{2}, they have no notion of translation or any useful group action, but can be given a groupoid structure. Our goal is to prove results analogous to Lenz and Veselić [11] for these spaces.

Triangulation spaces can be given a discrete or continuous structure. The continuous space is a foliated space as constructed by Moore and Schochet [12], and the discrete space is a transversal of that space. Although our work uses both of these structures, our results will focus primarily on the discrete space, since it is generally easier to work with and has the same large-scale properties.

Because we are modeling random triangulations, our results involve measures, which can be approximated by sequences of measures on spheres or on a single leaf. In particular, approximating via spheres gives a way to approximate the IDS via computing eigenvalues on finite spaces.

Fabila Carrasco et al. [7] studied a discrete magnetic Laplacian on graphs, and we are interested in studying similar operators in our setting:

Δdiscf(x1,z1)=1w(x1)y1N(x1)w¯(x1,y1)(f(y1,z1)f(x1,z1))\Delta_{\mathrm{disc}}f(x_{1},z_{1})=\frac{1}{w(x_{1})}\sum_{y_{1}\in N(x_{1})}\bar{w}(x_{1},y_{1})\left(f(y_{1},z_{1})-f(x_{1},z_{1})\right)

and

Δdisc,V,Bf(x,z)=1w(x1)y1N(x1)w¯(x1,y1)(eiα(x1,y1)f(y1,z1))+(V(x1)1)f(x1,z1),\Delta_{\mathrm{disc},V,B}f(x,z)=\frac{1}{w(x_{1})}\sum_{y_{1}\in N(x_{1})}\bar{w}(x_{1},y_{1})\left(e^{i\alpha(x_{1},y_{1})}f(y_{1},z_{1})\right)+(V(x_{1})-1)f(x_{1},z_{1}),

where ww and w¯\bar{w} are weight functions on edges and vertices, respectively, and α\alpha and VV are data related to the magnetic field and potential. Although our results are formulated with these operators in mind, they apply to a large class of operators.

Our major results are as follows. In these results, G^\hat{G} is a groupoid consisting of twice-marked triangulations, ν\nu is a transverse measure on G^\hat{G}, and W(G^)W^{*}(\hat{G}) is a von Neumann algebra that we will define on G^\hat{G} using the measure ν\nu.

Theorem 8. Suppose that ν\nu is such that, for a.e. xx, GxG^{x} is recurrent and νx=ν\nu^{x}=\nu. If ν(x)>0\nu(x)>0 for some (equivalently, a.e.) xx, then W(G^)W^{*}(\hat{G}) is a hyperfinite type I|G^0|I_{|\hat{G}_{0}|} factor. Otherwise, W(G^)W^{*}(\hat{G}) is a hyperfinite type II1II_{1} factor.

Theorem 11. If HH has finite hopping range, and νn\nu^{\prime}_{n} is a sequence of measures converging to ν\nu, then we can obtain the density of states κH(ϕ)\kappa_{H}(\phi) as a limit of the leafwise local trace Tr(ϕ(H))\mathrm{Tr}(\phi(H)) averaged with respect to νn\nu^{\prime}_{n}; that is,

κH(ϕ)=limnTr(νn(G^A)ϕ(H)).\kappa_{H}(\phi)=\lim_{n\to\infty}\mathrm{Tr}\left(\nu^{\prime}_{n}(\hat{G}^{A})\phi(H)\right).

Theorem 12. Suppose that 𝒟\mathcal{D} is discrete, G^\hat{G} is ergodic, and that HW(G^)H\in W^{*}(\hat{G}) is GG-invariant and has finite hopping range. Suppose also that tt\in\mathbb{R}. The following are equivalent:

  1. (1)

    The density of states κH(t)>0\kappa_{H}(t)>0.

  2. (2)

    For some xx, H|G^xH|_{\hat{G}^{x}} has an eigenfunction with eigenvalue tt supported on some finite patch AA with ν(GA)>0\nu(G^{A})>0.

  3. (3)

    For almost all xx, ker(HG^xλI)\ker(H_{\hat{G}^{x}}-\lambda I) is nontrivial and spanned by compactly supported eigenfunctions.

2. Notation

ArA_{r}:

the rrth interior of the discrete decorated finite patch AA

Cr(G),Cr(G^C^{*}_{r}(G),C^{*}_{r}(\hat{G}):

the reduced CC^{*}-algebra associated with the groupoid GG or G^\hat{G}

dDd_{D}:

the distance on the decoration space

DD:

the diagonal function on G^\hat{G}

𝒟\mathcal{D}:

the decoration space

gA,Tg_{A,T}:

the inclusion map of a decorated (discrete or continuous) finite patch AA into the triangulation TT

GG:

the holonomy groupoid: the set of triangulations with two marked tangent vectors, up to automorphisms

G0G_{0}:

the set of units in the holonomy groupoid: the set of triangulations with a marked tangent vector, up to automorphisms

G^,G^0\hat{G},\hat{G}_{0}:

the discrete holonomy groupoid, and its set of units

G0A,GA,G^0A,G^AG_{0}^{A},G^{A},\hat{G}_{0}^{A},\hat{G}^{A}:

the set of (discrete, continuous) triangulations containing the (marked, twice-marked) finite patch AA

N(x)N(x):

the set of neighbors of a vertex xx

S(T)S(T):

the discrete unit tangent bundle of TT

S(|T|)S(|T|):

the unit tangent bundle of |T||T|

|T||T|:

the smooth geometric realization of TT

𝒯\mathcal{T}:

the set of abstract simplicial complexes

W(G),W(G^W^{*}(G),W^{*}(\hat{G}):

the von Neumann algebra associated with the groupoid GG or G^\hat{G}

xAx_{A}:

the marked tangent vector on the (discrete or continuous) finite patch AA

x1,x2x_{1},x_{2}:

the tail and head of the discrete tangent vector xx

δ\delta:

the metric defined on G0G_{0}

κH\kappa_{H}:

the density of states associated with the operator HW(G^)H\in W^{*}(\hat{G})

ν\nu:

the transverse measure on G0G_{0}

τ\tau:

the trace on W(G)W^{*}(G) or W(G^)W^{*}(\hat{G})

ψA,B\psi_{A,B}:

the coordinate patch associated with the patch AA and ball BB

ΩA\Omega_{A}:

the set of permissible decorations on the decorated finite patch AA

3. Triangulation spaces

3.1. The space of pointed triangulations

Definition 1 (Triangulation; geometric realization).

By a triangulation we mean an abstract simplicial complex TT^{\prime} of dimension 2. The sets of vertices, edges, directed edges, and faces of TT^{\prime} will be denoted V(T)V(T^{\prime}), E(T)E(T^{\prime}) E(T)\overrightarrow{E}(T^{\prime}), and F(T)F(T^{\prime}). The geometric realization of |T||T^{\prime}| of TT^{\prime} is a metric space formed by assigning to each vertex vv, edge ee, and face ff a Euclidean simplex |v|,|e|,|f||v|,|e|,|f| of the same dimension with marked points corresponding to vertices, respecting inclusion so that |v||v| is identified with the corresponding point in |e||e| if vv is a vertex of ee, and |e||e| is identified with the corresponding segment in |f||f| if ee is an edge of ff. The usual metric on |T||T^{\prime}| is simply the Euclidean metric on each simplex, joined by shortest paths.

Fix an integer d>6d>6 and a compact metric space 𝒟\mathcal{D}. Let dDd_{D} denote the metric on 𝒟\mathcal{D}.

Definition 2 (Decorated triangulation).

Let 𝒯\mathcal{T} be the set of all ordered pairs T=(T,DT)T=(T^{\prime},D_{T}), where TT^{\prime} is a triangulation satisfying:

  • each vertex of TT^{\prime} has degree at most dd, and

  • the geometric realization |T||T^{\prime}| of TT^{\prime} is a surface of genus 0,

and where DT:E(T)𝒟D_{T}:\overrightarrow{E}(T^{\prime})\to\mathcal{D}. That is, a decorated triangulation TT consists of a triangulation TT^{\prime} with decorations DD. (These decorations are analogous to what in the study of tilings are called markings; we call them decorations because we have a different use in mind for the term “markings”.) Any simplicial properties of TT^{\prime} will be regarded as properties of TT; e.g., we will define V(T)=V(T)V(T)=V(T^{\prime}). We require isomorphisms between triangulations to preserve decorations: TT is isomorphic to UU if and only if there is a simplicial isomorphism ϕ:TU\phi:T^{\prime}\to U^{\prime} such that DU(ϕ(e))=DT(e)D_{U}(\phi(e))=D_{T}(e) for all eE(T1)e\in\overrightarrow{E}(T_{1}). Although we have defined decorations on edges, we could also speak of decorations on vertices, for example by considering triangulations where some components of DD are required to be equal for all edges emanating from a vertex.

Next we will define a geometric structure on such a triangulation. There are many possible ways to define such a structure, including some that take decorations into account. For our purposes, the following structure will suffice.

Definition 3 (The smooth geometric realization of a triangulation).

Let T𝒯T\in\mathcal{T}, and let |T||T| denote its geometric realization. Denote by dvd_{v} the degree of vv for each vertex vv of TT. For each directed edge (v,w)(v,w) of TT, map B(0,2/3)2B(0,2/3)\subset\mathbb{R}^{2} to B(v,2/3)|T|B(v,2/3)\subset|T| by preserving the distance from 0 or vv, and subdividing the unit circle into dvd_{v} equal intervals and mapping the kkth interval linearly to the angles on the kkth face counterclockwise from ww. These maps form charts of |T||T|. Let ,v\langle\cdot,\cdot\rangle_{v} be the Euclidean inner product on B(0,2/3)B(0,2/3) in the aforementioned chart, and let ,pl\langle\cdot,\cdot\rangle_{pl} be the Euclidean inner product on the usual piecewise linear structure of |T||T|. We observe that ,v\langle\cdot,\cdot\rangle_{v} does not depend on the choice of the second vertex ww, and that ,pl\langle\cdot,\cdot\rangle_{pl} is defined on all points except for vertices. Hence we can define a Riemannian metric

,=j(r),v+(1j(r)),pl,\langle\cdot,\cdot\rangle=j(r)\langle\cdot,\cdot\rangle_{v}+(1-j(r))\langle\cdot,\cdot\rangle_{pl},

where jj is a smooth function with j(0)=1j(0)=1 and j(r)=0j(r)=0 for r13r\geq\frac{1}{3}, and rr is the distance from the point xx to the nearest vertex.

Let S(|T|)S(|T|) denote the unit tangent bundle of |T||T|, with the local product metric given by its structure as an S1S^{1}-fiber bundle.

We note that the isometry class of a face |F||F| depends only on the degrees of the vertices of FF.

For a triangulation T𝒯T\in\mathcal{T}, we will denote by Aut(T)\mathrm{Aut}(T) the group of simplicial automorphisms of TT (which is trivial in most cases). The group Aut(T)\mathrm{Aut}(T) acts naturally on TT, |T||T|, and S(|T|)S(|T|).

We are also interested in the discrete sphere bundle of a triangulation.

Definition 4 (Discrete tangent vector; the discrete sphere bundle of a triangulation).

Let TT be a triangulation. A discrete tangent vector of TT is an ordered edge (x1,x2)(x_{1},x_{2}) in TT. The discrete sphere bundle of TT, denoted S(T)S(T), is the set of discrete tangent vectors of TT equipped with the metric d((x1,x2),(y1,y2))=max(d(x1,x2),d(y1,y2))d((x_{1},x_{2}),(y_{1},y_{2}))=max(d(x_{1},x_{2}),d(y_{1},y_{2})). Thus two distinct discrete tangent vectors (x1,x2)(x_{1},x_{2}) and (y1,y2)(y_{1},y_{2}) are adjacent if and only if x1x_{1} and y1y_{1} are either equal or adjacent, and x2x_{2} and y2y_{2} are either equal or adjacent. We will often abbreviate, for example, (x1,x2)(x_{1},x_{2}) to xx.

We can embed S(T)S(T) in S(|T|)S(|T|) by mapping to each (x1,x2)(x_{1},x_{2}) the tangent vector at |x1||x_{1}| along the edge |(x1,x2)||(x_{1},x_{2})|.

Let G0G_{0} be the set of pairs (T,[x])(T,[x]) where [x][x] is an orbit of the action of Aut(T)\mathrm{Aut}(T) on S(|T|)S(|T|). Since automorphisms are rare, we will usually think of such an orbit as a single point and write (T,x)(T,x) or simply xx if the context is clear. When we define the holonomy groupoid later on, G0G_{0} will be the set of units in that groupoid.

3.2. Topology on the space of pointed triangulations

Our next task is to define a topology on G0G_{0}. Loosely speaking, this topology is defined by a metric in which points that are close together on the same triangulation to be close, and points on different triangulations are close if those triangulations agree on a large radius up to a small change in decorations.

Definition 5 (Nearly decoration-preserving isometry).

Let TT and UU be triangulations with AS(|T|)A\subset S(|T|). If ϕ:AS(|U|)\phi:A\to S(|U|) is an isometry such that, for every discrete tangent vector xAx\in A, ϕ(x)\phi(x) is a discrete tangent vector in UU with dD(DT(x),DU(ϕ(x)))<ϵd_{D}(D_{T}(x),D_{U}(\phi(x)))<\epsilon, we say that ϕ\phi is an ϵ\epsilon-nearly decoration-preserving isometry (henceforth ϵ\epsilon-NDPI).

Definition 6 (Asymmetric distance in G0G_{0}).

Suppose that (T,[x]),(U,[y])𝒯(T,[x]),(U,[y])\in\mathcal{T}. For each x[x]x^{\prime}\in[x], rϕ0r_{\phi}\geq 0, and ϕ:B¯rϕ(x)S(|U|)\phi:\overline{B}_{r_{\phi}}(x^{\prime})\to S(|U|) an ϵϕ\epsilon_{\phi}-NDPI, let dϕd_{\phi} be the distance from ϕ(x)\phi(x^{\prime}) to [y][y]. We define

δ^((T,[x]),(U,[y]))=infmax(erϕ,edϕ,ϵϕ),\hat{\delta}((T,[x]),(U,[y]))=\inf\max(e^{-r_{\phi}},ed_{\phi},\epsilon_{\phi}),

where the infimum is over all xx^{\prime}, rϕr_{\phi}, and dϕd_{\phi} satisfying these conditions.

Lemma 1.

The function δ^\hat{\delta} defined this way is nondegenerate; that is,

δ^((T,[x]),(U,[y]))=0\hat{\delta}((T,[x]),(U,[y]))=0

only if (T,[x])(U,[y])(T,[x])\cong(U,[y]).

Proof.

If δ^((T,[x]),(U,[y]))=0\hat{\delta}((T,[x]),(U,[y]))=0, then there exist xx^{\prime} and ϕ\phi as in the definition with arbitrarily large rϕr_{\phi} and arbitrarily small dϕd_{\phi} and ϵϕ\epsilon_{\phi}. By composing with automorphisms if necessary, we may assume without loss of generality that x=xx^{\prime}=x and that d(ϕ(x),y)d(\phi(x),y) is arbitrarily small. Let ϕn\phi_{n} be such that rϕn>nr_{\phi_{n}}>n and d(ϕn(x),y)<1/nd(\phi_{n}(x),y)<1/n. Each zS(|T|)z\in S(|T|) can be written as expx(v)\exp_{x}(v) for some v3v\in\mathbb{R}^{3}. For sufficiently large nn, we have ϕn(B|v|(x))Bn(y)\phi_{n}(B_{|v|}(x))\in B_{n}(y). Then ϕn(z)=expϕn(x)(v)\phi_{n}(z)=\exp_{\phi_{n}(x)}(v) and ϕn(z)expy(v)\phi_{n}(z)\to\exp_{y}(v). This means that

ϕ:=limnϕn,\phi:=\lim_{n\to\infty}\phi_{n},

where the limit is taken pointwise, is a well-defined isometry from S(|T|)S(|T|) to S(|U|)S(|U|) that sends xx to yy. Furthermore, if zz is a discrete tangent vector in TT, then dD(DT(z),DU(ϕn(z)))<ϵd_{D}(D_{T}(z),D_{U}(\phi_{n}(z)))<\epsilon for every ϵ>0\epsilon>0 because some ϕn\phi_{n} is ϵ\epsilon-nearly decoration-preserving, and hence DT(z)=DU(ϕ(z))D_{T}(z)=D_{U}(\phi(z)). ∎

Lemma 2.

The function δ^\hat{\delta} defined this way satisfies the triangle inequality:

δ^((T,[x]),(V,[z]))δ^((T,[x]),(U,[y]))+δ^((U,[y]),(V,[z])).\hat{\delta}((T,[x]),(V,[z]))\leq\hat{\delta}((T,[x]),(U,[y]))+\hat{\delta}((U,[y]),(V,[z])).
Proof.

We can always take rϕr_{\phi} to be 0 and ϕ:{x}{y}\phi:\{x\}\to\{y\} to be the map sending xx to yy. Hence

δ^((T,[x]),(U,[y]))e,\hat{\delta}((T,[x]),(U,[y]))\leq e,

and we can restrict our attention to maps with dϕ<1d_{\phi}<1.

If (T,[x]),(U,[y]),(V,[z])𝒯(T,[x]),(U,[y]),(V,[z])\in\mathcal{T}, and we have a pair of maps ϕ1:B¯rϕ1(x)S(|U|)\phi_{1}:\overline{B}_{r_{\phi_{1}}}(x^{\prime})\to S(|U|) and ϕ2:B¯rϕ2(y)S(|V|)\phi_{2}:\overline{B}_{r_{\phi_{2}}}(y^{\prime})\to S(|V|), we can compose them as follows. By composing with an action of Aut(U)\mathrm{Aut}(U), we can assume without loss of generality that the distance from ϕ1(x)\phi_{1}(x^{\prime}) to yy^{\prime} is at most dϕ1d_{\phi_{1}}. We can then define

rϕ=min(rϕ1,rϕ2dϕ1),r_{\phi}=\min(r_{\phi_{1}},r_{\phi_{2}}-d_{\phi_{1}}),
ϵϕ=ϵϕ1+ϵϕ2,\epsilon_{\phi}=\epsilon_{\phi_{1}}+\epsilon_{\phi_{2}},

and

ϕ=ϕ2ϕ1.\phi=\phi_{2}\circ\phi_{1}.

If wB¯rϕ1(x)w\in\overline{B}_{r_{\phi_{1}}}(x^{\prime}) is a discrete tangent vector, we have

dD(DT(w),DV(ϕ(w)))\displaystyle d_{D}(D_{T}(w),D_{V}(\phi(w))) \displaystyle\leq dD(DT(w),DU(ϕ1(w)))+dD(DU(ϕ1(w)),DV(ϕ(w)))\displaystyle d_{D}(D_{T}(w),D_{U}(\phi_{1}(w)))+d_{D}(D_{U}(\phi_{1}(w)),D_{V}(\phi(w)))
\displaystyle\leq ϵϕ1+ϵϕ2\displaystyle\epsilon_{\phi_{1}}+\epsilon_{\phi_{2}}
=\displaystyle= ϵϕ,\displaystyle\epsilon_{\phi},

so ϕ\phi is ϵϕ\epsilon_{\phi}-nearly decoration-preserving. Then

edϕ\displaystyle ed_{\phi} \displaystyle\leq edϕ1+edϕ2\displaystyle ed_{\phi_{1}}+ed_{\phi_{2}}
\displaystyle\leq (δ^((T,[x]),(U,[y]))+δ^((U,[y]),(V,[z])))\displaystyle\left(\hat{\delta}((T,[x]),(U,[y]))+\hat{\delta}((U,[y]),(V,[z]))\right)

and

erϕ\displaystyle e^{-r_{\phi}} \displaystyle\leq max(erϕ1,edϕ1rϕ2)\displaystyle\max(e^{-r_{\phi_{1}}},e^{d_{\phi_{1}}-r_{\phi_{2}}})
\displaystyle\leq max(erϕ1,edϕ1+erϕ2)\displaystyle\max(e^{-r_{\phi_{1}}},ed_{\phi_{1}}+e^{-r_{\phi_{2}}})
\displaystyle\leq δ^((T,[x]),(U,[y]))+δ^((U,[y]),(V,[z])),\displaystyle\hat{\delta}((T,[x]),(U,[y]))+\hat{\delta}((U,[y]),(V,[z])),

where the second inequality comes from the fact that dϕ1<1d_{\phi_{1}}<1 and

ddtet<e\frac{d}{dt}e^{t}<e

for t<1t<1. ∎

Definition 7 (Distance in G0G_{0}).

We define

δ((T,[x]),(U,[y]))=δ^((T,[x]),(U,[y]))+δ^((U,[y]),(T,[x])).\delta((T,[x]),(U,[y]))=\hat{\delta}((T,[x]),(U,[y]))+\hat{\delta}((U,[y]),(T,[x])).
Theorem 1.

The function δ\delta defined in this way is a metric.

Proof.

Since δ^\hat{\delta} is nondegenerate and satisfies the triangle inequality, so does δ\delta. Furthermore, δ\delta is symmetric by definition. ∎

Definition 8 (The triangulation topology).

The triangulation topology is the topology induced by the metric δ\delta.

3.3. Foliated structure

Next, we will give G0G_{0} a foliated structure as defined by, for example, Moore and Schochet [12]. Foliated spaces differ from classical foliations in that the total space is not required to be a manifold; instead, a foliated space locally looks like the product of n\mathbb{R}^{n} with some model space (often, as in this case, a Cantor-like set). In this case our model space will be the discrete space G^0\hat{G}_{0} with the metric defined above. Short distances in this metric correspond to small motions of the marked tangent vector (motion in the leafwise direction), changes in the triangulation far from the marked tangent vector (motion in the transverse direction), and small changes in decorations on the triangulation (also motion in the transverse direction).

We will start by defining a useful family of subsets of G0G_{0} and G^0\hat{G}_{0}. These will be used to define the foliated structure, but, more importantly, will be a basis of the σ\sigma-algebra upon which our transverse measure is defined.

Definition 9 (Decorated finite patch).

By a decorated discrete finite patch we mean an ordered triple (A,x,Ω)(A,x,\Omega) where AA is a finite triangulation, xx is a marked discrete tangent vector of AA, and Ω\Omega is a Borel subset of E(A)𝒟\prod_{E(A)}\mathcal{D} (identified with the space of functions from E(A)E(A) to 𝒟\mathcal{D}). Generally we will refer to such a patch simply as AA and write xAx_{A} and ΩA\Omega_{A}. We similarly define a decorated continuous finite patch as an ordered triple where AA and Ω\Omega are as above and xS(|A|)x\in S(|A|). We call Ω\Omega the set of permissible decorations on AA. If |Ω|=1|\Omega|=1 we say the finite patch is exact. Likewise, we define a twice-marked decorated (discrete or continuous) finite patch as an ordered quadruple (A,x,y,Ω)(A,x,y,\Omega) where xx and yy are marked (discrete or continuous) tangent vectors.

Definition 10 (Sets of triangulations with a particular decorated finite patch).

Let AA be a decorated discrete finite patch. We define G^0AG^0\hat{G}_{0}^{A}\subset\hat{G}_{0} to be the set of triangulations with a marked discrete tangent vector (T,[x])(T,[x]) such that there exists gA,T:ATg_{A,T}:A\to T with the properties that:

  • gA,Tg_{A,T} is a simplicial embedding;

  • gA,T(xA)=xg_{A,T}(x_{A})=x;

  • the function xDT(gA,T(x))ΩAx\mapsto D_{T}(g_{A,T}(x))\in\Omega_{A}.

For any particular AA and TT, if such a gA,Tg_{A,T} exists, it is unique up to isomorphism (or choice of x[x]x\in[x]). We analogously define G^AG^\hat{G}^{A}\subset\hat{G} if AA is a twice-marked decorated discrete finite patch, G0AG0G_{0}^{A}\subset G_{0} if AA is a decorated continuous finite patch, or GAGG^{A}\subset G if AA is a twice-marked decorated continuous finite patch.

Definition 11 (Coordinate patches of G0G_{0}).

Let AA be a decorated continuous finite patch. Let B:SAB:S\to A be any smooth, but not necessarily isometric, embedding of a region S3S\subset\mathbb{R}^{3} into AA. We define ψA,B:G^0A×B\psi_{A,B}:\hat{G}_{0}^{A}\times B to G0G_{0} by

ψA,B(T,s)=(T,[gA,T(B(s))]),\psi_{A,B}(T,s)=(T,[g_{A,T}(B(s))]),

and call it the coordinate patch associated with the patch AA and region BB.

We can think of such a patch as a triangulation that is fixed in the region AA and allowed to vary outside of AA and whose decorations are allowed to vary within ΩA\Omega_{A}, with a marked tangent vector that is allowed to move within the subset BB.

Lemma 3.

The space G0G_{0} equipped with the coordinate patches ψA\psi_{A} is a foliated space.

Proof.

Since BB in the definition contains at most one point in any Aut(T)\mathrm{Aut}(T) orbit of S(|T|)S(|T|) for any TT containing AA, the coordinate patch ψA\psi_{A} is one-to-one. It follows from the manifold structure of S(|T|)S(|T|) that ψA,B1ψA,B\psi_{A^{\prime},B^{\prime}}^{-1}\psi_{A,B} is smooth. ∎

3.4. Holonomy groupoid

Our next objective is to construct GG, the holonomy groupoid of G0G_{0}. The precise construction of the holonomy groupoid is given in Moore and Schochet [12].

Briefly, a groupoid may be thought of as a small category in which all morphisms are invertible. It consists of a groupoid GG with a space of units G0G_{0}, a diagonal map Δ:G0G\Delta:G_{0}\to G, an inversion map 1\cdot^{-1} on GG, range and source maps r,s:GG0r,s:G\to G_{0}, and an associative multiplication on pairs (u,v)(u,v) where r(v)=s(u)r(v)=s(u). These must satisfy Δr(x)=Δs(x)=x\Delta r(x)=\Delta s(x)=x, uΔs(u)=Δr(u)u=uu\cdot\Delta s(u)=\Delta r(u)\cdot u=u, r(u1)=s(u)r(u^{-1})=s(u), and uu1=Δr(u)uu^{-1}=\Delta r(u). We denote Gx=s1(x)G_{x}=s^{-1}(x) and Gx=r1(x)G^{x}=r^{-1}(x).

One simple example of a groupoid is an equivalence relation \sim, where (x,y)G(x,y)\in G iff xyx\sim y, and we have r(x,y)=xr(x,y)=x, s(x,y)=ys(x,y)=y, (x,y)1=(y,x)(x,y)^{-1}=(y,x), and (x,y)(y,z)=(x,z)(x,y)(y,z)=(x,z). In fact, we will see that our holonomy groupoid is such a groupoid in the absence of symmetry.

In general the holonomy groupoid is constructed as follows. By a plaque we mean a connected component of U\ell\cap U where \ell is a leaf and UU is an open set. Given a path γ\gamma from xx to yy, and transversals NxN_{x} and NyN_{y} through xx and yy, respectively, we cover γ\gamma by small open sets V1,,VnV_{1},\ldots,V_{n}. If the ViV_{i} are sufficiently small, any plaque in ViV_{i} intersects a unique plaque in Vi+1V_{i+1}, so that we can start with the plaque in V1V_{1} containing some xNxx^{\prime}\in N_{x} and follow the sequence of intersecting plaques to a unique plaque in VnV_{n}, which intersects a unique point yy^{\prime} in NyN_{y}. Hence this results in a map from a neighborhood in NxN_{x} to one in NyN_{y}, realized by following a path from NxN_{x} to NyN_{y} sufficiently close to γ\gamma. The germ of this map is independent of the choice of ViV_{i}, and two paths correspond to the same holonomy element if the germs of their corresponding maps on NxN_{x} and NyN_{y} are the same.

Lemma 4.

The holonomy groupoid GG consists of Aut(T)\mathrm{Aut}(T) orbits of pairs of points [(x,y)][(x,y)], where x,yS(|T|)x,y\in S(|T|) for some T𝒯T\in\mathcal{T}.

Proof.

The leaves of G0G_{0} correspond to S(|T|)/Aut(T)S(|T|)/\mathrm{Aut}(T) with TT a triangulation. A plaque on a leaf of G0G_{0} corresponds to a small open subset of S(|T|)S(|T|).

Suppose that x,yS(|T|)x,y\in S(|T|) and γ\gamma is a path from xx to yy. We can choose a sufficiently large decorated continuous finite patch AA such that x=gA,T(xA)x=g_{A,T}(x_{A}) and γ\gamma is contained in the image of gA,Tg_{A,T}. Suppose additionally that V1,,VnV_{1},\ldots,V_{n} are open subsets of S(|A|)S(|A|) such that the gA(Vi)g_{A}(V_{i}) cover γ\gamma.

Suppose that (T,x)(T^{\prime},x^{\prime}) is another triangulation with x=gA,T(xA)x^{\prime}=g_{A,T^{\prime}}(x_{A}). Let y=gA,T(gA,T1(y))y^{\prime}=g_{A,T^{\prime}}(g_{A,T}^{-1}(y)); define yy^{\prime}, γ\gamma^{\prime}, and ViV^{\prime}_{i} analogously.

Then V1,,VnV^{\prime}_{1},\ldots,V^{\prime}_{n} cover γ\gamma^{\prime}, which is a path from xx^{\prime} to yy^{\prime}. It follows that the holonomy element corresponding to γ\gamma sends [x][x^{\prime}] to [y][y^{\prime}]; it depends on xx and yy but not on γ\gamma itself. This holds for any (T,x)G0A(T^{\prime},x^{\prime})\in G_{0}^{A}. Hence there is a unique holonomy element for each Aut(T)\mathrm{Aut}(T) orbits of pairs of points [(x,y)][(x,y)] on some triangulation TT. ∎

Definition 12 (The discrete holonomy groupoid and its space of units).

The holonomy groupoid GG contains a subgroupoid G^\hat{G} consisting of elements of the form [(|x|,|y|)][(|x|,|y|)] where xx and yy are discrete tangent vectors in S(T)S(T) for some triangulation TT. We denote its set of units by G^0\hat{G}_{0}.

If Aut(T)\mathrm{Aut}(T) is trivial, as we expect to be true in many cases, we can simply think of an element of GG as a triangulation with two marked tangent vectors, and an element of G^\hat{G} as a triangulation with two marked discrete tangent vectors. In any case, we will often suppress the brackets and write such an element (x,y)(x,y).

Our primary objects of interest will be G^\hat{G} and G^0\hat{G}_{0}. However, we will often need to refer to GG and G0G_{0} in order to use results about foliated spaces, as G^0\hat{G}_{0} does not have (for our purposes) a particularly useful foliated structure in its own right (for example, its leaves would consist of a single point).

3.5. Transverse measure

The notion of measure needed to consider a ”random triangulation” is a transverse measure. Moore and Schochet [12] define a transversal of a groupoid GG as a Borel subset of its space of units G0G_{0} whose intersection with each equivalence class (leaf, in this case) is countable, and a transverse measure as a measure on transversals satisfying certain properties. We will not repeat the precise definition here, as we intend to use a simpler characterization, also from Moore and Schochet [12]: a transverse measure is equivalent to a measure on a single complete transversal (that is, a transversal that intersects every leaf). Since G^0\hat{G}_{0} is such a transversal, we will simply think of our transverse measures as measures on G^0\hat{G}_{0}.

We will construct our measures as limits of finitely supported ones, taken in the following sense:

Definition 13 (Convergence of transverse measures).

We say that a sequence of transverse measures ν1,ν2,\nu_{1},\nu_{2},\ldots converges to a transverse measure ν\nu if νk(G^0A)ν(G^0A)\nu_{k}(\hat{G}_{0}^{A})\to\nu(\hat{G}_{0}^{A}) for every decorated discrete finite patch AA.

There are (at least) two useful means of obtaining a measure as a limit of finitely supported measures.

Definition 14 (Transverse measures constructed as limits of measures on spheres).

Let G^0S2\hat{G}_{0}^{S^{2}} be the subset of G^0\hat{G}_{0} consisting of points on finite (sphere) triangulations, and νk\nu_{k} be a sequence of measures on G^0S2\hat{G}_{0}^{S^{2}}. We require that νk(x)=νk(y)\nu_{k}(x)=\nu_{k}(y) for (T,x,y)G^(T,x,y)\in\hat{G}. In other words, the restriction of each νk\nu_{k} to a single sphere is a discrete uniform measure.

There are many possible choices for νk\nu_{k}. A few examples may be instructive:

  1. (1)

    Consider the case where 𝒟\mathcal{D} is trivial and νk\nu_{k} is a uniform probability measure on the sphere triangulations with at most kk faces. To our knowledge, it has not been proven whether or not the resulting sequence of measures converges, but in any event it is easy to show via diagonalization that a subsequence of the νk\nu_{k} converges. That is, let A1,A2,A_{1},A_{2},\ldots be an enumeration of decorated finite patches that includes, for every finite patch AA, a basis of the Borel σ\sigma-algebra on E(A)𝒟\prod_{E(A)}\mathcal{D}. Let ν0,j=νj\nu_{0,j}=\nu_{j}. For each kk, choose a subsequence νk,j\nu_{k,j} of ν(k1),j\nu_{(k-1),j} such that νk,jpk\nu_{k,j}\to p_{k} for some pk[0,1]p_{k}\in[0,1] (by compactness, such a subsequence must exist). Then the diagonal sequence νj,j\nu_{j,j} must have νj,j(Ak)pk\nu_{j,j}(A_{k})\to p_{k} for each kk. Henceforth we will assume that such a sequence νk\nu_{k} is fixed, and denote the limit by νu\nu_{u}.

    Conjecture 1.

    The sequence of measures νk\nu_{k} in this construction converges.

  2. (2)

    Let 𝒟={0,1}\mathcal{D}=\{0,1\} and let αk\alpha_{k} be the sphere triangulation formed by identifying the boundaries of two balls of radius kk on a 2-dimensional triangular grid. Define νk\nu_{k} to be the measure that decorates each edge of αk\alpha_{k} independently with 0 or 1 with equal probability and chooses a marked edge uniformly, and let νf\nu_{f} be the limit of the νk\nu_{k}. If AA is a finite patch that is not a subset of a 2-dimensional triangular grid, then νf(A)=0\nu_{f}(A)=0 since the identified boundaries on each αk\alpha_{k} have measures approaching 0. Hence νf\nu_{f} is a measure on decorated triangular grids, and G^\hat{G} is essentially a discrete tiling space. Without the decorations almost every leaf would be periodic and G^0\hat{G}_{0} would consist of a single point. However, with the decorations almost every leaf is aperiodic.

  3. (3)

    We can construct spheres via a process analogous to a substitution tiling. Let 𝒟\mathcal{D} be a finite set. Then there are only finitely many possible decorated faces. Choose a map β\beta that assigns to each decorated face with an ordering on the vertices (F,a,b,c)(F,a,b,c) a finite patch with three distinct marked points (β(F),β(a),β(b),β(c))(\beta(F),\beta(a),\beta(b),\beta(c)) such that β(F)D2\beta(F)\cong D^{2} and β(a)\beta(a), β(b)\beta(b), and β(c)\beta(c) are on the boundary of β(F)\beta(F) with no interior edges. Let β(ab¯)\beta(\overline{ab}) denote the portion of the boundary from β(a)\beta(a) to β(b)\beta(b) that does not pass through β(c)\beta(c), and define β(bc¯)\beta(\overline{bc}) and β(ca¯)\beta(\overline{ca}) analogously. We require that β(ab¯)\beta(\overline{ab}) depend only on D(a)D(a) and D(b)D(b) and the analogous conditions hold for β(bc¯)\beta(\overline{bc}) and β(ca¯)\beta(\overline{ca}). We can define β\beta on any triangulation by applying it to each face. We say that β\beta is nondegenerate if there exists kk such that βk(F)\beta^{k}(F) has at least one interior face for every face FF. This implies that there exists r<1r<1 such that the proportion of boundary faces to total faces in βn(F)\beta^{n}(F) is at most rnr^{n}. Then we can set T0T_{0} to be any sphere triangulation, set Tn=βn(T0)T_{n}=\beta^{n}(T_{0}), and choose νk\nu_{k} to be the uniform measure on TkT_{k}. Since β\beta is nondegenerate, almost all faces are in the interior (and eventually arbitrarily far from the boundary) of βn(F)\beta^{n}(F) for some FF, and we can determine the limiting frequency of each patch from analyzing β\beta, as in tiling theory. We denote the limit by νs\nu_{s}.

  4. (4)

    Let θ1\theta_{1} be the finite triangulation consisting of a single vertex of degree 7, its neighboring faces, and their vertices and edges. For n>0n>0, let θn\theta_{n} be constructed from θn1\theta_{n-1} by adding a face on every external edge, then adding additional vertices adjacent to each of the previous external vertices so that it becomes a vertex of degree 7, and connecting the newly added vertices in a cycle. The first three iterations are shown in 1.

    Figure 1. First three stages in the construction of νh\nu_{h}: θ1\theta_{1} in black, θ2θ1\theta_{2}\setminus\theta_{1} in red, and θ3θ2\theta_{3}\setminus\theta_{2} in blue.

    We can see that each boundary vertex has degree 3 or 4. If AA denotes a vertex of degree 3 and BB denotes a vertex of degree 4, the boundary of θn\theta_{n} is obtained from that of θn1\theta_{n-1} via the substitution

    ABAA;BBA.A\mapsto BAA;B\mapsto BA.

    The largest eigenvalue of the matrix of this substitution is 3+52\frac{3+\sqrt{5}}{2}, and hence the number of vertices in the boundary is asymptotically proportional to

    (3+52)n;\left(\frac{3+\sqrt{5}}{2}\right)^{n};

    in particular, it grows exponentially in nn. Furthermore, since 3+52\frac{3+\sqrt{5}}{2} is irrational, the limit of this substitution is aperiodic.

    Now let Θk\Theta_{k} be the sphere triangulation formed by identifying the boundaries of two copies of θk\theta_{k}. The boundaries become a cycle, which we will call the ridge, of vertices of degree 4 and 6 in a pattern given by the aforementioned substitution. All vertices not on the ridge have degree 7, and the structure of vertices in a ball intersecting the ridge is determined by the ridge. We define νh\nu_{h} to be the limit of the measures νk\nu_{k} that are uniform on the Θk\Theta_{k}. The proportion of vertices of degree 7 is asymptotically

    45+5<1,\frac{4}{5+\sqrt{5}}<1,

    meaning that vertices of degree 4 and 6 have positive measure.

    It is not too hard to see that νh\nu_{h} is well-defined. Indeed, with some work, we could compute it for any ball of finite radius. If the ball contains vertices of degree 4 and/or 6, and the vertices are consistent with the structure of Θn\Theta_{n} for large nn, the ball’s frequency is determined by the frequency of the pattern of ridge vertices (which can be determined with the usual techniques for substitution tilings). All balls of radius kk with only degree 7 vertices are isomorphic, and their frequency is simply the limiting proportion of vertices that are a distance more than kk from the ridge. Furthermore, because the limit of the substitution that produces the ridge is aperiodic, we need not consider automorphisms.

Another means of constructing transverse measures uses a single leaf. Not every leaf produces a well-defined transverse measure. In general, it is more difficult to construct measures this way compared to constructing them on spheres. However, this construction also leads to the idea of patch density, which will be useful in subsequent sections.

Definition 15 (Transverse measures constructed as limits of measures on a leaf; patch density on a leaf).

Let xG^0x\in\hat{G}_{0} and let f1,f2,f_{1},f_{2},\ldots be a sequence of functions on G^x\hat{G}^{x} such that fk𝑑λx=1\int f_{k}\,d\lambda^{x}=1 for each fkf_{k}. Often we will consider a leaf G^x\hat{G}^{x} that is recurrent in the sense of random walks and take fkf_{k} to be the distribution of a random walk from xx after kk steps. For a decorated discrete finite patch AA, we define νkx(A)=G^AG^xfk𝑑λ(x)\nu_{k}^{x}(A)=\int_{\hat{G}^{A}\cap\hat{G}^{x}}f_{k}\,d\lambda(x).

If νx(A)=limkνkx(A)\nu^{x}(A)=\lim_{k\to\infty}\nu_{k}^{x}(A) exists, we call it the patch density of AA on GxG^{x} (with respect to the exhaustion (fk)(f_{k})). If the exhaustion is not specified, we will assume it to be the exhaustion with respect to random walks.

4. Properties

4.1. Amenability

Our goal is to study operators on G^\hat{G} by averaging along leaves. If G^\hat{G} has two properties, amenability and ergodicity, such averages will be well-behaved. Additionally, when we construct a von Neumann algebra on G^\hat{G}, these properties will be useful for classifying it. We will first deal with amenability.

There are multiple equivalent criteria for amenability. The most useful one to us is Reiter’s criterion, which in our terminology is that there exists a sequence of functions gn:G^g_{n}:\hat{G}\to\mathbb{R} satisfying the following for ν\nu-a.e. xx:

  1. (1)

    gn𝑑λx=1\int g_{n}\,d\lambda_{x}=1, and

  2. (2)

    for all yGxy\in G^{x}, we have limngn(x,z)gn(y,z)dλx(z)=0\lim_{n\to\infty}\int g_{n}(x,z)-g_{n}(y,z)\,d\lambda^{x}(z)=0.

If almost all leaves had subexponential growth, we could simply take gng_{n} to be 1|Bn(x)|χBn(x)(y)\frac{1}{|B_{n}(x)|}\chi_{B_{n}(x)}(y). This does not work in the case of, for example, νh\nu_{h}, and so we resort to an argument using random walks.

Theorem 2.

If the groupoid G^\hat{G} is given a transverse measure such that GxG^{x} is recurrent for a.e. x, then G^\hat{G} is amenable.

Proof.

For TT a triangulation, xx a vertex of TT, and nn a positive integer, let B1x,B2x,B^{x}_{1},B^{x}_{2},\ldots denote a simple random walk on TT (that is, on the 1-skeleton of TT). To make our notation less cumbersome, we will use the shorthand x𝑛yx\xrightarrow[n]{}y to denote Bnx=yB^{x}_{n}=y. Let gng_{n} be the distribution of BxB^{x} after a uniformly random number of steps from 1,,n1,\ldots,n; that is,

gn(x,y)=1nk=1nP(x,y,k).g_{n}(x,y)=\frac{1}{n}\sum_{k=1}^{n}P(x,y,k).

Let ϵ>0\epsilon>0.

Let AxyA_{x}^{y} be the first time that the walk BxB^{x} hits yy - that is, the smallest positive integer such that BAxx=yB^{x}_{A_{x}}=y. If TT is recurrent and tt\in\mathbb{N} is sufficiently large, then P(Axy>t)<ϵ2P(A_{x}^{y}>t)<\frac{\epsilon}{2} and P(Ayx>t)<ϵ2P(A_{y}^{x}>t)<\frac{\epsilon}{2}. We will use x𝑛yx\xrightarrow{n}y to denote Axy=nA_{x}^{y}=n.

Now suppose that n>2tϵn>\frac{2t}{\epsilon}.

We have

gn(x,z)gn(y,z)\displaystyle g_{n}(x,z)-g_{n}(y,z) =\displaystyle= 1nk=1n(P(x𝑘z)P(y𝑘z))\displaystyle\frac{1}{n}\sum_{k=1}^{n}\left(P(x\xrightarrow[k]{}z)-P(y\xrightarrow[k]{}z)\right)
=\displaystyle= 1nm=1P(x𝑚y)k=1n(P(x𝑘z|x𝑚y)P(y𝑘z|x𝑚y)).\displaystyle\frac{1}{n}\sum_{m=1}^{\infty}P(x\xrightarrow{m}y)\sum_{k=1}^{n}\left(P(x\xrightarrow[k]{}z|x\xrightarrow{m}y)-P(y\xrightarrow[k]{}z|x\xrightarrow{m}y)\right).

Our next step is to decompose this sum into two parts: one with mtm\leq t and the other with m>tm>t. (That is, either the random walk starting at xx hits yy before time tt, in which case the distribution of a random walk starting at xx is similar to one starting at yy, or it does not hit yy before time tt, which is unlikely because of our choice of tt.)

First, we will consider the sum where mtm\leq t; that is, the random walk starting at xx hits yy before time tt. If mtm\leq t, we have

k=1n(P(x𝑘z|x𝑚y)P(y𝑘z|x𝑚y))\displaystyle\sum_{k=1}^{n}\left(P(x\xrightarrow[k]{}z|x\xrightarrow{m}y)-P(y\xrightarrow[k]{}z|x\xrightarrow{m}y)\right)
=\displaystyle= k=1mP(x𝑘z|x𝑚y)+k=m+1nP(x𝑘z|x𝑚y)\displaystyle\sum_{k=1}^{m}P(x\xrightarrow[k]{}z|x\xrightarrow{m}y)+\sum_{k=m+1}^{n}P(x\xrightarrow[k]{}z|x\xrightarrow{m}y)
j=1nmP(y𝑗z=z|x𝑚y)j=nm+1nP(y𝑗z|x𝑚y)\displaystyle-\sum_{j=1}^{n-m}P(y\xrightarrow[j]{}z=z|x\xrightarrow{m}y)-\sum_{j=n-m+1}^{n}P(y\xrightarrow[j]{}z|x\xrightarrow{m}y)
\displaystyle\leq k=m+1nP(x𝑘z|x𝑚y)j=1nmP(y𝑗z|x𝑚y)+j=nm+1nP(x𝑗z|x𝑚y)\displaystyle\sum_{k=m+1}^{n}P(x\xrightarrow[k]{}z|x\xrightarrow{m}y)-\sum_{j=1}^{n-m}P(y\xrightarrow[j]{}z|x\xrightarrow{m}y)+\sum_{j=n-m+1}^{n}P(x\xrightarrow[j]{}z|x\xrightarrow{m}y)
=\displaystyle= j=1nm(P(xj+mz|x𝑚y)P(y𝑗z=z|x𝑚y))+k=nm+1nP(x𝑘z|x𝑚y).\displaystyle\sum_{j=1}^{n-m}\left(P(x\xrightarrow[j+m]{}z|x\xrightarrow{m}y)-P(y\xrightarrow[j]{}z=z|x\xrightarrow{m}y)\right)+\sum_{k=n-m+1}^{n}P(x\xrightarrow[k]{}z|x\xrightarrow{m}y).

Now, if x𝑚yx\xrightarrow{m}y, then x𝑚yx\xrightarrow[m]{}y and Bj+mxB_{j+m}^{x} is a random walk of jj steps starting at yy. Hence each term in the first sum is 0, and we are left with

k=1n(P(x𝑘z|x𝑚y)P(y𝑘z|x𝑚y))\displaystyle\sum_{k=1}^{n}\left(P(x\xrightarrow[k]{}z|x\xrightarrow{m}y)-P(y\xrightarrow[k]{}z|x\xrightarrow{m}y)\right) \displaystyle\leq k=nm+1nP(x𝑘z|x𝑚y)\displaystyle\sum_{k=n-m+1}^{n}P(x\xrightarrow[k]{}z|x\xrightarrow{m}y)
\displaystyle\leq m\displaystyle m
\displaystyle\leq t\displaystyle t
m=1tP(x𝑚y)k=1n(P(x𝑘z|x𝑚y)P(y𝑘z|x𝑚y))\displaystyle\sum_{m=1}^{t}P(x\xrightarrow{m}y)\sum_{k=1}^{n}\left(P(x\xrightarrow[k]{}z|x\xrightarrow{m}y)-P(y\xrightarrow[k]{}z|x\xrightarrow{m}y)\right) \displaystyle\leq tm=1tP(x𝑚y)\displaystyle t\sum_{m=1}^{t}P(x\xrightarrow{m}y)
\displaystyle\leq t.\displaystyle t.

Next, consider the sum with m>tm>t. We have

m=t+1P(x𝑚y)k=1n(P(x𝑘z|x𝑚y)P(y𝑘z|x𝑚y))\displaystyle\sum_{m=t+1}^{\infty}P(x\xrightarrow{m}y)\sum_{k=1}^{n}\left(P(x\xrightarrow[k]{}z|x\xrightarrow{m}y)-P(y\xrightarrow[k]{}z|x\xrightarrow{m}y)\right)
\displaystyle\leq m=t+1P(x𝑚y)k=1nP(x𝑘z|x𝑚y)\displaystyle\sum_{m=t+1}^{\infty}P(x\xrightarrow{m}y)\sum_{k=1}^{n}P(x\xrightarrow[k]{}z|x\xrightarrow{m}y)
\displaystyle\leq m=t+1nP(x𝑚y)\displaystyle\sum_{m=t+1}^{\infty}nP(x\xrightarrow{m}y)
=\displaystyle= nP(Axy>t)\displaystyle nP(A_{x}^{y}>t)
<\displaystyle< nϵ2.\displaystyle\frac{n\epsilon}{2}.

Hence

m=1P(x𝑚y)k=1n(P(x𝑘z|x𝑚y)P(y𝑘z|x𝑚y))\displaystyle\sum_{m=1}^{\infty}P(x\xrightarrow{m}y)\sum_{k=1}^{n}\left(P(x\xrightarrow[k]{}z|x\xrightarrow{m}y)-P(y\xrightarrow[k]{}z|x\xrightarrow{m}y)\right) \displaystyle\leq t+nϵ2\displaystyle t+\frac{n\epsilon}{2}
gn(x,z)gn(y,z)\displaystyle g_{n}(x,z)-g_{n}(y,z) \displaystyle\leq 1n(t+nϵ2)\displaystyle\frac{1}{n}\left(t+\frac{n\epsilon}{2}\right)
=\displaystyle= tn+ϵ2\displaystyle\frac{t}{n}+\frac{\epsilon}{2}
\displaystyle\leq ϵ2+ϵ2\displaystyle\frac{\epsilon}{2}+\frac{\epsilon}{2}
=\displaystyle= ϵ.\displaystyle\epsilon.

Exchanging xx and yy gives

gn(y,z)gn(x,z)<ϵg_{n}(y,z)-g_{n}(x,z)<\epsilon

and hence

|gn(x,z)gn(y,z)|<ϵ,|g_{n}(x,z)-g_{n}(y,z)|<\epsilon,

and we have shown that G^\hat{G} is amenable. ∎

Corollary 1.

The groupoid G^\hat{G} with any transverse measure ν\nu constructed as in Definition 14 is amenable.

Proof.

Using the result of Benjamini and Schramm [6], TT is recurrent for ν\nu-a.e. (T,[x])(T,[x]). By the preceding lemma, it suffices to show that this implies GxG^{x} is recurrent. Recall that each point yGxy\in G^{x} is actually a discrete tangent vector, or directed edge, (y1,y2)E(T)(y_{1},y_{2})\in E(T), and that two such distinct tangent vectors (y1,y2)(y_{1},y_{2}) and (z1,z2)(z_{1},z_{2}) are adjacent if and only if y1y_{1} and z1z_{1} are either equal or adjacent, and y2y_{2} and z2z_{2} are either equal or adjacent. We will use the flow theorem, which states that a graph Γ\Gamma is transient if and only if it has a flow ρ\rho with a single source, no sinks, and a finite energy

(Γ)=(x,y)E(Γ)(ρ(x,y))2<.\mathcal{E}(\Gamma)=\sum_{(x,y)\in E(\Gamma)}\left(\rho(x,y)\right)^{2}<\infty.

Suppose that GxG^{x}, where xx is a tangent vector on the triangulation TT, is transient. Then there exists a flow ρ\rho with a single source w=(w1,w2)Gxw=(w_{1},w_{2})\in G^{x} and finite energy. Define a flow on TT by

ρ(y,z)=z2B1(z)y2B1(y)ρ((y,y2),(z,z2)).\rho^{\prime}(y,z)=\sum_{\stackrel{{\scriptstyle y_{2}\in B_{1}(y)}}{{z_{2}\in B_{1}(z)}}}\rho((y,y_{2}),(z,z_{2})).

We note that

div(ρ)(y)=y2B1(y)div(ρ)(y,y2),\mathrm{div}(\rho^{\prime})(y)=\sum_{y_{2}\in B_{1}(y)}\mathrm{div}(\rho)(y,y_{2}),

and hence ρ\rho^{\prime} has a single source at ww and no sinks.

Furthermore,

(ρ(y,z))2\displaystyle\left(\rho^{\prime}(y,z)\right)^{2} =\displaystyle= (z2B1(z)y2B1(y)ρ((y,y2),(z,z2)))2\displaystyle\left(\sum_{\stackrel{{\scriptstyle y_{2}\in B_{1}(y)}}{{z_{2}\in B_{1}(z)}}}\rho((y,y_{2}),(z,z_{2}))\right)^{2}
\displaystyle\leq z2B1(z)y2B1(y)ρ((y,y2),(z,z2))2|B1(y)||B1(z)|\displaystyle\sum_{\stackrel{{\scriptstyle y_{2}\in B_{1}(y)}}{{z_{2}\in B_{1}(z)}}}\rho((y,y_{2}),(z,z_{2}))^{2}|B_{1}(y)||B_{1}(z)|
\displaystyle\leq (d+1)2z2B1(z)y2B1(y)ρ((y,y2),(z,z2))2.\displaystyle(d+1)^{2}\sum_{\stackrel{{\scriptstyle y_{2}\in B_{1}(y)}}{{z_{2}\in B_{1}(z)}}}\rho((y,y_{2}),(z,z_{2}))^{2}.

Taking sums gives

(ρ)(d+1)2(ρ)<,\mathcal{E}(\rho^{\prime})\leq(d+1)^{2}\mathcal{E}(\rho)<\infty,

from which we conclude that TT is transient. By contrapositive, if TT is recurrent, then GxG^{x} is recurrent. ∎

The case of measures constructed using a leaf as in Definition 15 is less clear.

Conjecture 2.

If GxG^{x} is recurrent and νx\nu_{x} exists, then GyG^{y} is recurrent for νx\nu_{x}-a.e. y.

If this conjecture holds, it will also imply that G^\hat{G} is amenable when equipped with the transverse measure νx\nu_{x} constructed subject to these conditions.

4.2. Ergodicity

Next we consider ergodicity. We will start by proving the following lemma:

Lemma 5.

Suppose that ν\nu is chosen such that, for any decorated finite patch AA, νx=ν\nu^{x}=\nu for ν\nu-almost every xx (that is, the patch density of any finite patch AA almost surely does not depend on xx). For all f,gL1(G0)f,g\in L^{1}(G^{0}), we have

limnf(x)P(x𝑛y)g(y)𝑑λx(y)𝑑μ(x)=f𝑑μg𝑑μ.\lim_{n\to\infty}\int f(x)\int P(x\xrightarrow[n]{}y)g(y)\,d\lambda^{x}(y)\,d\mu(x)=\int f\,d\mu\int g\,d\mu.
Proof.

For each finite patch AA, let ϕ(A)=ν(G^A)\phi(A)=\nu(\hat{G}^{A}). If AA and BB are decorated finite patches and f=χG^Af=\chi_{\hat{G}^{A}} and g=χG^Bg=\chi_{\hat{G}^{B}}, the claim holds since both sides equal ϕ(A)ϕ(B)\phi(A)\phi(B). The claim in general follows from bilinearity and the fact that functions of the form χG^A\chi_{\hat{G}^{A}} span a dense subset of L1(G^0)L^{1}(\hat{G}^{0}). ∎

Theorem 3.

Suppose that ν\nu is chosen such that, for any decorated finite patch AA, νx=ν\nu^{x}=\nu for ν\nu-almost every xx. Then GG equipped with the measure ν\nu is ergodic.

Proof.

Suppose that ff is an invariant, real-valued function (that is, g(y)=f(x)g(y)=f(x) for any yGxy\in G^{x}) with f𝑑μ=0\int f\,d\mu=0. Setting g=fg=f in Lemma 5,

0\displaystyle 0 =\displaystyle= (f𝑑μ)2\displaystyle\left(\int f\,d\mu\right)^{2}
=\displaystyle= limnf(x)P(x𝑛y)f(x)𝑑λx(y)𝑑μ(x)\displaystyle\lim_{n\to\infty}\int f(x)\int P(x\xrightarrow[n]{}y)f(x)\,d\lambda^{x}(y)\,d\mu(x)
=\displaystyle= limn(f(x))2𝑑μ(x)\displaystyle\lim_{n\to\infty}\int(f(x))^{2}\,d\mu(x)

and hence f0f\equiv 0 μ\mu-a.e. Since every such ff is zero, GG is ergodic. ∎

Conjecture 3.

The measure νu\nu_{u} is ergodic.

5. Operator algebras

5.1. Construction

We are interested in the properties of operators on GG. To study these operators, we will consider operator algebras. Both the CC^{*}-algebra and the von Neumann algebra are generated using leafwise convolution operators.

Definition 16 (The reduced CC^{*}-algebra of GG).

We construct Cr(G)C^{*}_{r}(G) as a CC^{*}-algebra operating on xG0L2(Gx)\bigoplus_{x\in G_{0}}L^{2}(G_{x}). To each continuous, compactly supported function ff on GG, and for each xG0x\in G_{0}, we associate the operator πx(f)\pi_{x}(f) on L2(Gx)L^{2}(G_{x}) given by

πx(f)(ϕ)(z)=yGxϕ(y)f(y,z)𝑑λx(y)\pi_{x}(f)(\phi)(z)=\int_{y\in G_{x}}\phi(y)f(y,z)\,d\lambda_{x}(y)

for each ϕL2(Gx)\phi\in L^{2}(G_{x}). The closure of all πC(f)={πx(f):xG0}\pi_{C^{*}}(f)=\{\pi_{x}(f):x\in G_{0}\} under the norm

πC(f)=sup(T,[x])Gπx(f)\|\pi_{C^{*}}(f)\|=\sup_{(T,[x])\in G}\|\pi_{x}(f)\|

forms the CC^{*}-algebra Cr(G)C^{*}_{r}(G) associated to GG.

Definition 17 (The von Neumann algebras of GG and G^\hat{G}).

We construct W(G)W^{*}(G) as a von Neumann algebra operating on L2(G)L^{2}(G). Using the transverse measure ν\nu defined above, we can construct the spaces Lp(G)L^{p}(G). For fL1(G)f\in L^{1}(G) and ϕL2(G)\phi\in L^{2}(G), we define the operator

π(f)(ϕ)([(x,z)])=yf(x,y)ϕ(y,z)𝑑λx(y).\pi(f)(\phi)([(x,z)])=\int_{y}f(x,y)\phi(y,z)\,d\lambda_{x}(y).

The weak closure of all such π(f)\pi(f) is the von Neumann algebra W(G)W^{*}(G) associated with GG. That is, HB(L2(G))H\in B(L^{2}(G)) belongs to W(G)W^{*}(G) if there exists a sequence (fn)(f_{n}) such that π(fn)ϕ,ηHϕ,η\langle\pi(f_{n})\phi,\eta\rangle\to\langle H\phi,\eta\rangle for all ϕ,ηL2(G)\phi,\eta\in L^{2}(G).

We can repeat this construction using G^\hat{G} in place of GG. Since G^\hat{G} is a complete transversal of GG, we have

W(G)W(G^)(L2([0,1]))W^{*}(G)\cong W^{*}(\hat{G})\otimes\mathcal{B}(L^{2}([0,1]))

by Moore and Schochet [12] (Proposition 6.21).

The CC^{*} and von Neumann algebras are related as follows.

Lemma 6.

If ff is a continuous, compactly supported function, then

|π(f)||πC(f)|.|\pi(f)|\leq|\pi_{C^{*}}(f)|.
Proof.

For ϕL2(G)\phi\in L^{2}(G),

|π(f)ϕ|\displaystyle|\pi(f)\phi| =\displaystyle= (|π(f)ϕ(x,y)|2𝑑λx(y)𝑑μ(x))1/2\displaystyle\left(\int\int|\pi(f)\phi(x,y)|^{2}\,d\lambda_{x}(y)\,d\mu(x)\right)^{1/2}
=\displaystyle= (|πx(f)ϕ(x,y)|2𝑑λx(y)𝑑μ(x))1/2\displaystyle\left(\int\int|\pi_{x}(f)\phi(x,y)|^{2}\,d\lambda_{x}(y)\,d\mu(x)\right)^{1/2}
=\displaystyle= (|πx(f)ϕ|Gx|2dμ(x))1/2\displaystyle\left(\int|\pi_{x}(f)\phi|_{G^{x}}|^{2}\,d\mu(x)\right)^{1/2}
\displaystyle\leq (|πx(f)|2|ϕGx|2𝑑μ(x))1/2\displaystyle\left(\int|\pi_{x}(f)|^{2}|\phi_{G^{x}}|^{2}\,d\mu(x)\right)^{1/2}
\displaystyle\leq (|πC(f)|2ϕGx|2dμ(x))1/2\displaystyle\left(\int|\pi_{C^{*}}(f)|^{2}\phi_{G^{x}}|^{2}\,d\mu(x)\right)^{1/2}
=\displaystyle= |ϕ||πC(f)|.\displaystyle|\phi||\pi_{C^{*}}(f)|.

Corollary 2.

There exists an algebra homomorphism Φ:Cr(G)W(G)\Phi:C^{*}_{r}(G)\to W^{*}(G) such that Φ(πC(f))=π(f)\Phi(\pi_{C}^{*}(f))=\pi(f).

Proof.

Let HCr(G)H\in C^{*}_{r}(G). Choose a sequence of continuous, compactly supported functions (fn)(f_{n}) with πC(fn)H\pi_{C^{*}}(f_{n})\to H. Since πC(fn)\pi_{C^{*}}(f_{n}) is Cauchy, π(fn)\pi(f_{n}) is also Cauchy by the preceding lemma, and hence converges. Let Φ(H)=limπ(fn)\Phi(H)=\lim\pi(f_{n}). To show that this is well-defined, let (gn)(g_{n}) be another such sequence of functions. Then πC(fngn)0\pi_{C^{*}}(f_{n}-g_{n})\to 0 and the preceding lemma gives limπ(fn)=limπ(gn)\lim\pi(f_{n})=\lim\pi(g_{n}). That Φ\Phi is an algebra homomorphism follows from the definition. ∎

For our results the most important algebra is W(G^)W^{*}(\hat{G}). We will first characterize the operators in W(G^)W^{*}(\hat{G}), then discuss its classification.

Definition 18 (The diagonal function).

Define D:G^D:\hat{G}\to\mathbb{C} by D(T,x,y)=1D(T,x,y)=1 if x=yx=y and D(T,x,y)=0D(T,x,y)=0 otherwise.

It is easily seen that DL1(G^)L2(G^)D\in L^{1}(\hat{G})\cap L^{2}(\hat{G}) and that DD is the identity of leafwise convolution; that is, π(D)=1\pi(D)=1.

Lemma 7.

Let f,gL2(G^)f,g\in L^{2}(\hat{G}) and hL1(G^)h\in L^{1}(\hat{G}). Then hg,f=h,fg\langle h*g,f\rangle=\langle h,f*g^{*}\rangle.

Proof.

We have

hg,f\displaystyle\langle h*g,f\rangle =\displaystyle= (hg)(x,y)f(x,y)𝑑λx(y)𝑑μ(x)\displaystyle\int(h*g)(x,y)f(x,y)\,d\lambda^{x}(y)\,d\mu(x)
=\displaystyle= h(x,z)g(z,y)f(x,y)𝑑λx(z)𝑑λx(y)𝑑μ(x)\displaystyle\int h(x,z)g(z,y)f(x,y)\,d\lambda^{x}(z)\,d\lambda^{x}(y)\,d\mu(x)
=\displaystyle= h,fg.\displaystyle\langle h,f*g^{*}\rangle.

Theorem 4.

Let HH be a bounded linear operator on L2(G)L^{2}(G). The following are equivalent:

  1. (1)

    HW(G^)H\in W^{*}(\hat{G}).

  2. (2)

    H(D)g=H(g)H(D)*g=H(g) for every gL2(G^)g\in L^{2}(\hat{G}).

  3. (3)

    H=π(h)H=\pi(h) for some hL2(G^)h\in L^{2}(\hat{G}).

Proof.

If HH satisfies (2), let h=H(D)h=H(D). For every gL2(G^)g\in L^{2}(\hat{G}), we have H(g)=H(D)g=hgH(g)=H(D)*g=h*g; that is, H=π(h)H=\pi(h).

If HH satisfies (3), we can approximate hh as a norm, hence weak, limit of compactly supported functions hkL2(G^)h_{k}\in L^{2}(\hat{G}). For every g,fL2(G^)g,f\in L^{2}(\hat{G}),

limkhkg,f\displaystyle\lim_{k\to\infty}\langle h_{k}*g,f\rangle =\displaystyle= limkhk,fg\displaystyle\lim_{k\to\infty}\langle h_{k},f*g^{*}\rangle
=\displaystyle= h,fg;\displaystyle\langle h,f*g^{*}\rangle;

that is, HH is a weak limit of the π(hk)\pi(h_{k}).

Finally, suppose that HH satisfies (1). Then we can write HH as the weak limit of Hk=π(hk)H_{k}=\pi(h_{k}) with hkh_{k} compactly supported. Define h=H(D)h=H(D). Since

Hk(D)g,f=Hk(D),fg,\langle H_{k}(D)*g,f\rangle=\langle H_{k}(D),f*g^{*}\rangle,

weak convergence implies

H(D)g,f\displaystyle\langle H(D)*g,f\rangle =\displaystyle= H(D),fg\displaystyle\langle H(D),f*g^{*}\rangle
=\displaystyle= limkHk(D),fg\displaystyle\lim_{k\to\infty}\langle H_{k}(D),f*g^{*}\rangle
=\displaystyle= limkHk(D)g,f\displaystyle\lim_{k\to\infty}\langle H_{k}(D)*g,f\rangle
=\displaystyle= limkHk(g),f\displaystyle\lim_{k\to\infty}\langle H_{k}(g),f\rangle
=\displaystyle= H(g),f.\displaystyle\langle H(g),f\rangle.

Because ff was arbitrary, this implies H(D)g=H(g)H(D)*g=H(g). ∎

This result is convenient because it allows us to reduce questions about operators in W(G^)W^{*}(\hat{G}) to questions about convolutions of functions on L2(G^)L^{2}(\hat{G}). In the following discussion, we will identify a function ff with π(f)\pi(f). In addition, if (A,x,y)(A,x,y) is a twice-marked decorated discrete finite patch we will write χ(A,x,y)\chi_{(A,x,y)} in place of χG(A,x,y)\chi_{G^{(}A,x,y)}.

Although the structure of G^\hat{G} itself is essentially symmetric with respect to its first and second argument, W(G^)W^{*}(\hat{G}) treats the two arguments quite differently, essentially because it is defined using left rather than right convolution. The following corollary illustrates this.

Corollary 3.

Suppose that xG0x\in G^{0}, y,yGxy,y^{\prime}\in G^{x}, HW(G^)H^{*}\in W^{*}(\hat{G}), and ϕ(z,y)=ϕ(z,y)\phi(z,y)=\phi(z,y^{\prime}) for all zGxz\in G^{x}. Then Hϕ(x,y)=Hϕ(x,y)H\phi(x,y)=H\phi(x,y^{\prime}).

Proof.

By Theorem 4, we can write H=π(h)H=\pi(h). Then

Hϕ(x,y)\displaystyle H\phi(x,y) =\displaystyle= h(x,z)ϕ(z,y)𝑑λx(z)\displaystyle\int h(x,z)\phi(z,y)\,d\lambda_{x}(z)
=\displaystyle= h(x,z)ϕ(z,y)𝑑λx(z)\displaystyle\int h(x,z)\phi(z,y^{\prime})\,d\lambda_{x}(z)
=\displaystyle= Hϕ(x,y).\displaystyle H\phi(x,y^{\prime}).

In other words, operators in W(G^)W^{*}(\hat{G}) preserve invariance with respect to the second argument. (They do not preserve invariance with respect to the first argument; consider the example where hh is a diagonal function taking on two different values.) We might think of operators in W(G^)W^{*}(\hat{G}) as acting on the first argument only, with the second argument being a convenient way to keep track of the foliated structure.

Our next goal is to classify W(G^)W^{*}(\hat{G}).

Lemma 8.

Suppose that fL2(G^)f\in L^{2}(\hat{G}) and gL2(G0^)g\in L^{2}(\hat{G_{0}}). Define gD(x,z)=D(x,z)g(x)=D(x,z)g(z)g_{D}(x,z)=D(x,z)g(x)=D(x,z)g(z). Then

fgD=gDff*g_{D}=g_{D}*f

if and only if

g(x)f(x,z)=g(z)f(x,z)g(x)f(x,z)=g(z)f(x,z)

for almost every (x,z)G^(x,z)\in\hat{G}.

Proof.

We have

gDf(x,z)\displaystyle g_{D}*f(x,z) =\displaystyle= gD(x,y)f(y,z)𝑑λx(y)\displaystyle\int g_{D}(x,y)f(y,z)\,d\lambda^{x}(y)
=\displaystyle= g(x)f(x,z)\displaystyle g(x)f(x,z)

and, similarly,

fgD(x,z)=g(z)f(x,z).f*g_{D}(x,z)=g(z)f(x,z).

Lemma 9.

Let fL2(G^)f\in L^{2}(\hat{G}). Then π(f)Z(G^)\pi(f)\in Z(\hat{G}) if and only if f(x,y)=0f(x,y)=0 and f(x,x)=f(x,y)f(x,x)=f(x,y) for almost every (T,x,y)(T,x,y) with xyx\not=y.

Proof.

Suppose that f(x,y)=0f(x,y)=0 and f(x,x)=f(x,y)f(x,x)=f(x,y) for almost every (T,x,y)(T,x,y) with xyx\not=y. Then f=gDf=g_{D} for some gL2(G0^)g\in L^{2}(\hat{G_{0}}). For every hL2(G^)h\in L^{2}(\hat{G}) and (x,z)G^(x,z)\in\hat{G}, we have

g(x)h(x,z)=g(z)h(x,z),g(x)h(x,z)=g(z)h(x,z),

and the preceding lemma implies hgD=gDhh*g_{D}=g_{D}*h; i. e., f=gDZ(W(G^))f=g_{D}\in Z(W^{*}(\hat{G})).

Conversely, suppose that fZ(G^)f\in Z(\hat{G}). For every gL2(G0^)g\in L^{2}(\hat{G_{0}}), the preceding lemma gives

g(x)f(x,z)=g(z)f(x,z)g(x)f(x,z)=g(z)f(x,z)

for almost all (x,z)(x,z). This in particular holds when g=χ(A,x)g=\chi_{(A,x)} for some discrete decorated finite patch (A,x)(A,x). We can choose a countable set of finite patches (Ak,vk)(A_{k},v_{k}) such that the χ(Ak,vk)\chi_{(A_{k},v_{k})} generate the σ\sigma-algebra of G^\hat{G}. If xzx\not=z, then there must exist some kk such that χ(Ak,vk)(x)χ(Ak,vk)(z)\chi_{(A_{k},v_{k})}(x)\not=\chi_{(A-k,v_{k})}(z); otherwise (T,x)(T,x) would be isomorphic to (T,z)(T,z), and xx and zz would be two representatives of the same orbit of Aut(T)\mathrm{Aut}(T). Hence f(x,z)=0f(x,z)=0 for almost all (x,z)(x,z) with xzx\not=z, and we may assume that f=hDf=h_{D} for some hL2(G0^)h\in L^{2}(\hat{G_{0}}).

If gL2(G^)g\in L^{2}(\hat{G}) is arbitrary, Lemma 9 gives

h(x)g(x,z)=h(z)g(x,z)h(x)g(x,z)=h(z)g(x,z)

for almost all (x,z)(x,z). By choosing a gL2(G^)g\in L^{2}(\hat{G}) that is nowhere zero, this implies that h(x)=h(z)h(x)=h(z) for almost all (x,z)(x,z); that is, hh is constant on almost every leaf. ∎

Corollary 4.

The algebra W(G^)W^{*}(\hat{G}) is a factor if and only if ν\nu is ergodic.

Proof.

If G^\hat{G} is ergodic and fZ(G^)f\in Z(\hat{G}), the above lemma implies that f(x,y)=0f(x,y)=0 and f(x,x)=f(x,y)f(x,x)=f(x,y) for almost every (T,x,y)(T,x,y) with xyx\not=y; that is, f=hDf=h_{D} for some hL2(G^)h\in L^{2}(\hat{G}), and hh is constant on almost every leaf. Since GG is ergodic, this means that that hh is constant, and ff is a multiple of DD.

Conversely, if G^\hat{G} is not ergodic, there exists a function hL2(G^0)h\in L^{2}(\hat{G}_{0}) that is not constant but is constant on almost every leaf. Then hDZ(G^)h_{D}\in Z(\hat{G}), but hDh_{D} is not a multiple of DD. ∎

5.2. Trace

Many of our results will involve traces on W(G)W^{*}(G) and W(G^)W^{*}(\hat{G}). These traces are closely related to the transverse measure ν\nu. First we will show:

Theorem 5.

The transverse measure ν\nu is invariant; that is, νr=νs\nu^{r}=\nu^{s} where rr denotes the range map, ss denotes the source map, and

νr(S)=G0|Sr1(x)dν(x)\nu_{r}(S)=\int_{G_{0}}|S\cap r^{-1}(x)\,d\nu(x)

and

νs(S)=G0|Ss1(x)dν(x).\nu_{s}(S)=\int_{G_{0}}|S\cap s^{-1}(x)\,d\nu(x).
Proof.

Suppose that AA is a finite triangulation with marked tangent vectors pp and qq. For each x=(T,v)G0x=(T,v)\in G^{0},

G0Ar1(x)G_{0}^{A}\cap r^{-1}(x)

is the number of times (A,q)(A,q) appears in (T,v)(T,v). This is 1 if vv is contained in a marked patch that looks like (A,q)(A,q) in TT, and 0 otherwise. Furthermore,

νr(G0A)=X|G0Ar1(x)|𝑑ν(x)\nu_{r}(G_{0}^{A})=\int_{X}|G_{0}^{A}\cap r^{-1}(x)|\,d\nu(x)

is simply the probability of the marked patch (A,q)(A,q) appearing in a randomly chosen marked triangulation in G0G_{0}. Similarly, νs(G0A)\nu_{s}(G_{0}^{A}) is the probability of (A,p)(A,p) appearing in a randomly chosen marked triangulation.

Suppose that a sphere triangulation contains kk distinct (though not necessarily disjoint) copies of AA. Each copy of AA must contain exactly |Aut(A)||Aut(A)| copies of pp (that is, tangent vectors pp^{\prime} such that (A,p)(A,p^{\prime}) is isomorphic to (A,p)(A,p)), and exactly |Aut(A)||Aut(A)| copies of qq. Furthermore, no two distinct copies of AA can share a copy of pp (respectively, qq), since knowing the location of pp specifies the location of all other faces of AA. Since the marked tangent vector is chosen uniformly, (A,p)(A,p) and (A,q)(A,q) have the same probability of appearing in the sphere triangulation, and taking limits gives

νr(G0A)=νs(G0A).\nu_{r}(G_{0}^{A})=\nu_{s}(G_{0}^{A}).

Because sets of the form G0AG_{0}^{A} generate the σ\sigma-algebra of transversals, νr=νs\nu_{r}=\nu_{s}, and ν\nu is invariant.

As proven in Moore and Schochet [12] (Theorem 6.30), the invariance of ν\nu implies the existence of a trace on W(G)W^{*}(G), which we will denote τ\tau, given by

τ(H)=Tr(H)𝑑ν(),\tau(H)=\int\mathrm{Tr}(H_{\ell})\,d\nu(\ell),

where HH_{\ell} is the restriction of HH to the leaf \ell. ∎

Theorem 6.

Let u:G0u:G\to\mathbb{R}_{\geq 0} such that, for each xx, u|Gxu|_{G_{x}} is compactly supported and has total mass 1. Then

τ(H)=Tr(u(x,)(HGx))𝑑ν(x),\tau(H)=\int\mathrm{Tr}(u(x,*)(H_{G_{x}}))\,d\nu(x),

independently of uu.

Proof.

We note that

τ(H)\displaystyle\tau(H) =\displaystyle= Tr((HGy)(y))𝑑ν(y)\displaystyle\int\mathrm{Tr}((H_{G^{y}})(y))\,d\nu(y)
=\displaystyle= u(x,y)Tr((HGy)(y)dλy(x))𝑑ν(y)\displaystyle\int\int u(x,y)\mathrm{Tr}((H_{G^{y}})(y)\,d\lambda^{y}(x))\,d\nu(y)
=\displaystyle= u(x,y)Tr((HGx)(y)dλx(y))𝑑ν(x)\displaystyle\int\int u(x,y)\mathrm{Tr}((H_{G_{x}})(y)\,d\lambda_{x}(y))\,d\nu(x)
=\displaystyle= Tr(u(x,y)(HGx)(y)dλx(y))𝑑ν(x)\displaystyle\int\int\mathrm{Tr}(u(x,y)(H_{G_{x}})(y)\,d\lambda_{x}(y))\,d\nu(x)
=\displaystyle= Tr(u(x,)(HGx))𝑑ν(x).\displaystyle\int\mathrm{Tr}(u(x,*)(H_{G_{x}}))\,d\nu(x).

that is, τ\tau can also be given by Tr(u(x,)(HGx))𝑑ν(x)\int\mathrm{Tr}(u(x,*)(H_{G_{x}}))\,d\nu(x), independently of uu. ∎

We can also define, in an analogous manner, a trace (also denoted τ\tau) on W(G^)W^{*}(\hat{G}). In this case, the construction can be made simpler and more explicit.

Definition 19 (Trace on W(G^)W^{*}(\hat{G})).

For HW(G^)H\in W^{*}(\hat{G}), let

τ(H)=H(δx)(x)𝑑ν(x).\tau(H)=\int H(\delta_{x})(x)\,d\nu(x).
Theorem 7.

The function τ\tau is a trace with total mass 1.

Proof.

We note that

τ(HJ)\displaystyle\tau(HJ) =\displaystyle= HJ(δx)(x)𝑑ν(x)\displaystyle\int HJ(\delta_{x})(x)\,d\nu(x)
=\displaystyle= J(δx)(y)H(δy)(x)𝑑λx(y)𝑑ν(x)\displaystyle\int\int J(\delta_{x})(y)H(\delta_{y})(x)\,d\lambda_{x}(y)\,d\nu(x)
=\displaystyle= H(δy)(x)J(δx)(y)𝑑λy(x)𝑑ν(y)\displaystyle\int\int H(\delta_{y})(x)J(\delta_{x})(y)\,d\lambda_{y}(x)\,d\nu(y)
=\displaystyle= JH(δy)(y)𝑑ν(y)\displaystyle\int JH(\delta_{y})(y)\,d\nu(y)
=\displaystyle= τ(JH),\displaystyle\tau(JH),

and hence τ\tau satisfies the definition of a trace. That τ(1)=1\tau(1)=1 follows from the definition. ∎

We can now classify W(G^)W^{*}(\hat{G}), subject to the conditions that we used to guarantee amenability and ergodicity.

Theorem 8.

Suppose that ν\nu is such that, for a.e. xx, GxG^{x} is recurrent and νx=ν\nu^{x}=\nu. If ν(x)>0\nu(x)>0 for some (equivalently, a.e.) xx, then W(G^)W^{*}(\hat{G}) is a hyperfinite type I|G^0|I_{|\hat{G}_{0}|} factor. Otherwise, W(G^)W^{*}(\hat{G}) is a hyperfinite type II1II_{1} factor.

Proof.

Since G^\hat{G} is amenable, W(G^)W^{*}(\hat{G}) is hyperfinite by Delaroche and Renault [1]. Since G^\hat{G} is ergodic, W(G^)W^{*}(\hat{G}) is a factor by corollary 4. The type of W(G^)W^{*}(\hat{G}) follows by considering the possible values of τ\tau. ∎

Next we introduce a class of operators that will be particularly well-behaved.

Definition 20 (Finite hopping range).

Suppose that HW(G^)H\in W^{*}(\hat{G}) and there exists a radius rr such that HGx(δx)H_{G^{x}}(\delta_{x}) is supported in Br(x)B_{r}(x) for all xx, and Br(x)Br(y)B_{r}(x)\cong B_{r}(y) implies HGx(δx)=HGy(δy)H_{G^{x}}(\delta_{x})=H_{G^{y}}(\delta_{y}) where we identify Br(x)B_{r}(x) with Br(y)B_{r}(y). Then we say that HH has hopping range rr.

Similarly, suppose that HW(G)H\in W^{*}(G) and there exists a radius r>1r>1 with HGx(δx)H_{G^{x}}(\delta_{x}) supported in Br(x)B_{r}(x) for all xx. Suppose that, for all xx and yy with Br(x)Br(y)B_{r}(x)\cong B_{r}(y) and ff supported on B1(x)B_{1}(x), then HGx(f)=HGy(f)H_{G^{x}}(f)=H_{G^{y}}(f) (where we define ff on GyG^{y} via the isomorphism between GxG^{x} and GyG^{y}). Then we say that HH has hopping range rr.

Theorem 9.

If νx=ν\nu^{x}=\nu for almost all xx and HW(G^)H\in W^{*}(\hat{G}) has finite hopping range, then τ(H)\tau(H) is the limit of the leafwise trace averaged along a random walk on an almost arbitrary leaf; that is,

τ(H)=limnTr(νnx(GA)HGxA)\tau(H)=\lim_{n\to\infty}\mathrm{Tr}(\nu^{x}_{n}(G^{A})H_{G^{x_{A}}})

for almost all xx.

Proof.

Since HH has finite hopping range, H(δx)(x)H(\delta_{x})(x) is constant on GAG^{A} for AA a patch of radius at least rr. It follows that

τ(H)=HGxA(δxA)(xA)𝑑ν(GA),\tau(H)=\int H_{G^{x_{A}}}(\delta_{x_{A}})(x_{A})\,d\nu(G^{A}),

where the integral is over all patches AA of radius rr, and xAx_{A} is an arbitrary element of GAG^{A}. (If 𝒟\mathcal{D} is discrete, this integral is just a sum.) If xx is chosen such that νx=ν\nu^{x}=\nu, we can write

τ(H)\displaystyle\tau(H) =\displaystyle= HGxA(δxA)(xA)𝑑νx(GA)\displaystyle\int H_{G^{x_{A}}}(\delta_{x_{A}})(x_{A})\,d\nu^{x}(G^{A})
=\displaystyle= limnHGxA(δxA)(xA)𝑑νnx(GA)\displaystyle\lim_{n\to\infty}\int H_{G^{x_{A}}}(\delta_{x_{A}})(x_{A})\,d\nu^{x}_{n}(G^{A})
=\displaystyle= limnTr(νnx(GA)HGxA),\displaystyle\lim_{n\to\infty}\mathrm{Tr}(\nu^{x}_{n}(G^{A})H_{G^{x_{A}}}),

which is true for almost all xx. ∎

This argument can also be applied to any other sequence of measures that converges to ν\nu. In particular, we have this theorem, whose proof is analogous to Theorem 9:

Theorem 10.

If ν=limnνn\nu=\lim_{n\to\infty}\nu_{n} and HW(G^)H\in W^{*}(\hat{G}) has finite hopping range, then

τ(H)=limnTr(νn(GA)HGxA).\tau(H)=\lim_{n\to\infty}\mathrm{Tr}(\nu_{n}(G^{A})H_{G^{x_{A}}}).

5.3. Density of states

Definition 21 (Density of states).

For HW(G^)H\in W^{*}(\hat{G}) and ϕCc()\phi\in C_{c}(\mathbb{R}), define κH(ϕ)=τ(ϕ(H))\kappa_{H}(\phi)=\tau(\phi(H)). We call κH\kappa_{H} the density of states of HH.

Theorem 11.

If HH has finite hopping range, and νn\nu^{\prime}_{n} is a sequence of measures converging to ν\nu, then we can obtain the density of states κH(ϕ)\kappa_{H}(\phi) as a limit of the leafwise local trace Tr(ϕ(H))\mathrm{Tr}(\phi(H)) averaged with respect to νn\nu^{\prime}_{n}; that is,

κH(ϕ)=limnTr(νn(G^A)ϕ(H)).\kappa_{H}(\phi)=\lim_{n\to\infty}\mathrm{Tr}\left(\nu^{\prime}_{n}(\hat{G}^{A})\phi(H)\right).
Proof.

Given any continuous ϕ:\phi:\mathbb{R}\to\mathbb{R} and ϵ>0\epsilon>0, there exists a Weierstrass polynomial pp such that |ϕ(t)p(t)|<ϵ/3|\phi(t)-p(t)|<\epsilon/3 for |t|<H|t|<\|H\|. Let ϕ^=ϕp\hat{\phi}=\phi-p. Now p(H)p(H) is a polynomial of HH and therefore also has finite hopping range. We can therefore write

κH(p)\displaystyle\kappa_{H}(p) =\displaystyle= τ(p(H))\displaystyle\tau(p(H))
=\displaystyle= limnTr(νnx(G^A)p(H)G^xA).\displaystyle\lim_{n\to\infty}\mathrm{Tr}\left(\nu^{x}_{n}(\hat{G}^{A})p(H)_{\hat{G}^{x_{A}}}\right).

for almost all xx. In other words, for sufficiently large nn and almost all xx, we have

|κH(p)Tr(νn(G^A)p(H)G^xA)|<ϵ3.\left|\kappa_{H}(p)-\mathrm{Tr}\left(\nu^{\prime}_{n}(\hat{G}^{A})p(H)_{\hat{G}^{x_{A}}}\right)\right|<\frac{\epsilon}{3}.

Furthermore,

|κH(ϕ)κH(p)|\displaystyle\left|\kappa_{H}(\phi)-\kappa_{H}(p)\right| =\displaystyle= τ(ϕ(H))τ(p(H))\displaystyle\|\tau(\phi(H))-\tau(p(H))\|
=\displaystyle= |τ(ϕ^)|\displaystyle|\tau(\hat{\phi})|
\displaystyle\leq ϵ3\displaystyle\frac{\epsilon}{3}

and

|Tr(νn(G^A)ϕ(H)G^xA)Tr(νn(G^A)p(H)G^xA)|\displaystyle\left|\mathrm{Tr}\left(\nu^{\prime}_{n}(\hat{G}^{A})\phi(H)_{\hat{G}^{x_{A}}}\right)-\mathrm{Tr}\left(\nu^{\prime}_{n}(\hat{G}^{A})p(H)_{\hat{G}^{x_{A}}}\right)\right| =\displaystyle= |Tr(νn(G^A)ϕ^(H)G^xA)|\displaystyle\left|\mathrm{Tr}\left(\nu^{\prime}_{n}(\hat{G}^{A})\hat{\phi}(H)_{\hat{G}^{x_{A}}}\right)\right|
\displaystyle\leq |Tr(νn(G^A)ϵ/3)|\displaystyle\left|\mathrm{Tr}\left(\nu^{\prime}_{n}(\hat{G}^{A})\epsilon/3\right)\right|
\displaystyle\leq ϵ3,\displaystyle\frac{\epsilon}{3},

and we conclude that

|κH(ϕ)Tr(νn(G^A)ϕ(H)G^xA)|ϵ.\left|\kappa_{H}(\phi)-\mathrm{Tr}\left(\nu^{\prime}_{n}(\hat{G}^{A})\phi(H)_{\hat{G}^{x_{A}}}\right)\right|\leq\epsilon.

It follows that

κH(ϕ)=limnTr(νn(G^A)ϕ(H)G^xA).\kappa_{H}(\phi)=\lim_{n\to\infty}\mathrm{Tr}\left(\nu^{\prime}_{n}(\hat{G}^{A})\phi(H)_{\hat{G}^{x_{A}}}\right).

Two natural choices for the νn\nu^{\prime}_{n} are the measures νn\nu_{n} on spheres and the measures νnx\nu^{x}_{n} obtained from a random walk on a generic leaf (assuming GG is ergodic). Hence we have the following corollaries:

Corollary 5.

If HH has finite hopping range and GG is ergodic, then we can obtain κH(ϕ)\kappa_{H}(\phi) as a limit of Tr(ϕ(H))\mathrm{Tr}(\phi(H)) averaged along a random walk on an almost arbitrary leaf; that is,

κH(ϕ)=limnTr(νnx(G^A)ϕ(H)G^xA)\kappa_{H}(\phi)=\lim_{n\to\infty}\mathrm{Tr}\left(\nu^{x}_{n}(\hat{G}^{A})\phi(H)_{\hat{G}^{x_{A}}}\right)

for almost all xx.

Corollary 6.

If HH has finite hopping range, then we can obtain κH(ϕ)\kappa_{H}(\phi) as a limit of Tr(ϕ(H))\mathrm{Tr}(\phi(H)) averaged on sphere triangulations; that is,

κH(ϕ)=limnTr(νn(G^A)ϕ(H).)\kappa_{H}(\phi)=\lim_{n\to\infty}\mathrm{Tr}\left(\nu_{n}(\hat{G}^{A})\phi(H).\right)

The latter provides a convenient means of approximating κH\kappa_{H}, since Trϕ(H)\mathrm{Tr}\phi(H) on a finite sphere triangulation can be calculated by computing eigenvalues. We also wish to extend this result to some operators that do not have finite hopping range. Recall that DL2(G^)D\in L^{2}(\hat{G}) denotes the diagonal function.

Lemma 10.

We have

HD,D=τ(H).\langle HD,D\rangle=\tau(H).
Proof.

We have

HD,D\displaystyle\langle HD,D\rangle =\displaystyle= (HD)(x,y)D(x,y)𝑑λx(y)𝑑μ(x)\displaystyle\int\int(HD)(x,y)D(x,y)\,d\lambda_{x}(y)\,d\mu(x)
=\displaystyle= (HD)(x,x)𝑑μ(x)\displaystyle\int(HD)(x,x)\,d\mu(x)
=\displaystyle= H(δx)(x)𝑑μ(x),\displaystyle\int H(\delta_{x})(x)\,d\mu(x),

which is precisely our definition of τ(H)\tau(H) on G^\hat{G}. ∎

Corollary 7.

Suppose that Hn,HW(G^)H_{n},H\in W^{*}(\hat{G}) such that HnHH_{n}\to H weakly; that is,

Hnx,y=Hx,y\langle H_{n}x,y\rangle=\langle Hx,y\rangle

for all x,yL2(G^)x,y\in L^{2}(\hat{G}). Then τ(Hn)τ(H)\tau(H_{n})\to\tau(H).

Lemma 11.

Suppose that GG is ergodic; that Hn,HW(G^)H_{n},H\in W^{*}(\hat{G}) such that HnHH_{n}\to H in norm; and that ϕCc()\phi\in C_{c}(\mathbb{R}). Then

κHn(ϕ)κH(ϕ).\kappa_{H_{n}}(\phi)\to\kappa_{H}(\phi).
Proof.

We first claim that ϕ(Hn)ϕ(H)\phi(H_{n})\to\phi(H) in norm. Suppose that ϵ>0\epsilon>0. The hypotheses imply HnH_{n} is bounded, so there exists MM with |Hn|<M|H_{n}|<M for all nn and |H|<M|H|<M. We can choose a Weierstrass polynomial pp with |pϕ|<ϵ/3M|p-\phi|_{\infty}<\epsilon/3M. Furthermore, since HnHH_{n}\to H, for sufficiently large nn we have |p(Hn)p(H)|<ϵ/3|p(H_{n})-p(H)|<\epsilon/3, and

|ϕ(Hn)ϕ(H)|\displaystyle|\phi(H_{n})-\phi(H)| \displaystyle\leq |ϕ(Hn)p(Hn)|+|p(Hn)p(H)|+|p(H)ϕ(H)|\displaystyle|\phi(H_{n})-p(H_{n})|+|p(H_{n})-p(H)|+|p(H)-\phi(H)|
\displaystyle\leq ϵ.\displaystyle\epsilon.

Since norm convergence is stronger than weak convergence, it follows from the above corollary and the definition of κ\kappa that κHn(ϕ)κH(ϕ)\kappa_{H_{n}}(\phi)\to\kappa_{H}(\phi). ∎

Corollary 8.

Let HmH_{m} and HH be operators such that HmH_{m} has finite hopping range for each mm, HmHH_{m}\to H in norm, and νnν\nu^{\prime}_{n}\to\nu. Then

κH(ϕ)=limmlimnTr(νn(G^A)ϕ(Hm)G^xA).\kappa_{H}(\phi)=\lim_{m\to\infty}\lim_{n\to\infty}\mathrm{Tr}\left(\nu^{\prime}_{n}(\hat{G}^{A})\phi(H_{m})_{\hat{G}^{x_{A}}}\right).

5.4. Jumps of the IDS

One of the main results of Lenz and Veselić [11] relates compactly supported eigenfunctions to jumps of the IDS. We now have the framework to prove a similar result in our setting.

Theorem 12.

Suppose that 𝒟\mathcal{D} is discrete, G^\hat{G} is ergodic, and that HW(G^)H\in W^{*}(\hat{G}) is GG-invariant and has finite hopping range. Suppose also that tt\in\mathbb{R}. The following are equivalent:

  1. (1)

    The density of states κH(t)>0\kappa_{H}(t)>0.

  2. (2)

    For some xx, H|G^xH|_{\hat{G}^{x}} has an eigenfunction with eigenvalue tt supported on some finite patch AA with ν(GA)>0\nu(G^{A})>0.

  3. (3)

    For almost all xx, ker(HG^xλI)\ker(H_{\hat{G}^{x}}-\lambda I) is nontrivial and spanned by compactly supported eigenfunctions.

Proof.

The implication (3)(2)(3)\Rightarrow(2) is trivial. To prove (2)(1)(2)\Rightarrow(1), suppose that (2) holds. Let ff be a compactly supported eigenfunction on GxG^{x} with eigenvalue tt. Since HH has finite hopping range, there exists a radius rr such that Hf|G^x=Hf|G^yHf|_{\hat{G}^{x}}=Hf|_{\hat{G}^{y}} for any yG^0Br(x)y\in\hat{G}_{0}^{B_{r}(x)}.

Define u(x,y)=χBr(x)(y)|Br(x)|u(x,y)=\frac{\chi_{B_{r}(x)}(y)}{|B_{r}(x)|}. For each yG^0Br(x)y\in\hat{G}_{0}^{B_{r}(x)}, HH has ff as an eigenfunction, and hence Tr(H|Ay)>0\mathrm{Tr}(H|_{A_{y}})>0. Because G^0Br(x)\hat{G}_{0}^{B_{r}(x)} has positive measure, (1) follows. It only remains to prove (1)(3)(1)\Rightarrow(3).

For nn\in\mathbb{N}, let AnA_{n} denote the nnth interior of AA; that is, the set of xAx\in A with Bn(x)AB_{n}(x)\subset A.

Lemma 12.

Let PW(G^)P\in W^{*}(\hat{G}) with P>0P>0 and let R>0R>0. Then there exists a finite patch AA such that ran(χAP|G^A)2(G^0AR){0}\mathrm{ran}(\chi_{A}P|_{\hat{G}^{A}})\cap\ell^{2}(\hat{G}_{0}^{A_{R}})\not=\{0\}.

Proof.

Since P>0P>0, τ(P)>0\tau(P)>0. Because PP is bounded, we may assume without loss of generality that |P|=1|P|=1. By amenability and ergodicity, for almost every xx, G^x\hat{G}^{x} has a Følner exhaustion Λn\Lambda_{n}, and limnTr(χΛnP|Λn)|Λn|=τ(P)\lim_{n\to\infty}\frac{\mathrm{Tr}(\chi_{\Lambda_{n}}P|_{\Lambda_{n}})}{|\Lambda_{n}|}=\tau(P). Choose xx such that this is true. For nn sufficiently large,

Tr(χΛnP|Λn)12τ(P)|Λn|\mathrm{Tr}(\chi_{\Lambda_{n}}P|_{\Lambda_{n}})\geq\frac{1}{2}\tau(P)|\Lambda_{n}|

and, since Λn\Lambda_{n} is a Følner exhaustion,

|ΛnΛn,R|<12τ(P)|Λn|.|\Lambda_{n}\setminus\Lambda_{n,R}|<\frac{1}{2}\tau(P)|\Lambda_{n}|.

This implies

|ΛnΛn,R|\displaystyle|\Lambda_{n}\setminus\Lambda_{n,R}| <\displaystyle< Tr(χΛnP|Λn)\displaystyle\mathrm{Tr}(\chi_{\Lambda_{n}}P|_{\Lambda_{n}})
\displaystyle\leq dim(ran(χΛnP|Λn));\displaystyle\dim(\mathrm{ran}(\chi_{\Lambda_{n}}P|_{\Lambda_{n}}));

that is,

dim(ran(χΛnP|Λn))\displaystyle\dim(\mathrm{ran}(\chi_{\Lambda_{n}}P_{|}{\Lambda_{n}})) \displaystyle\geq ΛnΛn,R\displaystyle\|\Lambda_{n}\|-\|\Lambda_{n,R}\|
=\displaystyle= dim(2(Λn))dim(2(Λn,R)),\displaystyle\dim(\ell^{2}(\Lambda_{n}))-\dim(\ell^{2}(\Lambda_{n,R})),

implying that

ran(χΛnP|Λn)2(Λn,R){0}.\mathrm{ran}(\chi_{\Lambda_{n}}P|_{\Lambda_{n}})\cap\ell^{2}(\Lambda_{n,R})\not=\{0\}.

Lemma 13.

Let HW(G^)H\in W^{*}(\hat{G}) have finite hopping range and let λ\lambda\in\mathbb{R}. Define P=EH({λ})P=E_{H}(\{\lambda\}) and let PcP_{c} be the projection of PP onto the subspace generated by compactly supported eigenfunctions of HH. Then P=PcP=P_{c}.

Proof.

Suppose not. Define Q=PPcQ=P-P_{c}. Then Q>0Q>0. Define RR to be twice the hopping range of HH and apply the previous lemma, so that AA is a finite patch and we can choose a nonzero f2(G^0A)f\in\ell^{2}(\hat{G}_{0}^{A}) with χAQ(f)2(G^0AR){0}\chi_{A}Q(f)\in\ell^{2}(\hat{G}_{0}^{A_{R}})\setminus\{0\}. We note that

λχAQ(f)\displaystyle\lambda\chi_{A}Q(f) =\displaystyle= χAλQ(f)\displaystyle\chi_{A}\lambda Q(f)
=\displaystyle= χAHQ(f)\displaystyle\chi_{A}HQ(f)
=\displaystyle= χA(HχAQ(f))+χA(HχAcQ(f)).\displaystyle\chi_{A}(H\chi_{A}Q(f))+\chi_{A}(H\chi_{A^{c}}Q(f)).

Since χAQ(f)2(G^0AR)\chi_{A}Q(f)\in\ell^{2}(\hat{G}_{0}^{A_{R}}), HχAQ(f)2(G^0AR/2)H\chi_{A}Q(f)\in\ell^{2}(\hat{G}_{0}^{A_{R/2}}) since HH has hopping range R/2R/2. Again using the hopping range of HH, HχAcQ(f)2(G^0AR/2c)H\chi_{A^{c}}Q(f)\in\ell^{2}(\hat{G}_{0}^{A_{R/2}^{c}}). But λχAQ(f)2(G^0AR)\lambda\chi_{A}Q(f)\in\ell^{2}(\hat{G}_{0}^{A_{R}}), and hence the second term is 0, giving us

λχAQ(f)\displaystyle\lambda\chi_{A}Q(f) =\displaystyle= χA(HχAQ(f))\displaystyle\chi_{A}(H\chi_{A}Q(f))
=\displaystyle= HχAQ(f).\displaystyle H\chi_{A}Q(f).

This means that χAQ(f)\chi_{A}Q(f) is a compactly supported eigenfunction of HH, contradicting the definition of QQ. ∎

We can now prove (1)(3)(1)\Rightarrow(3). If (1) is true, then EH({λ})E_{H}(\{\lambda\}) is nontrivial, and hence ker(HG^xλI)\ker(H_{\hat{G}^{x}}-\lambda I) is nontrivial. The previous lemma implies that ker(HG^xλI)\ker(H_{\hat{G}^{x}}-\lambda I) is spanned by compactly supported eigenfunctions. ∎

5.5. Approximate eigenfunctions

Our goal is to study some properties of operators on G^\hat{G} by examining them on sphere triangulations. One way to do so is by relating a property to another property that can be described in terms of finite patches.

Definition 22 (Spectrally localized function).

Let HW(G^)H\in W^{*}(\hat{G}), xG0x\in G^{0}, and fL2(G^x)f\in L^{2}(\hat{G}^{x}). Let ϕ:[0,1]\phi:\mathbb{R}\to[0,1] and ξ,δ\xi,\delta\in\mathbb{R}. We say that ff is (H,ξ,δ,ϵ)(H,\xi,\delta,\epsilon)-spectrally localized if

(1χ[ξδ,ξ+δ])H(f)ϵf.\|(1-\chi_{[\xi-\delta,\xi+\delta]})H(f)\leq\epsilon\|f\|.\|
Definition 23 (Approximate eigenfunction).

Let HW(G^)H\in W^{*}(\hat{G}) and let ξ\xi\in\mathbb{C}. Let fL2(Gx){0}f\in L^{2}(G^{x})\setminus\{0\}. If (Hfξf)2<ζf2\|(Hf-\xi f)\|_{2}<\zeta\|f\|_{2}, we say that ff is a ζ\zeta-approximate eigenfunction for HH with eigenvalue ξ\xi supported on ArA_{r}.

Lemma 14.

Let HWG^H\in W^{*}{\hat{G}} be self-adjoint and let xG0x\in G^{0} and ξ\xi\in\mathbb{R}. Let δ>0\delta>0. Suppose that fL2(Gx)f\in L^{2}(G^{x}) such that fL2(Gx)=1\|f\|_{L^{2}(G^{x})}=1. Then:

  1. (1)

    If ff is an ϵδ\epsilon\delta-approximate eigenfunction, then ff is (H,ξ,δ,ϵ)(H,\xi,\delta,\epsilon)-spectrally localized.

  2. (2)

    If ff is (H,ξ,δ,ϵ)(H,\xi,\delta,\epsilon)-spectrally localized, then ff is a δ+Hϵ\delta+\|H\|\epsilon-approximate eigenfunction.

Proof.

(1) Assume without loss of generality that ξ=0\xi=0. Because (1χ[ξδ,ξ+δ])(t)t/δ(1-\chi_{[\xi-\delta,\xi+\delta]})(t)\leq t/\delta for all tt, we have

(1χ[ξδ,ξ+δ])HGx(f)\displaystyle\|(1-\chi_{[\xi-\delta,\xi+\delta]})H\|_{G^{x}}(f) \displaystyle\leq Hf/δ\displaystyle\|Hf\|/\delta
<\displaystyle< ϵ/δ.\displaystyle\epsilon/\delta.

(2) Again assume without loss of generality that ξ=0\xi=0. Because (1χ[ξδ,ξ+δ])(t)t/δ(1-\chi_{[\xi-\delta,\xi+\delta]})(t)\leq t/\delta for all tt, we have

Hf\displaystyle\|Hf\| =\displaystyle= H(ϕH(f)+(1χ[ξδ,ξ+δ])H(f))\displaystyle\|H(\phi H(f)+(1-\chi_{[\xi-\delta,\xi+\delta]})H(f))\|
\displaystyle\leq HϕH(f)+H(1χ[ξδ,ξ+δ])H(f)\displaystyle\|H\phi H(f)\|+\|H(1-\chi_{[\xi-\delta,\xi+\delta]})H(f)\|
\displaystyle\leq δ+Hϵ.\displaystyle\delta+\|H\|\epsilon.

Corollary 9.

Let HWG^H\in W^{*}{\hat{G}} be self-adjoint with finite hopping range rr and let ξ\xi\in\mathbb{R}. Let R>rR>r and δ>0\delta>0. Then

P(f|Gx has a BRrsupported ϵδ- approximate eigenfunction)\displaystyle P(f|_{G^{x}}\mbox{ has a }B_{R-r}-\mbox{supported }\epsilon\delta\mbox{- approximate eigenfunction})
\displaystyle\leq P(f|Gx has a BRrsupported (H,ξ,δ,ϵ) spectrally localized function)\displaystyle P(f|_{G^{x}}\mbox{ has a }B_{R-r}-\mbox{supported }(H,\xi,\delta,\epsilon)-\mbox{ spectrally localized function})
\displaystyle\leq P(f|Gx has a BRrsupported δ+Hϵ -approximate eigenfunction).\displaystyle P(f|_{G^{x}}\mbox{ has a }B_{R-r}-\mbox{supported }\delta+\|H\|\epsilon\mbox{ -approximate eigenfunction}).

Since the left and right sides of the inequality depend only on the patch densities of patches of radius RR, this gives a means by which we might estimate the frequency of compactly supported, spectrally localized functions. We might also compute limits as, for example, RR\to\infty, δ=CRu\delta=CR^{u}, and ϵ=DRv\epsilon=DR^{v} for some C,D>0C,D>0 and u,v<0u,v<0.

Corollary 10.

Let u,v0u,v\leq 0 and C,D>0C,D>0. Then

limRP(f|Gx has a BRrsupported CDRu+v- approximate eigenfunction)\displaystyle\lim_{R\to\infty}P(f|_{G^{x}}\mbox{ has a }B_{R-r}-\mbox{supported }CDR^{u+v}\mbox{- approximate eigenfunction})
\displaystyle\leq limRP(f|Gx has a BRrsupported (H,ξ,δ,DRv) spectrally localized function)\displaystyle\lim_{R\to\infty}P(f|_{G^{x}}\mbox{ has a }B_{R-r}-\mbox{supported }(H,\xi,\delta,DR^{v})-\mbox{ spectrally localized function})
\displaystyle\leq limRP(f|Gx has a BRrsupported CRu+HDRv -approximate eigenfunction).\displaystyle\lim_{R\to\infty}P(f|_{G^{x}}\mbox{ has a }B_{R-r}-\mbox{supported }CR^{u}+\|H\|DR^{v}\mbox{ -approximate eigenfunction}).

Let AA be a decorated discrete finite patch. Suppose that ff is an η\eta-approximate eigenfunction of HH with eigenvalue ξ\xi supported on ArA_{r}. Let δ,ι>0\delta,\iota>0 and ϕ=χ[ξδ,ξ+δ]\phi=\chi[\xi-\delta,\xi+\delta].

Corollary 11.

Suppose that BB is an exact decorated discrete finite patch containing kk disjoint copies of AA. On each copy of AA we have a copy of ff, which we can denote f1,,fkf_{1},\ldots,f_{k}. Then fif_{i} is (H,ϕ,η/δ)(H,\phi,\eta/\delta)-spectrally localized, and

fi,ϕH(fi)1ηδ.\langle f_{i},\phi H(f_{i})\rangle\geq 1-\frac{\eta}{\delta}.

6. Directionally invariant functions and the discrete Laplacian

Motivated by physics, we are interested in studying properties of the Laplacian and related operators. In order to work with the discrete spaces G^\hat{G} and G^0\hat{G}_{0}, we need to consider discrete analogues to these operators. Fabila Carrasco et al. [7] constructed a discrete magnetic Laplacian on graphs; we will adapt this construction to our setting.

Our choice of using tangent vectors to construct G^\hat{G} may seem inconvenient when constructing operators. However, the following mechanism allows us to consider operators on functions that essentially ignore the directions of tangent vectors.

Definition 24 (Directionally invariant function).

A function fLp(G^)f\in L^{p}(\hat{G}) is directionally invariant if f(x1,x2,y1,y2)=f(x1,x3,y1,y3)f(x_{1},x_{2},y_{1},y_{2})=f(x_{1},x_{3},y_{1},y_{3}) for all xk,ykx_{k},y_{k} such that (x1,x2,y1,y2),(x1,x3,y1,y3)G^(x_{1},x_{2},y_{1},y_{2}),(x_{1},x_{3},y_{1},y_{3})\in\hat{G}. We denote the subspace of directionally invariant functions by LDIp(G^)L^{p}_{DI}(\hat{G}).

Lemma 15.

If ff and gg are directionally invariant functions such that fgf*g is defined, then fgf*g is directionally invariant.

Proof.

We have

fg(x1,x2,z1,z2)\displaystyle f*g(x_{1},x_{2},z_{1},z_{2}) =\displaystyle= f(x1,x2,y1,y2)g(y1,y2,z1,z2)𝑑λ(x1,x2)(y1,y2)\displaystyle\int f(x_{1},x_{2},y_{1},y_{2})g(y_{1},y_{2},z_{1},z_{2})\,d\lambda_{(x_{1},x_{2})}(y_{1},y_{2})
=\displaystyle= f(x1,x3,y1,y2)g(y1,y2,z1,z3)𝑑λ(x1,x3)(y1,y2)\displaystyle\int f(x_{1},x_{3},y_{1},y_{2})g(y_{1},y_{2},z_{1},z_{3})\,d\lambda_{(x_{1},x_{3})}(y_{1},y_{2})
=\displaystyle= fg(x1,x3,z1,z3).\displaystyle f*g(x_{1},x_{3},z_{1},z_{3}).

Definition 25 (Directional symmetrization of a function).

Let fL2(G^)f\in L^{2}(\hat{G}). Define N(x)N(x) to be the set of vertices neighboring a vertex xx. The directional symmetrization of ff is

f~(x1,x2,y1,y2)=1|N(x1)||N(y1)|y3N(y1)x3N(x1)f(x1,x3,y1,y3).\tilde{f}(x_{1},x_{2},y_{1},y_{2})=\frac{1}{|N(x_{1})||N(y_{1})|}\sum_{\stackrel{{\scriptstyle x_{3}\in N(x_{1})}}{{y_{3}\in N(y_{1})}}}f(x_{1},x_{3},y_{1},y_{3}).

By construction, f~\tilde{f} is directionally invariant, and f~=f\tilde{f}=f when ff is directionally invariant.

Definition 26 (Extension of an operator on directionally invariant functions).

Let HB(LDI2(G^))H\in B(L^{2}_{DI}(\hat{G})). Define

H~(f)=H(f~).\tilde{H}(f)=H(\tilde{f}).

Then H~\tilde{H} is an operator on L2(G^)L^{2}(\hat{G}) that agrees with HH on LDI2(G^)L^{2}_{DI}(\hat{G}) (in particular, eigenfunctions of HH are also eigenfunctions of H~\tilde{H}).

We will construct the discrete Laplacian as an operator on directionally invariant functions.

Definition 27 (The discrete Laplacian with a potential and magnetic field).

Let 𝒟=4\mathcal{D}=\mathbb{R}^{4} and denote the components D(x1,x2)=(w(x1),w¯(x1,x2),V(x1),α(x1,x2))D(x_{1},x_{2})=(w(x_{1}),\bar{w}(x_{1},x_{2}),V(x_{1}),\alpha(x_{1},x_{2})). Furthermore, choose ν\nu such that, for almost all (x1,x2)(x_{1},x_{2}), ww and VV depend only on x1x_{1}, w¯\bar{w} is symmetric, and α\alpha is antisymmetric.

We think of ww as the vertex weight and w¯\bar{w} as the edge weight. We could construct ν\nu so that these weights are chosen either deterministically based on the surrounding geometry (one natural choice is w(x1)=deg(x1)w(x_{1})=\deg(x_{1}) and w¯(x1,x2)=1\bar{w}(x_{1},x_{2})=1) or randomly. We define

Δdisc,V,Bf(x,z)=1w(x1)y1N(x1)w¯(x1,y1)(eiα(x1,y1)f(y1,z1))+(V(x1)1)f(x1,z1).\Delta_{\mathrm{disc},V,B}f(x,z)=\frac{1}{w(x_{1})}\sum_{y_{1}\in N(x_{1})}\bar{w}(x_{1},y_{1})\left(e^{i\alpha(x_{1},y_{1})}f(y_{1},z_{1})\right)+(V(x_{1})-1)f(x_{1},z_{1}).

When α\alpha and VV are everywhere 0, this reduces to

Δdiscf(x1,z1)=1w(x1)y1N(x1)w¯(x1,y1)(f(y1,z1)f(x1,z1)).\Delta_{\mathrm{disc}}f(x_{1},z_{1})=\frac{1}{w(x_{1})}\sum_{y_{1}\in N(x_{1})}\bar{w}(x_{1},y_{1})\left(f(y_{1},z_{1})-f(x_{1},z_{1})\right).

In the case where α=V=0\alpha=V=0, w(x)=deg(x)w(x)=\deg(x), and w¯=1\bar{w}=1 everywhere, Figure 2 shows an eigenfunction of Δdisc\Delta_{\mathrm{disc}}.

1-11-1
Figure 2. Example of a compactly supported eigenfunction of Δdisc\Delta_{\mathrm{disc}} with eigenvalue 4/3-4/3. The function takes the value 0 on all unlabeled vertices.

7. Questions

  • Our amenability result used the result of Benjamini and Schramm [6] which requires a bound on the degrees of vertices. If we allow vertices to have unbounded degrees, can we still ensure amenability?

  • Do any of our results hold in higher dimensions? (Again, Benjamini and Schramm [6] would not apply.)

  • Do Conjectures 1, 2, and 3 hold?

  • In the case of a measure ν\nu derived from a generalized substitution tiling, can we prove results similar to those that have been proven for substitution tilings?

  • Can we determine any stronger results about “convergence” of eigenvectors; that is, results about properties of eigenvectors or near-eigenvectors of an operator on GG by examining analogous operators on spheres?

  • Can we say anything specific about Δdisc\Delta_{\mathrm{disc}} or any other interesting operator for any particular choice of ν\nu? In particular, can we prove any results analogous to those of Fabila Carrasco et al. [7]?

8. Acknowledgements

This paper is based on a dissertation written under the supervision of Eric Babson. The author is grateful to him and to Jerry Kaminker for their guidance and support.

References

  • [1] C. Anantharaman-Delaroche and J. Renault. Amenable groupoids. Number 36 in Monographie … de l’Enseignement mathématique. L’enseignement mathématique, Genève 2000.
  • [2] S. Beckus, J. Bellissard, and G. De Nittis. Spectral continuity for aperiodic quantum systems i. general theory. J. Funct. Anal., (11) 275 (2018), 2917–2977.
  • [3] S. Beckus and F. Pogorzelski. Delone dynamical systems and spectral convergence. Ergod. Theor. Dyn. Syst., (6) 40 (2020), 1510–1544.
  • [4] M. T. Benameur and V. Mathai. Gap-labelling conjecture with nonzero magnetic field. Adv. Math., 325 (2018), 116–164.
  • [5] M. T. Benameur and V. Mathai. Proof of the magnetic gap-labelling conjecture for principal solenoidal tori. J. Funct. Anal., (3) 278 (2020), 108323.
  • [6] I. Benjamini and O. Schramm. Recurrence of distributional limits of finite planar graphs. Electron. J. Probab., 6 (2001), 1–13.
  • [7] J. Fabila Carrasco, F. Lledó, and O. Post. Spectral gaps and discrete magnetic laplacians. Linear Algebra Appl., 547 (2017).
  • [8] A. Julien and J. Savinien. Transverse laplacians for substitution tilings. Commun. Math. Phys., 301 (2009), 1–29.
  • [9] J. Kaminker and I. Putnam. A proof of the gap labeling conjecture. Mich. Math. J., 51 (2003), 537–546.
  • [10] D. Lenz, N. Peyerimhoff, and I. Veselić. Groupoids, von neumann algebras and the integrated density of states. Math. Phys. Anal. Geom., 10 (2002), 1–41.
  • [11] D. Lenz and I. Veselić. Hamiltonians on discrete structures: Jumps of the integrated density of states and uniform convergence. Math. Z., 263 (2007), 813–835.
  • [12] C. C. Moore and C. L. Schochet. Global Analysis on Foliated Spaces. Mathematical Sciences Research Institute Publications. Cambridge University Press, 2 edition, Cambridge 2005.
  • [13] L. Sadun. Topology of Tiling Spaces. University lecture series. American Mathematical Society, Providence 2008.