Spaces of random plane triangulations and the density of states
1. Introduction
An important tool for studying tilings is the tiling space. [13] The set of all tilings of 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 , also called the hull of , is the closure in this space of the set of all translates of . Another, perhaps more intuitive, characterization of the hull of is the set of all tilings such that every finite patch of can be found, up to translation, in . 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 - 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 , 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:
and
where and are weight functions on edges and vertices, respectively, and and 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, is a groupoid consisting of twice-marked triangulations, is a transverse measure on , and is a von Neumann algebra that we will define on using the measure .
Theorem 8. Suppose that is such that, for a.e. , is recurrent and . If for some (equivalently, a.e.) , then is a hyperfinite type factor. Otherwise, is a hyperfinite type factor.
Theorem 11. If has finite hopping range, and is a sequence of measures converging to , then we can obtain the density of states as a limit of the leafwise local trace averaged with respect to ; that is,
Theorem 12. Suppose that is discrete, is ergodic, and that is -invariant and has finite hopping range. Suppose also that . The following are equivalent:
-
(1)
The density of states .
-
(2)
For some , has an eigenfunction with eigenvalue supported on some finite patch with .
-
(3)
For almost all , is nontrivial and spanned by compactly supported eigenfunctions.
2. Notation
- :
-
the th interior of the discrete decorated finite patch
- ):
-
the reduced -algebra associated with the groupoid or
- :
-
the distance on the decoration space
- :
-
the diagonal function on
- :
-
the decoration space
- :
-
the inclusion map of a decorated (discrete or continuous) finite patch into the triangulation
- :
-
the holonomy groupoid: the set of triangulations with two marked tangent vectors, up to automorphisms
- :
-
the set of units in the holonomy groupoid: the set of triangulations with a marked tangent vector, up to automorphisms
- :
-
the discrete holonomy groupoid, and its set of units
- :
-
the set of (discrete, continuous) triangulations containing the (marked, twice-marked) finite patch
- :
-
the set of neighbors of a vertex
- :
-
the discrete unit tangent bundle of
- :
-
the unit tangent bundle of
- :
-
the smooth geometric realization of
- :
-
the set of abstract simplicial complexes
- ):
-
the von Neumann algebra associated with the groupoid or
- :
-
the marked tangent vector on the (discrete or continuous) finite patch
- :
-
the tail and head of the discrete tangent vector
- :
-
the metric defined on
- :
-
the density of states associated with the operator
- :
-
the transverse measure on
- :
-
the trace on or
- :
-
the coordinate patch associated with the patch and ball
- :
-
the set of permissible decorations on the decorated finite patch
3. Triangulation spaces
3.1. The space of pointed triangulations
Definition 1 (Triangulation; geometric realization).
By a triangulation we mean an abstract simplicial complex of dimension 2. The sets of vertices, edges, directed edges, and faces of will be denoted , , and . The geometric realization of of is a metric space formed by assigning to each vertex , edge , and face a Euclidean simplex of the same dimension with marked points corresponding to vertices, respecting inclusion so that is identified with the corresponding point in if is a vertex of , and is identified with the corresponding segment in if is an edge of . The usual metric on is simply the Euclidean metric on each simplex, joined by shortest paths.
Fix an integer and a compact metric space . Let denote the metric on .
Definition 2 (Decorated triangulation).
Let be the set of all ordered pairs , where is a triangulation satisfying:
-
•
each vertex of has degree at most , and
-
•
the geometric realization of is a surface of genus 0,
and where . That is, a decorated triangulation consists of a triangulation with decorations . (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 will be regarded as properties of ; e.g., we will define . We require isomorphisms between triangulations to preserve decorations: is isomorphic to if and only if there is a simplicial isomorphism such that for all . Although we have defined decorations on edges, we could also speak of decorations on vertices, for example by considering triangulations where some components of 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 , and let denote its geometric realization. Denote by the degree of for each vertex of . For each directed edge of , map to by preserving the distance from 0 or , and subdividing the unit circle into equal intervals and mapping the th interval linearly to the angles on the th face counterclockwise from . These maps form charts of . Let be the Euclidean inner product on in the aforementioned chart, and let be the Euclidean inner product on the usual piecewise linear structure of . We observe that does not depend on the choice of the second vertex , and that is defined on all points except for vertices. Hence we can define a Riemannian metric
where is a smooth function with and for , and is the distance from the point to the nearest vertex.
Let denote the unit tangent bundle of , with the local product metric given by its structure as an -fiber bundle.
We note that the isometry class of a face depends only on the degrees of the vertices of .
For a triangulation , we will denote by the group of simplicial automorphisms of (which is trivial in most cases). The group acts naturally on , , and .
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 be a triangulation. A discrete tangent vector of is an ordered edge in . The discrete sphere bundle of , denoted , is the set of discrete tangent vectors of equipped with the metric . Thus two distinct discrete tangent vectors and are adjacent if and only if and are either equal or adjacent, and and are either equal or adjacent. We will often abbreviate, for example, to .
We can embed in by mapping to each the tangent vector at along the edge .
Let be the set of pairs where is an orbit of the action of on . Since automorphisms are rare, we will usually think of such an orbit as a single point and write or simply if the context is clear. When we define the holonomy groupoid later on, 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 . 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 and be triangulations with . If is an isometry such that, for every discrete tangent vector , is a discrete tangent vector in with , we say that is an -nearly decoration-preserving isometry (henceforth -NDPI).
Definition 6 (Asymmetric distance in ).
Suppose that . For each , , and an -NDPI, let be the distance from to . We define
where the infimum is over all , , and satisfying these conditions.
Lemma 1.
The function defined this way is nondegenerate; that is,
only if .
Proof.
If , then there exist and as in the definition with arbitrarily large and arbitrarily small and . By composing with automorphisms if necessary, we may assume without loss of generality that and that is arbitrarily small. Let be such that and . Each can be written as for some . For sufficiently large , we have . Then and . This means that
where the limit is taken pointwise, is a well-defined isometry from to that sends to . Furthermore, if is a discrete tangent vector in , then for every because some is -nearly decoration-preserving, and hence . ∎
Lemma 2.
The function defined this way satisfies the triangle inequality:
Proof.
We can always take to be and to be the map sending to . Hence
and we can restrict our attention to maps with .
If , and we have a pair of maps and , we can compose them as follows. By composing with an action of , we can assume without loss of generality that the distance from to is at most . We can then define
and
If is a discrete tangent vector, we have
so is -nearly decoration-preserving. Then
and
where the second inequality comes from the fact that and
for . ∎
Definition 7 (Distance in ).
We define
Theorem 1.
The function defined in this way is a metric.
Proof.
Since is nondegenerate and satisfies the triangle inequality, so does . Furthermore, is symmetric by definition. ∎
Definition 8 (The triangulation topology).
The triangulation topology is the topology induced by the metric .
3.3. Foliated structure
Next, we will give 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 with some model space (often, as in this case, a Cantor-like set). In this case our model space will be the discrete space 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 and . These will be used to define the foliated structure, but, more importantly, will be a basis of the -algebra upon which our transverse measure is defined.
Definition 9 (Decorated finite patch).
By a decorated discrete finite patch we mean an ordered triple where is a finite triangulation, is a marked discrete tangent vector of , and is a Borel subset of (identified with the space of functions from to ). Generally we will refer to such a patch simply as and write and . We similarly define a decorated continuous finite patch as an ordered triple where and are as above and . We call the set of permissible decorations on . If we say the finite patch is exact. Likewise, we define a twice-marked decorated (discrete or continuous) finite patch as an ordered quadruple where and are marked (discrete or continuous) tangent vectors.
Definition 10 (Sets of triangulations with a particular decorated finite patch).
Let be a decorated discrete finite patch. We define to be the set of triangulations with a marked discrete tangent vector such that there exists with the properties that:
-
•
is a simplicial embedding;
-
•
;
-
•
the function .
For any particular and , if such a exists, it is unique up to isomorphism (or choice of ). We analogously define if is a twice-marked decorated discrete finite patch, if is a decorated continuous finite patch, or if is a twice-marked decorated continuous finite patch.
Definition 11 (Coordinate patches of ).
Let be a decorated continuous finite patch. Let be any smooth, but not necessarily isometric, embedding of a region into . We define to by
and call it the coordinate patch associated with the patch and region .
We can think of such a patch as a triangulation that is fixed in the region and allowed to vary outside of and whose decorations are allowed to vary within , with a marked tangent vector that is allowed to move within the subset .
Lemma 3.
The space equipped with the coordinate patches is a foliated space.
Proof.
Since in the definition contains at most one point in any orbit of for any containing , the coordinate patch is one-to-one. It follows from the manifold structure of that is smooth. ∎
3.4. Holonomy groupoid
Our next objective is to construct , the holonomy groupoid of . 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 with a space of units , a diagonal map , an inversion map on , range and source maps , and an associative multiplication on pairs where . These must satisfy , , , and . We denote and .
One simple example of a groupoid is an equivalence relation , where iff , and we have , , , and . 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 where is a leaf and is an open set. Given a path from to , and transversals and through and , respectively, we cover by small open sets . If the are sufficiently small, any plaque in intersects a unique plaque in , so that we can start with the plaque in containing some and follow the sequence of intersecting plaques to a unique plaque in , which intersects a unique point in . Hence this results in a map from a neighborhood in to one in , realized by following a path from to sufficiently close to . The germ of this map is independent of the choice of , and two paths correspond to the same holonomy element if the germs of their corresponding maps on and are the same.
Lemma 4.
The holonomy groupoid consists of orbits of pairs of points , where for some .
Proof.
The leaves of correspond to with a triangulation. A plaque on a leaf of corresponds to a small open subset of .
Suppose that and is a path from to . We can choose a sufficiently large decorated continuous finite patch such that and is contained in the image of . Suppose additionally that are open subsets of such that the cover .
Suppose that is another triangulation with . Let ; define , , and analogously.
Then cover , which is a path from to . It follows that the holonomy element corresponding to sends to ; it depends on and but not on itself. This holds for any . Hence there is a unique holonomy element for each orbits of pairs of points on some triangulation . ∎
Definition 12 (The discrete holonomy groupoid and its space of units).
The holonomy groupoid contains a subgroupoid consisting of elements of the form where and are discrete tangent vectors in for some triangulation . We denote its set of units by .
If is trivial, as we expect to be true in many cases, we can simply think of an element of as a triangulation with two marked tangent vectors, and an element of as a triangulation with two marked discrete tangent vectors. In any case, we will often suppress the brackets and write such an element .
Our primary objects of interest will be and . However, we will often need to refer to and in order to use results about foliated spaces, as 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 as a Borel subset of its space of units 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 is such a transversal, we will simply think of our transverse measures as measures on .
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 converges to a transverse measure if for every decorated discrete finite patch .
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 be the subset of consisting of points on finite (sphere) triangulations, and be a sequence of measures on . We require that for . In other words, the restriction of each to a single sphere is a discrete uniform measure.
There are many possible choices for . A few examples may be instructive:
-
(1)
Consider the case where is trivial and is a uniform probability measure on the sphere triangulations with at most 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 converges. That is, let be an enumeration of decorated finite patches that includes, for every finite patch , a basis of the Borel -algebra on . Let . For each , choose a subsequence of such that for some (by compactness, such a subsequence must exist). Then the diagonal sequence must have for each . Henceforth we will assume that such a sequence is fixed, and denote the limit by .
Conjecture 1.
The sequence of measures in this construction converges.
-
(2)
Let and let be the sphere triangulation formed by identifying the boundaries of two balls of radius on a 2-dimensional triangular grid. Define to be the measure that decorates each edge of independently with 0 or 1 with equal probability and chooses a marked edge uniformly, and let be the limit of the . If is a finite patch that is not a subset of a 2-dimensional triangular grid, then since the identified boundaries on each have measures approaching 0. Hence is a measure on decorated triangular grids, and is essentially a discrete tiling space. Without the decorations almost every leaf would be periodic and would consist of a single point. However, with the decorations almost every leaf is aperiodic.
-
(3)
We can construct spheres via a process analogous to a substitution tiling. Let be a finite set. Then there are only finitely many possible decorated faces. Choose a map that assigns to each decorated face with an ordering on the vertices a finite patch with three distinct marked points such that and , , and are on the boundary of with no interior edges. Let denote the portion of the boundary from to that does not pass through , and define and analogously. We require that depend only on and and the analogous conditions hold for and . We can define on any triangulation by applying it to each face. We say that is nondegenerate if there exists such that has at least one interior face for every face . This implies that there exists such that the proportion of boundary faces to total faces in is at most . Then we can set to be any sphere triangulation, set , and choose to be the uniform measure on . Since is nondegenerate, almost all faces are in the interior (and eventually arbitrarily far from the boundary) of for some , and we can determine the limiting frequency of each patch from analyzing , as in tiling theory. We denote the limit by .
-
(4)
Let be the finite triangulation consisting of a single vertex of degree 7, its neighboring faces, and their vertices and edges. For , let be constructed from 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 : in black, in red, and in blue. We can see that each boundary vertex has degree 3 or 4. If denotes a vertex of degree 3 and denotes a vertex of degree 4, the boundary of is obtained from that of via the substitution
The largest eigenvalue of the matrix of this substitution is , and hence the number of vertices in the boundary is asymptotically proportional to
in particular, it grows exponentially in . Furthermore, since is irrational, the limit of this substitution is aperiodic.
Now let be the sphere triangulation formed by identifying the boundaries of two copies of . 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 to be the limit of the measures that are uniform on the . The proportion of vertices of degree 7 is asymptotically
meaning that vertices of degree 4 and 6 have positive measure.
It is not too hard to see that 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 for large , 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 with only degree 7 vertices are isomorphic, and their frequency is simply the limiting proportion of vertices that are a distance more than 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 and let be a sequence of functions on such that for each . Often we will consider a leaf that is recurrent in the sense of random walks and take to be the distribution of a random walk from after steps. For a decorated discrete finite patch , we define .
If exists, we call it the patch density of on (with respect to the exhaustion ). 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 by averaging along leaves. If has two properties, amenability and ergodicity, such averages will be well-behaved. Additionally, when we construct a von Neumann algebra on , 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 satisfying the following for -a.e. :
-
(1)
, and
-
(2)
for all , we have .
If almost all leaves had subexponential growth, we could simply take to be . This does not work in the case of, for example, , and so we resort to an argument using random walks.
Theorem 2.
If the groupoid is given a transverse measure such that is recurrent for a.e. x, then is amenable.
Proof.
For a triangulation, a vertex of , and a positive integer, let denote a simple random walk on (that is, on the 1-skeleton of ). To make our notation less cumbersome, we will use the shorthand to denote . Let be the distribution of after a uniformly random number of steps from ; that is,
Let .
Let be the first time that the walk hits - that is, the smallest positive integer such that . If is recurrent and is sufficiently large, then and . We will use to denote .
Now suppose that .
We have
Our next step is to decompose this sum into two parts: one with and the other with . (That is, either the random walk starting at hits before time , in which case the distribution of a random walk starting at is similar to one starting at , or it does not hit before time , which is unlikely because of our choice of .)
First, we will consider the sum where ; that is, the random walk starting at hits before time . If , we have
Now, if , then and is a random walk of steps starting at . Hence each term in the first sum is , and we are left with
Next, consider the sum with . We have
Hence
Exchanging and gives
and hence
and we have shown that is amenable. ∎
Corollary 1.
The groupoid with any transverse measure constructed as in Definition 14 is amenable.
Proof.
Using the result of Benjamini and Schramm [6], is recurrent for -a.e. . By the preceding lemma, it suffices to show that this implies is recurrent. Recall that each point is actually a discrete tangent vector, or directed edge, , and that two such distinct tangent vectors and are adjacent if and only if and are either equal or adjacent, and and are either equal or adjacent. We will use the flow theorem, which states that a graph is transient if and only if it has a flow with a single source, no sinks, and a finite energy
Suppose that , where is a tangent vector on the triangulation , is transient. Then there exists a flow with a single source and finite energy. Define a flow on by
We note that
and hence has a single source at and no sinks.
Furthermore,
Taking sums gives
from which we conclude that is transient. By contrapositive, if is recurrent, then is recurrent. ∎
The case of measures constructed using a leaf as in Definition 15 is less clear.
Conjecture 2.
If is recurrent and exists, then is recurrent for -a.e. y.
If this conjecture holds, it will also imply that is amenable when equipped with the transverse measure constructed subject to these conditions.
4.2. Ergodicity
Next we consider ergodicity. We will start by proving the following lemma:
Lemma 5.
Suppose that is chosen such that, for any decorated finite patch , for -almost every (that is, the patch density of any finite patch almost surely does not depend on ). For all , we have
Proof.
For each finite patch , let . If and are decorated finite patches and and , the claim holds since both sides equal . The claim in general follows from bilinearity and the fact that functions of the form span a dense subset of . ∎
Theorem 3.
Suppose that is chosen such that, for any decorated finite patch , for -almost every . Then equipped with the measure is ergodic.
Proof.
Suppose that is an invariant, real-valued function (that is, for any ) with . Setting in Lemma 5,
and hence -a.e. Since every such is zero, is ergodic. ∎
Conjecture 3.
The measure is ergodic.
5. Operator algebras
5.1. Construction
We are interested in the properties of operators on . To study these operators, we will consider operator algebras. Both the -algebra and the von Neumann algebra are generated using leafwise convolution operators.
Definition 16 (The reduced -algebra of ).
We construct as a -algebra operating on . To each continuous, compactly supported function on , and for each , we associate the operator on given by
for each . The closure of all under the norm
forms the -algebra associated to .
Definition 17 (The von Neumann algebras of and ).
We construct as a von Neumann algebra operating on . Using the transverse measure defined above, we can construct the spaces . For and , we define the operator
The weak closure of all such is the von Neumann algebra associated with . That is, belongs to if there exists a sequence such that for all .
We can repeat this construction using in place of . Since is a complete transversal of , we have
by Moore and Schochet [12] (Proposition 6.21).
The and von Neumann algebras are related as follows.
Lemma 6.
If is a continuous, compactly supported function, then
Proof.
For ,
∎
Corollary 2.
There exists an algebra homomorphism such that .
Proof.
Let . Choose a sequence of continuous, compactly supported functions with . Since is Cauchy, is also Cauchy by the preceding lemma, and hence converges. Let . To show that this is well-defined, let be another such sequence of functions. Then and the preceding lemma gives . That is an algebra homomorphism follows from the definition. ∎
For our results the most important algebra is . We will first characterize the operators in , then discuss its classification.
Definition 18 (The diagonal function).
Define by if and otherwise.
It is easily seen that and that is the identity of leafwise convolution; that is, .
Lemma 7.
Let and . Then .
Proof.
We have
∎
Theorem 4.
Let be a bounded linear operator on . The following are equivalent:
-
(1)
.
-
(2)
for every .
-
(3)
for some .
Proof.
If satisfies (2), let . For every , we have ; that is, .
If satisfies (3), we can approximate as a norm, hence weak, limit of compactly supported functions . For every ,
that is, is a weak limit of the .
Finally, suppose that satisfies (1). Then we can write as the weak limit of with compactly supported. Define . Since
weak convergence implies
Because was arbitrary, this implies . ∎
This result is convenient because it allows us to reduce questions about operators in to questions about convolutions of functions on . In the following discussion, we will identify a function with . In addition, if is a twice-marked decorated discrete finite patch we will write in place of .
Although the structure of itself is essentially symmetric with respect to its first and second argument, 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 , , , and for all . Then .
In other words, operators in preserve invariance with respect to the second argument. (They do not preserve invariance with respect to the first argument; consider the example where is a diagonal function taking on two different values.) We might think of operators in 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 .
Lemma 8.
Suppose that and . Define . Then
if and only if
for almost every .
Proof.
We have
and, similarly,
∎
Lemma 9.
Let . Then if and only if and for almost every with .
Proof.
Suppose that and for almost every with . Then for some . For every and , we have
and the preceding lemma implies ; i. e., .
Conversely, suppose that . For every , the preceding lemma gives
for almost all . This in particular holds when for some discrete decorated finite patch . We can choose a countable set of finite patches such that the generate the -algebra of . If , then there must exist some such that ; otherwise would be isomorphic to , and and would be two representatives of the same orbit of . Hence for almost all with , and we may assume that for some .
If is arbitrary, Lemma 9 gives
for almost all . By choosing a that is nowhere zero, this implies that for almost all ; that is, is constant on almost every leaf. ∎
Corollary 4.
The algebra is a factor if and only if is ergodic.
Proof.
If is ergodic and , the above lemma implies that and for almost every with ; that is, for some , and is constant on almost every leaf. Since is ergodic, this means that that is constant, and is a multiple of .
Conversely, if is not ergodic, there exists a function that is not constant but is constant on almost every leaf. Then , but is not a multiple of . ∎
5.2. Trace
Many of our results will involve traces on and . These traces are closely related to the transverse measure . First we will show:
Theorem 5.
The transverse measure is invariant; that is, where denotes the range map, denotes the source map, and
and
Proof.
Suppose that is a finite triangulation with marked tangent vectors and . For each ,
is the number of times appears in . This is 1 if is contained in a marked patch that looks like in , and 0 otherwise. Furthermore,
is simply the probability of the marked patch appearing in a randomly chosen marked triangulation in . Similarly, is the probability of appearing in a randomly chosen marked triangulation.
Suppose that a sphere triangulation contains distinct (though not necessarily disjoint) copies of . Each copy of must contain exactly copies of (that is, tangent vectors such that is isomorphic to ), and exactly copies of . Furthermore, no two distinct copies of can share a copy of (respectively, ), since knowing the location of specifies the location of all other faces of . Since the marked tangent vector is chosen uniformly, and have the same probability of appearing in the sphere triangulation, and taking limits gives
Because sets of the form generate the -algebra of transversals, , and is invariant.
As proven in Moore and Schochet [12] (Theorem 6.30), the invariance of implies the existence of a trace on , which we will denote , given by
where is the restriction of to the leaf . ∎
Theorem 6.
Let such that, for each , is compactly supported and has total mass 1. Then
independently of .
Proof.
We note that
that is, can also be given by , independently of . ∎
We can also define, in an analogous manner, a trace (also denoted ) on . In this case, the construction can be made simpler and more explicit.
Definition 19 (Trace on ).
For , let
Theorem 7.
The function is a trace with total mass 1.
Proof.
We note that
and hence satisfies the definition of a trace. That follows from the definition. ∎
We can now classify , subject to the conditions that we used to guarantee amenability and ergodicity.
Theorem 8.
Suppose that is such that, for a.e. , is recurrent and . If for some (equivalently, a.e.) , then is a hyperfinite type factor. Otherwise, is a hyperfinite type factor.
Proof.
Next we introduce a class of operators that will be particularly well-behaved.
Definition 20 (Finite hopping range).
Suppose that and there exists a radius such that is supported in for all , and implies where we identify with . Then we say that has hopping range .
Similarly, suppose that and there exists a radius with supported in for all . Suppose that, for all and with and supported on , then (where we define on via the isomorphism between and ). Then we say that has hopping range .
Theorem 9.
If for almost all and has finite hopping range, then is the limit of the leafwise trace averaged along a random walk on an almost arbitrary leaf; that is,
for almost all .
Proof.
Since has finite hopping range, is constant on for a patch of radius at least . It follows that
where the integral is over all patches of radius , and is an arbitrary element of . (If is discrete, this integral is just a sum.) If is chosen such that , we can write
which is true for almost all . ∎
This argument can also be applied to any other sequence of measures that converges to . In particular, we have this theorem, whose proof is analogous to Theorem 9:
Theorem 10.
If and has finite hopping range, then
5.3. Density of states
Definition 21 (Density of states).
For and , define . We call the density of states of .
Theorem 11.
If has finite hopping range, and is a sequence of measures converging to , then we can obtain the density of states as a limit of the leafwise local trace averaged with respect to ; that is,
Proof.
Given any continuous and , there exists a Weierstrass polynomial such that for . Let . Now is a polynomial of and therefore also has finite hopping range. We can therefore write
for almost all . In other words, for sufficiently large and almost all , we have
Furthermore,
and
and we conclude that
It follows that
∎
Two natural choices for the are the measures on spheres and the measures obtained from a random walk on a generic leaf (assuming is ergodic). Hence we have the following corollaries:
Corollary 5.
If has finite hopping range and is ergodic, then we can obtain as a limit of averaged along a random walk on an almost arbitrary leaf; that is,
for almost all .
Corollary 6.
If has finite hopping range, then we can obtain as a limit of averaged on sphere triangulations; that is,
The latter provides a convenient means of approximating , since 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 denotes the diagonal function.
Lemma 10.
We have
Proof.
We have
which is precisely our definition of on . ∎
Corollary 7.
Suppose that such that weakly; that is,
for all . Then .
Lemma 11.
Suppose that is ergodic; that such that in norm; and that . Then
Proof.
We first claim that in norm. Suppose that . The hypotheses imply is bounded, so there exists with for all and . We can choose a Weierstrass polynomial with . Furthermore, since , for sufficiently large we have , and
Since norm convergence is stronger than weak convergence, it follows from the above corollary and the definition of that . ∎
Corollary 8.
Let and be operators such that has finite hopping range for each , in norm, and . Then
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 is discrete, is ergodic, and that is -invariant and has finite hopping range. Suppose also that . The following are equivalent:
-
(1)
The density of states .
-
(2)
For some , has an eigenfunction with eigenvalue supported on some finite patch with .
-
(3)
For almost all , is nontrivial and spanned by compactly supported eigenfunctions.
Proof.
The implication is trivial. To prove , suppose that (2) holds. Let be a compactly supported eigenfunction on with eigenvalue . Since has finite hopping range, there exists a radius such that for any .
Define . For each , has as an eigenfunction, and hence . Because has positive measure, (1) follows. It only remains to prove .
For , let denote the th interior of ; that is, the set of with .
Lemma 12.
Let with and let . Then there exists a finite patch such that .
Proof.
Since , . Because is bounded, we may assume without loss of generality that . By amenability and ergodicity, for almost every , has a Følner exhaustion , and . Choose such that this is true. For sufficiently large,
and, since is a Følner exhaustion,
This implies
that is,
implying that
∎
Lemma 13.
Let have finite hopping range and let . Define and let be the projection of onto the subspace generated by compactly supported eigenfunctions of . Then .
Proof.
Suppose not. Define . Then . Define to be twice the hopping range of and apply the previous lemma, so that is a finite patch and we can choose a nonzero with . We note that
Since , since has hopping range . Again using the hopping range of , . But , and hence the second term is 0, giving us
This means that is a compactly supported eigenfunction of , contradicting the definition of . ∎
We can now prove . If (1) is true, then is nontrivial, and hence is nontrivial. The previous lemma implies that is spanned by compactly supported eigenfunctions. ∎
5.5. Approximate eigenfunctions
Our goal is to study some properties of operators on 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 , , and . Let and . We say that is -spectrally localized if
Definition 23 (Approximate eigenfunction).
Let and let . Let . If , we say that is a -approximate eigenfunction for with eigenvalue supported on .
Lemma 14.
Let be self-adjoint and let and . Let . Suppose that such that . Then:
-
(1)
If is an -approximate eigenfunction, then is -spectrally localized.
-
(2)
If is -spectrally localized, then is a -approximate eigenfunction.
Proof.
(1) Assume without loss of generality that . Because for all , we have
(2) Again assume without loss of generality that . Because for all , we have
∎
Corollary 9.
Let be self-adjoint with finite hopping range and let . Let and . Then
Since the left and right sides of the inequality depend only on the patch densities of patches of radius , 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, , , and for some and .
Corollary 10.
Let and . Then
Let be a decorated discrete finite patch. Suppose that is an -approximate eigenfunction of with eigenvalue supported on . Let and .
Corollary 11.
Suppose that is an exact decorated discrete finite patch containing disjoint copies of . On each copy of we have a copy of , which we can denote . Then is -spectrally localized, and
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 and , 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 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 is directionally invariant if for all such that . We denote the subspace of directionally invariant functions by .
Lemma 15.
If and are directionally invariant functions such that is defined, then is directionally invariant.
Proof.
We have
∎
Definition 25 (Directional symmetrization of a function).
Let . Define to be the set of vertices neighboring a vertex . The directional symmetrization of is
By construction, is directionally invariant, and when is directionally invariant.
Definition 26 (Extension of an operator on directionally invariant functions).
Let . Define
Then is an operator on that agrees with on (in particular, eigenfunctions of are also eigenfunctions of ).
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 and denote the components . Furthermore, choose such that, for almost all , and depend only on , is symmetric, and is antisymmetric.
We think of as the vertex weight and as the edge weight. We could construct so that these weights are chosen either deterministically based on the surrounding geometry (one natural choice is and ) or randomly. We define
When and are everywhere 0, this reduces to
In the case where , , and everywhere, Figure 2 shows an eigenfunction of .
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.)
- •
-
•
In the case of a measure 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 by examining analogous operators on spheres?
-
•
Can we say anything specific about or any other interesting operator for any particular choice of ? 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.