Action convergence of general hypergraphs and tensors
Abstract
Action convergence provides a limit theory for linear bounded operators where are potentially different probability spaces. This notion of convergence emerged in graph limits theory as it unifies and generalizes many notions of graph limits. We generalize the theory of action convergence to sequences of multi-linear bounded operators . Similarly to the linear case, we obtain that for a uniformly bounded (under an appropriate norm) sequence of multi-linear operators, there exists an action convergent subsequence. Additionally, we explain how to associate different types of multi-linear operators to a tensor and we study the different notions of convergence that we obtain for tensors and in particular for adjacency tensors of hypergraphs. We obtain several hypergraphs convergence notions and we link these with the hierarchy of notions of quasirandomness for hypergraph sequences. This convergence also covers sparse and inhomogeneous hypergraph sequences and it preserves many properties of adjacency tensors of hypergraphs. Moreover, we explain how to obtain a meaningful convergence for sequences of non-uniform hypergraphs and, therefore, also for simplicial complexes. Additionally, we highlight many connections with the theory of dense uniform hypergraph limits (hypergraphons) and we conjecture the equivalence of this theory with a modification of multi-linear action convergence.
Keywords: Graph limits, Hypergraphs, Action convergence, Tensors, Higher-order interactions Mathematics Subject Classification Number: 05C65
1 Introduction
In the last 20 years, the study of complex networks has permeated many areas of social and natural sciences. Important examples are computer, telecommunication, biological, cognitive, semantic and social networks. In particular, in all of these areas, understanding large networks is a fundamental problem.
Network structures are usually modelled using graph theory to represent pairwise interactions between the elements of the network. However, for very large networks, such as the internet, the brain and social networks among others, exact information about the number of nodes and other specific features of the underlying graph is not available. For this reason, there is the need for a mathematical definition of synthetic structures containing only the relevant information for a very large graph. This is equivalent to assuming that the number of nodes is so big that the graph can be well approximated with a “graph-like” object with infinite number of nodes. This motivated the development of graph limits theory, the study of graph sequences, their convergence and their limit objects. In mathematical terms, one is interested in finding a metric on the space of graphs and a completion of this space with respect to this metric. This is a very active field of mathematics that connects graph theory with many other mathematical areas such as stochastic processes, ergodic theory, spectral theory and several branches of analysis and topology.
From the rise of graph limits theory, two different cases have been mostly considered. The first case is the limits of dense graphs, i.e. when the number of edges of the graphs in the sequence is asymptotically proportional to the square of the number of vertices. This case, where the limit objects are called graphons (from graph functions), is now very well understood thanks to the contributions of L. Lovász, B. Szegedy, C. Borgs and J. Chayes among others [12, 33, 35]. The dense graph limit convergence is metrized by the cut-metric and is equivalent to the convergence of homomorphism densities. The completion of the set of all graphs in this metric is compact, i.e. every graph sequence has a convergent subsequence, which is a very useful property. A shortcoming of the dense graph limit theory is that it has not enough expressive power to study graph sequences in which the number of edges is sub-quadratic in the number of vertices. In fact, every sparse graph is considered to be similar to the empty graph in this metric. An important generalization of this theory are graphons [14, 13]. The other case that has been well studied are graph sequences with uniformly bounded degree and the associated notion of convergence was introduced by I. Benjamini and O. Schramm [9] and it has a stronger version called local-global convergence [11, 23]. The limits of such convergent sequences can be represented as objects called graphings. For a thorough treatment of these topics see the monograph by L. Lovász [34].
Unfortunately, for most applications, the really interesting case is the intermediate degree case, not covered by the previously presented theories. Real networks are usually sparse but not very sparse and heterogeneous. For this reason, the intermediate case attracted a lot of attention recently, see for example [32, 30, 31]. In particular, in a recent work Á. Backhausz and B. Szegedy introduced a new functional analytic/measure theoretic notion of convergence [6], that not only covers the intermediate degree case but also unifies the graph limit theories previously presented. This notion of convergence is called action convergence and the limit objects for graph sequences are called graphops (from graph operators). More generally this is a notion of convergence for operators, i.e. linear bounded operators
where is a probability space. As a matrix can be naturally interpreted as a operator, we obtain as a special case a notion of convergence for matrices. The notions of convergence for graphs are derived by associating to graphs (properly normalized) matrices, for example, adjacency matrices or Laplacian matrices. In this work we extend the notion of action convergence to multi-linear operators. More specifically, we consider multi-linear operators of the form
where is a probability space and is the cartesian product of with itself times. We name such operators multi-operators.
This convergence notion comes with an associated pseudo-metric . We therefore say that two multi-operators and are isomorphic if . The space of classes of isomorphism of multi-operators equipped with is a metric space.
We obtain a compactness result for multi-operators analogous to the compactness result for the case of operators: Sequences of multi-operators that have a uniform bound on the quantity
for all have a convergent subsequence in the space of isomorphism classes of multi-operators equipped with the metric . Moreover,
if the sequence is convergent with limit multi-operator .
We focus on using multi-linear action convergence to define meaningful convergence notions for tensors and hypergraphs.
Definition 1.1.
Let . An -th order -dimensional tensor consists of entries
where .
The tensor is symmetric if its entries are invariant under any permutation of their indices.
First of all, we explain how symmetric tensors can be associated with multi-operators in multiple ways. For example, for a rd order symmetric tensor
we can consider the operator
where are vectors or alternatively
where are symmetric matrices. These different choices of associating a multi-operator to a tensor give rise in general to different convergence notions for tensors. In the case of rd order symmetric tensors the second choice presented seems to be in many cases more appropriate. However, one can require action convergence of both multi-operators and associated to the tensor at the same time.
Recently, in network sciences, a lot of interest has been generated by higher-order interactions (interactions that are beyond pairwise) and the phenomena generated by them [7, 8, 16, 36, 37, 38, 15, 26, 27]. Hypergraphs are the natural mathematical/combinatorial structure to represent higher-order interactions.
Definition 1.2.
An hypergraph is a pair where is the set of vertices, is the set of edges and for each .
A hypergraph is -uniform if for every .
Limit theories for hypergraphs are much less developed than the ones for graphs due to the bigger combinatorial complexity and for this reason very limited to the uniform and dense hypergraph sequences case. The first contributions on hypergraph limits,[20, 19] by Elek and Szegedy used techniques from nonstandard analysis/model theory to define uniform and dense hypergraph limit objects. This approach using “ultralimits” is well explained in the recent book [44] by Towsner. A more classical approach using “quotients” and regularity partitions obtaining the same type of limits has been developed by Zhao in [45]. The limit objects of this convergence notions appeared earlier in the context of exchangeable arrays of random variables[28, 24, 3, 4, 5, 17]. For sparse uniform hypergraph sequences, a removal lemma is obtained in [43] using again techniques from logic but no limit theory/convergence for sparse hypergraphs is developed to the best of our knowledge. Our hypergraphs convergence based on multi-linear action convergence instead is based on functional analytic and measure-theoretic techniques and can be applied to any hypergraph sequence, also for non-uniform, very sparse and heterogeneous hypergraphs sequences.
To apply action convergence we associate to hypergraphs their adjacency tensor
Definition 1.3.
Let be a hypergraph on nodes with largest edge cardinality . The adjacency tensor of is the -th order -dimensional tensor with entries
(possibly multiplied by some normalizing constant) and as already explained we can associate a tensor with different multi-operators and therefore different convergence notions with a relationship between them. These different types of convergence are related to the different types of quasi-randomness for sequences of hypergraphs. In particular, we focus our attention on one notion of convergence obtained in such a way, which we consider being in many cases the most appropriate, and we compare it with the existing notion of convergence for dense hypergraphs (hypergraphon convergence). We underline many similarities in the two theories and we look at some motivating examples that bring us to conjecture that a modification of action convergence of the normalized adjacency tensor and hypergraphon convergence are equivalent.
The generalization of action convergence to multi-linear operators allows us to study hypergraph limits and therefore to represent conveniently large hypernetworks with objects that we will call hypergraphops. Hypergraphops are symmetric and positivity-preserving multi-operators and hypergraphs are obviously special cases of hypergraphops. We show that the space of hypergraphops (with a uniform bound on some operator norm) is closed. In fact, symmetry and positivity of multi-operators are preserved under action convergence, i.e. the limit of an action convergent sequence of symmetric, respectively positivity-preserving, multi-operators is again symmetric, respectively positivity-preserving.
We compare the action convergence metric with other norms and metrics in order to better understand this convergence. These comparisons allow us to give several examples of action convergent sequences of hypergraphs and their limits.
We also study other possible tensors associated with hypergraphs and their associated action convergence. In particular, we present possible choices to obtain meaningful limit objects for inhomogeneous and non-uniform hypergraph sequences. In particular, to the best of our knowledge, we are the first to introduce a meaningful convergence for non-uniform hypergraphs. Covering the case of non-uniform hypergraphs, our limit theory gives us a convergence for simplicial complexes as a special case answering a question in [10].
Generalising the results of [6] is technically challenging as it requires us to use multi-linear operators and tensors instead of linear operators and matrices, for which there are fewer results. Furthermore, this generalisation also significantly complicates the associated notation. However, on a more conceptual level, the main challenge of understanding the limit objects of hypergraphs requires a deeper understanding of action convergence to which we contribute here.
Structure of the paper
In Section 2 we introduce the notation and basic definitions from functional analysis and probability theory. In Section 3 we briefly recall the theory of action convergence and in Section 4 we introduce relevant notions for hypergraphs and tensors. In Section 5 we finally introduce the generalization of action convergence to multi-linear operators and in Section 6 we prove the main compactness result. Moreover, in Section 7 we study several properties of multi-linear action convergence and of the related limit objects. In Section 8 we compare the action convergence distance with other norms and distances for multi-linear operators. In Section 9 and 10 we investigate how action convergence for multi-linear operators can be specialized to tensors and hypergraphs in different ways and we study the different convergence notions obtained. In Section 11 we point out many relationships between hypergraph convergence notions obtained from action convergence and hypergraphon convergence in the context of dense hypergraph sequences and we conjecture the equivalence of a modification of action convergence and hypergraphon convergence.
2 Notation
In the following, we will denote with a standard probability space where is a algebra and is a probability measure on . We will denote with or shortened the set of probability measures on . Moreover, we will indicate the expectation of a real-valued measurable function (or in probabilistic language a random variable) on with . We indicate the (possibly infinite) norm of a real-valued measurable function with
If a measurable function has finite norm we say that is integrable (or has finite moment). We denote with the usual Banach space of the real-valued measurable integrable functions (identified if they are equal almost everywhere) on equipped with the usual norm or equivalently, in probabilistic language, the space of random variables with finite moment. We will use a lot of times the shortened notations or when there is no risk of confusion. For a set we will also denote with the space of the integrable random variables taking values in .
For a linear operator
we define the operator norm
The linear operator is said to be bounded (or equivalently continuous) if the operator norm is finite. We denote with the Banach space of linear bounded operators from to equipped with the operator norm.
A dimensional random vector is a measurable function from a probability space to and we can naturally represent it as
where are real-valued random variables on . Therefore, a real-valued random variable is a dimensional random vector. For a dimensional random vector , we denote with its distribution (or law), that is the measure on defined as
where is a set in the Borel -algebra of .
Given ,
we denote by the set . In the case of a finite probability space, the law of a random vector has a particularly easy representation. We show this with the following example that will be important in the next sections.
Example 2.1.
Let’s consider the probability space where is the uniform probability measure on and with the discrete algebra on . Then for any dimensional random vector
the law is
where is the Dirac measure centered in .
3 Action convergence
We briefly recall here, following [6], the notion of action convergence of operators, a very general notion of convergence for operators acting on spaces defined on different probability spaces, introduced in the context of graph limit theory. Other related works to this limit notion are [25, 39].
We start giving the following definition.
Definition 3.1.
A operator is a linear bounded operator
for any probability space .
A operator is acting on the probability space if the and spaces are defined on . We denote the set of all operators with and the set of all operators acting on with .
We give here an example that will be important in the following.
Example 3.2.
A matrix can be interpreted as a operator acting on the probability space where we denoted with the uniform probability measure on and with the discrete -algebra on . In fact, for
In particular, a graph can be associated to its adjacency matrix (or its Laplacian matrix) and therefore it can be interpreted as a operator.
We would now like to introduce a metric on operators possibly acting on different probability spaces. This means that we would like to equip with a metric and, therefore, with a natural notion of convergence. In reality, we will equip with a pseudo-metric and then quotient over equivalent classes (elements at distance ) of to obtain a proper metric space. We will see that for graphs (adjacency matrices of graphs) this identification of elements is exactly what we want as it identifies isomorphic graphs.
By definition, an element of is a real-valued bounded random variable on . Therefore, for a operator acting on ,
is, by definition, a real-valued random variable with finite expectation. Therefore, for functions we can consider the dimensional random vector
and in particular its distribution . For a operator , if a measure is such that
for some functions we say that is a measure generated by through . We now define the set of measures generated by . For reasons that will be clear in the following, it will be convenient to allow in our sets only measures generated by functions in , i.e. functions taking values between and almost everywhere. Therefore, we define the profile of , , as the set of measures generated by through functions in , i.e.
(1) |
This is a set of measures. To compare sets of measures, first of all, we will need a metric on the space of measures. For this reason, we recall the following well-known metric:
Definition 3.3 (Lévy-Prokhorov metric).
The Lévy-Prokhorov Metric on the space of probability measures is for
where is the Borel -algebra on , is the set of points that have Euclidean distance smaller than from .
The above metric metrizes the weak/narrow convergence for measures.
We want to be able to compare sets of measures. We, therefore, introduce the following (pseudo-)metric on the sets of measures.
Definition 3.4 (Hausdorff metric).
Given , their Hausdorff distance
Note that if and only if , where is the closure in It follows that is a pseudometric for all subsets in , and it is a metric for closed sets.
Moreover, observe that by definition, the Lévy-Prokhorov distance between probability measures is upper-bounded by and, therefore, the Hausdorff metric for sets of measures is upper-bounded by .
We are now ready to define the pseudo-metric we are interested in. Consider two operators
and
Definition 3.5 (Metrization of action convergence).
For the two -operators the action convergence metric is
Moreover, we will say that a sequence of -operators is a Cauchy sequence if the sequence is Cauchy in .
The metric has some nice compactness properties. In particular, the following theorem gives us that sets of operators with uniformly bounded norm with are pre-compact in the action convergence metric.
Theorem 3.6 (Theorem 2.14 in [6]).
Let and . Let be a Cauchy sequence of -operators with uniformly bounded norms. Then there is a -operator such that , and . Therefore, the sequence is action convergent.
Importantly, we can relate action convergence to the notions of convergence arising in the cases of dense graph sequences convergence (cut-metric, graphons) and uniformly bounded degree graph sequences convergence (local-global convergence).
In particular, consider the sequence of adjacency matrices of graphs , and let be the number of vertices of . Then,
-
•
The action convergence of the sequence
coincides with graphon convergence [6, Theorem 8.2 and Lemma 8.3]
- •
We refer to [6] for more details.
A lot of properties of matrices and graphs can be directly translated into the language of operators: self-adjointness, positivity, positivity-preservation and regularity, see Definition 3.1 in [6].
All these properties carry over in the limit if we assume a uniform bound on the norm with . For the rigorous results, see Lemma 3.2 in [6], Proposition 3.4 in[6], Corollary 2.2 in [25], where counterexamples are provided for when the assumptions of the Corollary 2.2 are not satisfied, and Proposition 3.4 in [6].
In particular, the following special class of operators is important in graph limits theory.
Definition 3.7.
A positivity-preserving and self-adjoint operator is called a graphop.
A graphop is the natural translation in the language of operators of the adjacency matrix of a graph when we consider the uniform probability measure on the nodes.
4 Tensors and hypergraphs
We start by giving some preliminary definitions and notations on tensors and hypergraphs.
We indicate a vector in by . For a set we denote with the cardinality of and with the powerset of .
Definition 4.1.
The symmetrization of an th order tensor is the -th order tensor where
where is the set of all permutations of .
It will be very convenient in the following to consider a tensor as many different possible operators.
Definition 4.2.
For an th order dimensional symmetric tensor and for , the action of on the th order dimensional symmetric tensors is the operation
The action of is an operator that sends th order dimensional symmetric tensors in an th order dimensional symmetric tensor. Therefore, the action is an operator acting on real-valued functions with domain the set of subsets of cardinality of
To make the definition more clear, we give here some examples of action of a tensor that we will also use in the following.
Example 4.3.
For an th order dimensional symmetric tensor the action of on the dimensional vectors (first-order tensors) is the operation
In the case of , a second-order tensor is a matrix and the action on a vector is just the classical matrix multiplication with the vector , i.e.
Example 4.4.
For an th order dimensional symmetric tensor and the action of on the th order dimensional symmetric tensors is the operation
In particular, for a third-order dimensional symmetric tensor the action of on the symmetric matrices and is the operation
Remark 4.5.
For an th order dimensional (not necessarily symmetric) tensor and for , we can also consider the non-symmetrized action of on the th order dimensional (not necessarily symmetric) tensors is the operation
We now introduce some notation and notions for hypergraphs.
Given an edge , we recall that we denote its cardinality by , and in the following we will denote with the maximal edge cardinality, i.e.
Moreover, we observe that for the adjacency tensor of a hypergraph on vertices with largest edge cardinality its adjacency tensor
is a standard notion for uniform hypergraphs. However, also edges with non-maximal cardinality are incorporated as repeated indices correspond to sets of lower cardinality.
We give here some examples of deterministic and random hypergraphs. We will use these examples in the following.
Example 4.6.
The complete uniform hypergraph on vertices is the hypergraph with and such that is the set of all subsets of with cardinality .
Recall that graphs are the uniform hypergraphs. Therefore, a random graph is a uniform random hypergraph. We recall here the Erdös-Renyi random graph model.
Example 4.7 (Erdös-Renyi graph).
Consider the vertex set and we connect each of the possible pairs independently with probability , i.e. following the law of independent Bernoulli random variables. This is the Erdös-Renyi random graph and we will denote it with .
A very common random uniform hypergraph that we will consider is the following.
Example 4.8 (uniform Erdős–Rényi random hypergraph).
We denote with the uniform random hypergraph with vertex set and with edge set defined as follows: For every , set of vertices of cardinality , is in with probability , i.e. every edge of cardinality is in following independent Bernoulli random variables with parameter p. In the case r=2, corresponds with the Erdős–Rényi random graph .
In the case the uniform Erdős–Rényi random hypergraph corresponds with the complete uniform hypergraph.
We give also another example of uniform random hypergraph:
Example 4.9.
We denote with the random uniform hypergraph constructed taking the vertex set and as edges the triangles of the Erdős–Rényi random graph on the same vertex set .
We can generalize naturally this random hypergraph model
Example 4.10.
We denote with the uniform random hypergraph on the vertex set constructed inductively on as follows:
-
•
for we define .
-
•
for we define as the uniform hypergraph constructed selecting as edges independently with probability the sets of vertices such that restricted to these vertices is the uniform complete hypergraph on vertices.
We notice that the random uniform hypergraph is the same as the random uniform hypergraph .
5 Multi-action convergence for multi-linear operators
In the previous section, we have seen how a hypergraph can be interpreted as a tensor and how there are various ways to interpret tensors as multi-linear operators. Therefore, we now want to generalize action convergence to general multi-linear operators.
Definition 5.1.
An order multi-operator is a multi-linear operator such that the multi-linear operator norm
is finite. We will say that a multi-operator is acting on the probability space if the and spaces are defined on . We denote the set of all th order multi-operators with and the set of all th order multi-operators acting on with .
We can relate multi- operators and tensors in multiple ways as in the following
Example 5.2.
We can interpret the action of an th order symmetric tensor as a multi-operator
where is the symmetric algebra on and we consider the uniform probability measure on , i.e.
for all . We just have to identify the set of th order symmetric tensors with in the canonical way.
Remark 5.3.
One also consider the non-symmetrized action as a multi- operator. The probability space to consider in that case is just with the discrete algebra and the uniform probability measure on
For functions we consider the dimensional random vector
for a multi-operator and we call the distribution of this random vector,
(2) |
which is a probability measure in the measure generated by the multi-operator through the ordered sequence of functions . Sometimes, we will use the abbreviation
We now define the set of measures generated by . Similarly to the action convergence in the linear case, it is convenient to allow in our sets only measures generated by functions in , i.e. functions taking values between and almost everywhere. Therefore, we define the profile of , , as the set of measures generated by by functions in .
This is a set of measures. To compare two different sets of measures we will use the Hausdorff metric (Definition 3.4) on sets of the space of probability measures (equipped with the Levy-Prokhorov metric (Definition 3.3)), that we will denote with
Remark 5.4.
We are now ready to define the pseudo-metric we are interested in. Consider two multi-operators
and
Definition 5.5 (Metrization of action convergence).
For the two order multi--operators the action convergence metric is
Remark 5.6.
As the Hausdorff metric is bounded by 1, we have that also the action convergence distance is bounded by 1.
We will say that a sequence of -operators is a Cauchy sequence if the sequence is Cauchy in .
We notice that a sequence is a Cauchy sequence if and only if for every the sequence is a Cauchy sequence in .
Remark 5.7.
The completeness of implies that the induced Hausdorff topology is also complete [21]. Therefore, a sequence is a Cauchy sequence if and only if for every there is a closed set of measures such that .
The following lemma is an equivalent of Lemma 2.6 in [6] for multi-operators and guarantees that a subsequence converges in to a closed set of measures under a uniform bound assumption on the norm.
Lemma 5.8.
Let be a sequence of th order multi-operators with uniformly bounded norms. Then, it has a subsequence that is a Cauchy sequence.
This lemma follows directly from the same standard arguments that we summarize here for completeness.
For a probability measure on let denote the maximal expectation of the marginals of ,
(3) |
For and let
Let furthermore denote the set of closed sets in the metric space .
Lemma 5.9.
The metric spaces and are both compact and complete metric spaces.
Proof.
Markov’s inequality gives uniform tightness in , which implies the compactness of for Prokhorov’s theorem. It is known that the set of closed subsets of a compact Polish space equipped with the Hausdorff metric is again compact. ∎
Lemma 5.10.
Let and let . Then for every the closure of with respect to is in .
Proof.
Let be a sequence of functions in . We have that for every and holds for . The result follows as the moments of the absolute values of the coordinates in , (3), are given by
for and
∎
As in the linear case, for a sequence of multi-operators, we will not be interested only in the convergence of the sequences of profiles but also in the existence of a multi-operator as limit object. This will actually be the convergence we are interested in.
Definition 5.11 (Action convergence of multi--operators).
We say that the sequence is action convergent to the th order multi-operator if it is a Cauchy sequence and it is such that for every positive integer the profile is the limit of the profiles sequence in the Hausdorff metric The multi-operator is the limit of the sequence .
Additionally, we will say that a sequence of multi-operators is action convergent if there exists a limit multi-operator.
Remark 5.12.
We will often use the following consequence of the definition of action convergence. For an action convergent sequence of operators to a multi-operator and for every , there are elements such that
weakly converges to
as goes to infinity.
We introduce now a multi-operator norm for spaces that is a natural generalization of the linear operator norm
Definition 5.13 (Multi-linear operator norm).
For an th order multi-operator the multi-linear operator norm is
We denote the set of all th order multi-operators with finite norm with and the set of all th order multi-operators acting on with finite norm with .
Remark 5.14.
With an abuse of notation, we can think of a multi-operator with bounded norm as a multi-linear bounded operator
by Lemma 6.6.
The following theorem is the generalization of Theorem 3.6 (Theorem 2.9 in [6]) to the multi-linear case and it states that sets of multi-operators with uniformly bounded norm with are pre-compact in the action convergence metric.
Theorem 5.15.
For and , let be a Cauchy sequence of th order multi--operators with uniformly bounded norms. Then there is a multi--operator such that , and . Therefore, the sequence is action convergent.
We give the technical proof of this theorem in the next section which is an adaptation of the proof of Theorem 2.9 in [6] to the multi-linear case.
Remark 5.16.
Observe that having a uniform bound on the norm for directly implies that we have a uniform bound on the norm for as
6 Construction of the limit object
For an th order multi--operator and let denote the closure of in the space .
This section is dedicated to showing Theorem 5.15. This technical proof is a generalization of the proof of Theorem 2.9 in [6] to the multi-linear case (the proof is similar but we have to deal with multi-linear operators and a heavier notation). Let’s consider a sequence of probability spaces and let’s assume that is a Cauchy sequence of -operators with for a fixed . For every , we can define
We aim to construct a multi--operator with -profile that is the limit of the -profiles of the operators in a given convergent sequence of operators for every fixed , i.e. we will prove that there is a -operator for some probability space such that for every we have that
Before the technical proof, we describe the main idea. For every we consider the limit of the -profiles of the sequence of operators , which is a set of measures, and we take a dense countable subset of this set. In this way, we have that each point in this dense subset can be approximated by elements in the -profiles of the sequence of operators . Moreover, every element in the -profile of involves measurable functions on (in the terminology used before the measure is generated through those functions). In probabilistic language, these functions are random variables, since is a probability space. Very roughly speaking, the main idea is to take, for every , enough functions needed to generate enough measures (contained in the profiles of the operators ) to approximate a dense countable subset of the limiting profile. These are countably many functions for each . By passing to a subsequence, we can assume that the joint distributions of these countably many functions (random variables) converge weakly and the limit is some probability measure on . Each coordinate function in the probability space on corresponds to a function involved in a -profile for some . Since every measure in the -profile comes from functions and their images, we obtain some information on a possible limiting operator. More precisely, we obtain that certain coordinate functions are the images of some other coordinate functions under the action of the candidate limit multi-linear operator. However, it is not clear that it is possible to extend the obtained multi-linear operator to the full function space on and so we need to refine the above idea.
We now make the above idea rigorous. We need to work with enough functions to represent the function space of a whole -algebra to extend the candidate limit multi-linear operator for the entire function space on . To do this, we extend the above function systems by new functions obtained by some natural operations. In order to do this, we introduce an abstract algebraic formalism involving semigroups. The most challenging part of the proof is to show that, at the end of this construction, the limit operator is well-defined and has the desired properties.
For this construction, we will use the following algebraic notion.
Definition 6.1 (Free semigroup with multi-operators).
Let and be sets. We denote by the free semigroup with generator set and multioperator set (freely acting on . More precisely, we have that is the smallest set of abstract words satisfying the following properties. (1) . (2) If , then . (3) If , then . There is a unique length function such that for , and
We give an example of a word in with set of multioperators:
where and . The length of this word is . Note that if both and are countable sets, then also is countable.
In the first technical part of the proof, we construct a function system
for some countable index set . Later, we will construct a probability measure and an operator using this function system. In the end, we will show that is an appropriate limit object for the sequence .
Construction of a function system: First, we define , the countable index set. For every , let’s consider a dense countable subset in the metric space , which is separable. Let’s define , the generator set. Therefore, the index set will be the free semigroup generated by and a set of appropriate nonlinear multi-operators . For any and let be the (bounded) continuous function defined by for and for . Finally, for every and we define , where is indexed by the pair . Observe that by definition, . Being these functions indexed by , with an abuse of notation, we will denote . Therefore, we let be as in Definition 6.1 and, thus, is countable. Furthermore, we define the functions . For every , and let be random variables in such that the joint distribution of
converges to as goes to .
At this point, we will define the functions recursively to the length of the words . The functions have been constructed above for words of length . Assume we have already constructed all the functions with for some . Consider a such that . If for some , then set . If , then set
Construction of the probability space: Let be the map such that for , and the coordinate of is equal to
where for is defined to be the projection on the th variable and . For the random variable we denote its distribution with , i.e. is the joint distribution of the functions and . Since holds (we recall the definition of , equation (3)), there exists a strictly increasing sequence in and a probability measure such that is weakly convergent to as goes to infinity. We will define and consider as a topological space, equipped with the product topology. Therefore, we constructed the probability space , where the algebra is its Borel -algebra and the probability measure obtained as weak limit of the sequence . We remark that is a probability measure, as it is the weak limit of probability distributions.
Construction of the operator: We now define an operator with the probability space defined above. For we denote with the projection to the coordinate. Observe that
(4) | ||||
Additionally, by the definition of , we also notice that for and . We want now to prove that there exists a unique -bounded th order multi-operator from to with such that holds for every .
Lemma 6.2.
For the coordinate functions on the following properties hold:
-
1.
If and , then holds in .
-
2.
If and then holds in .
-
3.
If , for every then
-
4.
For all , the linear span of the functions is dense in the space .
-
5.
Assume that and . Then holds for and we have
Remark 6.3.
When functions on are treated as functions in for some , they are identified if they differ on a set of measure zero. This standard identification of functions allows the correspondence between different coordinate functions. For example let us consider the uniform measure on which is a Borel measure on . The -coordinate function and the -coordinate function coincide in the space , as they agree on the support of . We will heavily exploit this fact in the rest of our proof.
For the proof of Lemma 6.2 we will need the following two lemmas.
Lemma 6.4 (Lemma 4.3 in [6]).
Let . For every we have that
The following lemma, which is easy to show, see Theorem 22.4 in the lecture notes [18], will be needed in the following.
Lemma 6.5.
Let . Let be a system of functions for some countable index set such that for every there is with . Let be the -algebra generated by the functions . Suppose that the constant 1 function on can be approximated by a uniformly bounded family of finite linear combinations of . Then the -closure of the linear span of is .
Finally, we come back to the proof of Lemma 6.2.
Proof.
The first statement of the lemma is shown as follows. By the construction of the function system, for every and , it holds that . Therefore, by equation (4) and the continuity of , it follows that each is supported on the closed set
Therefore, is also supported inside this set and hence the equality holds -almost everywhere for every .
The second statement is proven along the same lines as the first one. Again, by the construction of the function system, it follows that for every and we have . Thus, by the definition of , equation (4) and the continuity of we obtain that is supported inside the closed set
for every . Therefore, for every , the equality holds -almost everywhere.
To show the third claim, we recall that holds for every and thus
The sums in the factors on the right-hand side are functions in whose values for the respective are in the compact intervals for , therefore, we obtain that is a bounded, continuous function on the support of . Therefore, using that converges to weakly and equation (4) again (in particular, integrating the th power of the absolute values with respect to ), we obtain that
On the other side, weak convergence implies the following inequality:
as is a continuous non-negative function. Therefore, putting those inequalities together we obtain the third statement.
To prove the fourth statement, let be the -closure of the linear span of the function system for and .
First of all we notice that
for all and and, therefore, . From now on we will write as it does not depend on .
Now we prove that holds for every . From the second statement of the lemma, it follows that the following equality holds
(5) |
where is represented by the pair for .
Thus, we notice that the left-hand side is in as the right-hand side of (5) obviously is in . Moreover, by the third statement of the lemma. Hence, by Lemma 6.2, we have that, the left-hand side of (5) converges to in , as goes to , and hence .
For fixed , let denote the -algebra generated by the functions . Observe that already in the constant function can be approximated on . Therefore, we obtain by the first statement in this lemma and Lemma 6.2 that holds for every and . Thus, we obtained that for every the equality holds and, hence, all coordinate functions on are measurable in . This finally proves that holds for every .
From the definition of the functions and the definition of the probability measure , we directly obtain the last statement of the lemma. ∎
We will need also the following lemma to prove the existence of the multi-operator. This is the multi-linear version of a classical result about the extension of linear bounded operators defined on a dense set.
Lemma 6.6.
Let and be Banach spaces and where, for every , is a dense subspace of . For a multi-linear bounded operator
there exists a unique multi-linear bounded operator
and
Proof.
For every we define
where as where for every and the convergence is in the natural norm on . We show that this definition is independent of the sequence we choose. We consider two sequences
as . Moreover,
as the sets are dense in and therefore is dense in .
∎
We now finally define the limit operator . For , let
This defines a multi-linear operator on the linear span of . This operator is bounded by the third statement of Lemma 6.2.
Thus, there exists a unique continuous multi-linear extension on its -closure.
In fact, by the fourth statement of Lemma 6.2 and Lemma 6.6, we get that there is a unique operator with such that holds for every .
Last part of the proof: From the last statement of Lemma 6.2 together with the equality
we obtain that for every and it holds . Hence, for every we directly observe that . We now want to prove that for every and thus we still need to show the converse inclusion . Let and let . Hence, we aim to prove that
For arbitrary, it follows by the fourth statement of Lemma 6.2 that for some large enough natural number there are elements and real numbers such that for every we have , where for .
We recall that only vectors with norm bounded by are admitted in the profiles. For this reason, we will need to use a truncation function . Let be the continuous function with for for and for . We notice that holds almost everywhere as for every and By , we observe that for . Therefore, using the triangle inequality we obtain
(6) |
for . For and let and let
By the properties of convergence in distribution of random vectors (linear combinations of entries converge in distribution to the same linear combination of the entries of the limit random vector) and the definition of , it follows that
holds in . Moreover, we have
since and where the second last inequality follows from . From Lemma 11.4 we have that , where . Let
Observe that the function
is continuous. Moreover, the functions all take values in the compact interval where . Therefore, it follows that
for as is continuous, converge in distribution to , are uniformly bounded and the inequality in (6). Hence, if is large enough, then holds for and therefore by Lemma 11.4.
We choose now to be a subsequence of such that exists. Noticing that and , we get that
This inequality holds for arbitrary and, hence, we finally obtain .
Remark 6.7.
This proof works generally for any sequence of multi-operators with a uniform bound on their order. However, this proof cannot work for sequences of multi-operators in which the order of the multi-operators is diverging.
7 Properties of limit objects
In this section, we discuss some properties of multi-operators that are preserved under action convergence.
Definition 7.1.
Let be a multi--operator.
-
•
is symmetric if
holds for every and for every permutation of .
-
•
is positivity-preserving if for every with for almost every , we have that holds for almost every .
-
•
is -regular if for some .
-
•
is a hypergraphop if it is positivity-preserving and symmetric.
-
•
is atomless if is atomless.
In particular, we notice that the action of the adjacency tensor of a hypergraph is positivity-preserving and symmetric, i.e. a hypergraphop.
Remark 7.2.
The regularity property of a multi-operator is related to certain regularity properties (i.e. having constant degree) of hypergraphs. In particular, we can consider different notions of degrees for hypergraphs. For a uniform hypergraph we define for the degree as
for pairwise distinct (compare this degree notion with (17) in the following). We observe that the action of the adjacency tensor of an uniform hypergraph is regular if and only if the hypergraph has constant degrees equal to .
The following lemmas are generalizations to the multi-linear case of the results from Section 3 in [6] for action convergence and the proofs are similar.
Lemma 7.3.
Atomless multi--operators are closed with respect to .
Proof.
Let’s assume and to be two multi--operators with . Additionally, let’s suppose to be atomless. Therefore, there exists a random variable such that its distribution is uniform on . Let’s define . As we have that with and thus , where and are the marginals of and on the first coordinate. Thus, the distance between and the uniform distribution is at most . Therefore, the largest atom in is at most as by the definition of Levy-Prokhorov distance
and . Hence the largest atom in has weight at most . For this reason, if is the limit of atomless operators, then is atomless. ∎
Under uniform boundedness conditions, positivity and symmetry of multi-operators are preserved under action convergence.
Lemma 7.4.
Let and . Let be a sequence of multi--operators with a uniform bound on the -norms converging to a multi--operator . If is symmetric for every , then is also symmetric.
Proof.
Let be a permutation of To show the statement let and let . By the definition of action convergence, it follows that for every there exist functions such that weakly converges to . By Lemma 11.5, we have that goes to and goes to as goes to infinity. But additionally, we notice that
and therefore
This concludes the proof. ∎
Remark 7.5.
The action of the adjacency tensor of a hypergraph is positive and symmetric.
Moreover, positivity-preserving and regular multi-operators are also closed under action convergence, under slightly different uniform boundedness conditions.
Lemma 7.6.
Let and let be a sequence of multi--operators with a uniform bound on the -norms converging to a -operator . Then we have the following two statements.
-
1.
If is positivity-preserving for every , then is also positivity-preserving.
-
2.
If is -regular for every , then is also -regular.
Proof.
To show the first statement, let . By the definition of action convergence, there is a sequence such that weakly converges to as goes to infinity. As weakly converges to the non-negative distribution it follows that weakly converges to . Thus, by Lemma 11.6, we have that
for and, for this reason, weakly converges to the probability measure . The fact that is non-negative for every directly implies that is non-negative.
To show the second statement, let be a sequence of functions such that weakly converges to . We notice that weakly converges to and, for this reason, by Lemma 11.6 we have that
as . Hence, it follows that is the weak limit of . The result directly follows now by the fact that . ∎
Remark 7.7.
The action of the adjacency tensor of a hypergraph is positivity-preserving.
8 Norms and metrics comparison
In this section, we compare different norms and metrics for multi-operators.
The following two lemmas are generalizations of Lemmas 2.12 and 2.13 in [6].
Lemma 8.1.
Let and be positive integers and let be th order multi--operators both in for some probability space Then
Proof.
Let be arbitrary. We have that there exist functions such that is equal to the probability measure . Let . Since
holds for every , we have by Lemma 11.4 that . We obtained that
By switching the roles of and and repeating the same argument we get the above inequality with and switched. This implies the statement of the lemma. ∎
The following lemma is a direct consequence of Lemma 8.1.
Lemma 8.2.
Assume that are th order multi--operators acting on the same space . We have .
Proof.
For a multi-operator we define the multi cut norm as
We obtain that for an th order multi-operator the norm is equivalent to the multi cut norm. This is a generalization of Lemma 8.11 in [34] for graphons.
Lemma 8.3.
Let be a multi-operator The following inequality holds:
Proof.
We get the first inequality by definition:
We now show the second inequality. We observe the following equality:
Moreover, for any we have the following inequality
Therefore, we obtain ∎
Let be a bijective measure-preserving transformation. We denote with its measure-preserving inverse. The transformation induces a natural, linear action on , which we also indicate by , defined by . Furthermore, for let defined as
We observe that if , then and . Let and be two th order multi-operators such that The multi cut distance between and is defined as
where the infimum is taken over all invertible measure-preserving transformations from to
Lemma 8.4.
Assume that are th order multi--operators acting on the same space . Then .
Proof.
The proof follows directly from Lemma 8.2 and observing that for any bijective and measure preserving transformations from to itself. ∎
9 Multi-action convergence of hypergraphs and tensors
We have seen in the previous sections that hypergraphs can be naturally associated with symmetric tensors through, for example, the adjacency tensor. We can therefore study the convergence of sequences of tensors and see the convergence of hypergraphs as a particular case. Moreover, in the previous sections, we noticed that tensors can be associated with multi-linear operators in many different ways. We will compare the notions of convergence induced by the different operators associated to the same tensor. We mainly focus on symmetric tensors as we are originally motivated by undirected hypergraphs. We notice that the obtained notions of convergence are not equivalent. For simplicity, we will mainly present the convergence in the case of rd order symmetric tensors and therefore hypergraphs with maximal edge cardinality 3. However, the notions of convergence are general and cover tensors of any finite order and hypergraphs with any finite maximal edge cardinality. These convergence notions particularly make sense for uniform hypergraphs. However, we will explain, in Section 10, how one can extend these notions to not lose information regarding the non-maximal cardinality edges in non-uniform hypergraphs.
We recall the notion of action of a tensor from Section 4. For a rd-order tensor the 1-action and the action of the tensor are respectively
and
Therefore, we can interpret the action as an operator
and the action as
where is the symmetric algebra on .
Remark 9.1.
More generally the action of an th order symmetric tensor is acting on symmetric th order tensors and gives as an output another symmetric th order tensor. For this reason, this action can be interpreted as an operator
where is the symmetric algebra on .
In such a way, we obtain two notions of convergence for sequences of rd-order tensors , the one obtained by the action convergence of the sequence of multi-linear operators and the one obtained by the action convergence of the sequence of multi-linear operators .
Remark 9.2.
As already pointed out we can associate to an th order symmetric tensor its action for . These different actions can be interpreted as different multi-operators. Therefore, for a sequence of th order symmetric tensors, we obtain different notions of convergence.
We will use the results in this section later in this work to connect the metric with other norms and metrics for hypergraph limits.
9.1 Uniform bounds on sequences of actions
We recall that in the case of graphs, we typically consider the action of normalized adjacency matrices. In particular, for dense graphs, we consider
for a graph on the vertex set These linear bounded operators can be easily extended to linear bounded operators
and we have that these operators are uniformly bounded (independently by the cardinality of the vertex set ) as
We can observe from the following example that for hypergraphs with maximal edge cardinality this is not true.
Example 9.3.
For example, consider a (dense) hypergraph , its adjacency tensor and the associated multi-operator
and consider the matrices such
However, we can consider a smaller extension.
Lemma 9.4.
The sequence of operators
is uniformly bounded in operator norm.
Proof.
In these spaces, we have a uniform bound, in fact
where the last inequality follows by Cauchy-Schwartz inequality. Therefore, we obtain
where in the third inequality we used Minkowski inequality.
∎
Remark 9.5.
More in general, for , for a sequence of dense hypergraphs the sequence of actions of the relative normalized adjacency tensors cannot be extended/interpreted as a uniformly bounded sequence of linear operators from to . Therefore, one has to consider them as operators from to with and . This happens already in the case of graph limits for sparse graph sequences, and we know that this translates in larger classes of measures admitting also more irregular measures, possibly not absolutely continuous with respect to the Lebesgue measure on the interval . Instead, for every the sequence of actions of the normalized adjacency matrices of dense graphs is a uniformly bounded sequence of linear operators from to .
Remark 9.6.
The same estimates hold for the non-symmetrized action.
9.2 Properties of actions as operators
We underline here a few more properties of the action of (normalized) adjacency matrices of hypergraphs and, therefore, also of their limits by Lemma 7.4 and Lemma 7.6.
First of all, we notice that the actions of (normalized) adjacency tensors are obviously positivity-preserving multi-operators and, therefore, their action convergence limits are too. The following lemma and remark state that the action of a symmetric tensor is a symmetric multi-operator.
Lemma 9.7.
For a rd order symmetric tensor the multi-operator is symmetric.
Proof.
The result follows from the following equality
(7) | ||||
∎
Remark 9.8.
Similarly, the action of a symmetric th order dimensional symmetric tensor is symmetric for every by a similar computation.
Therefore, the limit of the sequence of symmetric tensors will also be symmetric for Lemma 7.4.
9.3 Generalization of actions
Recall the actions introduced in 4.2 for tensors. In this section, we generalize the notion of action to kernels (see below) and study its properties. We will also present a natural generalization of action, for to kernels.
Let be a probability space. We call a measurable function
such that an kernel.
We will say that an kernel is an graphon if takes values in
Remark 9.9.
This is a trivial generalization of real-valued graphons [34]. In particular, for we have that the graphons are the real-valued graphons.
Remark 9.10.
An th order dimensional tensor is an kernel where endowed with the uniform measure. One can also naturally represent a tensor with a kernel that is a step-function (as a trivial generalization of the step-representation of a graph (or matrix) for real-valued graphons).
We can identify an kernel with its action, the th order multi-operator
defined as
For a kernel we can define the cut norm as
Compare also [45].
From Lemma 8.3 we directly obtain that for an kernel the norm of the associated multi-operator is equivalent to the cut norm.
(8) |
This is a generalization of Lemma 8.11 in [34] for graphons.
For a bijective measure-preserving transformation and an kernel we denote with the kernel defined for every as
We observe that Moreover, for two kernels and on the same probability space , we define the cut metric
(9) | ||||
Therefore, from Lemma 8.4, we obtain
(10) |
This implies that convergence in the cut metric (or cut norm) of a sequence of kernels implies multi-linear action convergence of the sequence of the actions associated with the kernels.
Remark 9.11.
Similarly, we can consider the action of a symmetric kernel as the straightforward generalization of the action of a symmetric tensor to kernels. For brevity, we write down explicitly only the action for a symmetric kernel that is the multi-operator
Let’s now consider the (too) strong cut norm
Therefore, we can use the reasoning used for substituting with and with to obtain
(11) |
The (too) strong cut norm has been studied in the context of hypergraph limits before. The interested reader can find more information about it in Section 3 of [45]. There it is also explained that many interesting hypergraph sequences do not admit a convergent subsequence in this norm.
Therefore, from the results in this section, we directly get examples of convergent hypergraph limits in see the next section or Section 3 in [45].
9.4 Examples of hypergraph sequences and action convergence
The emergence of multiple operators and therefore of different notions of convergence of symmetric tensors is related to the emergence of different levels of quasi-randomness for sequences of hypergraphs [44, 2, 29].
We illustrate here this relationship with some examples.
Example 9.12.
Let’s consider the uniform Erdős–Rényi hypergraph from Example 4.8 and the uniform hypergraph , i.e. the uniform hypergraph with the triangles of an Erdős–Rényi graph from Example 4.9 as edges. We now consider the sequence of the normalized adjacency tensors associated with (a realization of) , i.e. , and the sequence of the normalized adjacency tensors associated with (a realization of) , i.e. . We remark that the normalization of the adjacency tensors we are choosing is necessary to satisfy the hypothesis of Theorem 5.15 and, therefore, to ensure a convergent (sub)sequence as shown in 9.4. However, different normalizations could be chosen as we will explore later.
If we now consider the sequences of multi-operators and they both action converge to the same limit object, the action of the constant graphon defined on where the unit interval is endowed with the Lebesgue measure. This can be easily seen using the results in Section 9.3 and known facts about these random hypergraph models and the cut norm (see Section 3 in [45]). However, the two random hypergraph models are very different. To see the combinatorial difference between these two random hypergraph models consider how many edges can be present in an induced hypergraph on vertices. In there cannot be exactly three edges but in this happens with probability
Instead, if we now consider the sequences of multi-operators and the two sequences are now converging to two different limits as we show in Lemma 9.13 below. Again one can easily see using the results in Section 9.3 and known facts about these random hypergraph models and the cut norm (see Section 3 in [45] again) that the sequence of multi-operators converges to the action of the graphon defined on where the unit interval is endowed with the Lebesgue measure. However, we cannot use the same method to say something about the sequence as is not convergent in
Lemma 9.13.
The (sub-)sequences of the multi-operators and , as defined in Example 9.12, have different action convergence limits.
Proof.
Let’s denote with the set of edges of (a realization of) the Erdős–Rényi graph from which is generated, that is the Erdős–Rényi graph from which the triangles are taken to create the edges of . Let be the matrix with every entry equal to . We can observe that the distribution
of the random vector
where
and the distribution
of the random vector
where
are very different. In fact, for we have that for any for any the probability that is an edge of is . Therefore, is a sum of Bernoulli (almost) independent random variables with parameter divided by Therefore, by (a standard argument using) the law of large numbers we obtain that
However, similarly, we obtain that
as if then for any the probability that is an edge of is but if then there is no such that is an edge of . Therefore, the profiles and of the action convergence limits and (passing to subsequences if it is necessary) of the sequences and are at Hausdorff distance bigger than a constant . Let’s suppose by contradiction that there exists such that for every
We recall that convergence in distribution to a constant and convergence in probability to the same constant are equivalent and, as the random variables are bounded between and , convergence in probability is equivalent to the convergence of the th moment. Therefore, for any , we can choose small enough such that
and, therefore,
Using Lemma 11.3, we obtain that
Therefore, for the triangular inequality we have
where and as . But this is in contradiction with
∎
Remark 9.14.
We could have deduced directly that the weak limit of
as we know the limit constant graphon on of Observe in fact that
9.5 Finite hypergraphs and action convergence
Now that we have given some motivating examples for sequences of hypergraphs with diverging number of vertices we study what action convergence and the profiles capture for finite tensors and hypergraphs.
The following theorem states that finite tensors are completely determined by the action convergence distance, up to relabelling of the indices. This is particularly interesting for adjacency tensors of hypergraphs because the following result implies that two adjacency tensors of two (finite) hypergraphs are identified if and only if the two hypergraphs are isomorphic.
Theorem 9.15.
For two rd order dimensional symmetric tensors and , the actions and are at distance zero in action convergence distance if and only if there exists a bijective map
such that
Proof.
The only non-trivial implication is the “only if” part. We observe that, in the finite case, it must exist a bijective and measure-preserving function
such that
for all symmetric matrices on .
Because, in general, to have
we need
Therefore, we can compare the two terms
and
Now, we choose and where is the indicator function of the set . Then we have
and
From the second expression, we can notice that for an element of the sum to be non-zero it is necessary that one of the following sets of conditions is satisfied:
(12) | ||||
(13) | ||||
(14) | ||||
(15) | ||||
We observe that varying we accordingly vary as is bijective. In fact, for all conditions (12), (13), (14) and (15) if there would be two distinct and in corresponding to the same then would fail to be bijective. For this reason, we obtain from the conditions (12),(13),(14) and (15) that (respectively ) depend only on the second variable. Moreover, we notice that a necessary condition to be bijective and measure-preserving (measurable) for is
(16) |
Therefore, we notice that conditions (13) and (14) would contradict condition (16). In conclusion, we can only have from (12) and (16) that
where is a permutation of Therefore, substituting and requiring that
we obtain that
∎
This result holds more generally as explained in the following remark.
Remark 9.16.
We can use the same reasoning as in the proof of Theorem 9.15 to show more generally that the -actions of two th order symmetric tensors and are completely determined by the action convergence distance, i.e. their actions are at action convergence distance zero if and only if
In fact, similarly to the case , there must exist a bijective and measure-preserving transformation
such that
for all symmetric th order tensors and where for a symmetric th order tensor we define
Moreover, using the test functions , where represents the indicator function of the set
the conditions on the imposed by the fact that is measure-preserving and bijective we obtain that for a permutation of we have
where is a permutation of
The previous theorem has the following direct important corollary:
Corollary 9.16.1.
For two hypergraphs and with maximal edge cardinality the actions of their adjacency tensors and (that are th order tensors) are identified by the action convergence metric if and only if the hypergraphs and are isomorphic.
We expect that the previous theorem and remark can be generalized to any action () of an th order tensor. The action case is trivial and we showed in the previous theorem and remark the action case.
10 Sparse and non-uniform hypergraphs and different tensors
In this section, we study how one can use action convergence for sparse hypergraph sequences and for hypergraphs with different edge cardinalities (non-uniform hypergraphs), without losing information about edges with non-maximal cardinality.
First of all, we discuss here how the sparseness of the hypergraphs interacts with our notions of action convergence. We underline that the action for uniform hypergraphs might not be the best choice for sparser hypergraphs and the action might be sometimes more appropriate as the following example shows.
Example 10.1.
Consider the uniform hypergraph given by the triangles of the sparse Erdős–Rényi random graph where and . For every we consider a realization of and the related graph on the same vertex set with the hyperedges of as triangles. Let’s denote with the (symmetric) set of edges of and recall that we denote with the adjacency tensor of . In this case, for every (sequences of) symmetric matrices, weakly converges to the Dirac function centered in In fact, if we consider the sequence of multi-operators
where is the uniform measure on , if and only if . But as , . For this reason, it might be appropriate to consider the action or change the probability measures in such a way that converges to some positive constant (for example choose as the uniform probability measures on ).
Now, we present some possible choices to adapt action convergence to non-uniform hypergraphs.
In fact, considering the action (Definition 4.2) associated with a (normalized) adjacency tensor of a hypergraph
we notice that considering the probability space with uniform probability (and the symmetric algebra) the diagonal, i.e. the set
has probability . Therefore, in the limit we have that the edges of cardinality do not play any role in the profile measures of the multi-linear operator. However, we can choose other probability measures different from the uniform distribution so that the information from the edges with lower cardinality is not lost. A natural choice for is the discrete measure defined by and . This obviously characterizes uniquely the discrete probability measure . In this case, and, therefore, the lower cardinality edges play a role in the construction of the profiles and therefore of the limit object.
Remark 10.2.
This construction of this probability measure can be naturally generalized for the case where with the symmetric -algebra.
As simplicial complexes are a special case of general hypergraphs we obtain in such a way a notion of convergence for dense simplicial complexes. Interest in a notion of convergence for dense simplicial complexes, similar to the one for dense graphs (graphons), has been expressed in [10] describing it as a “potentially very interesting direction of future research in mathematics of random complexes”. Therefore, the study of this convergence and the relative limit objects might be of special interest. In [41] the authors proposed a notion of limit for dense simplicial complexes, however, we remark that the counting lemma (Lemma 6) in [41] cannot hold as stated (the proof is incorrect and a minor adaptation of the counterexamples for uniform hypergraphs, see [45], gives a counterexample to the lemma).
We have seen that we have different possible choices for the probability measures . We obviously have also many possible options for choosing different tensors and different normalizations of these tensors.
In fact, the (normalized) adjacency tensor is not the only tensor we can associate with a hypergraph. One possibility is to normalize dividing every entry of the adjacency tensor by the quantity
(17) | |||
in the following way
It is easy to notice that
In the particular case we have
This is interesting for inhomogeneous hypergraphs and for hypergraphs with different edge cardinality. In fact, we can define on the probability measure
if and
These operators are also symmetric with respect to the right probability measure.
Lemma 10.3.
The operator is symmetric with respect to the probability measure .
Proof.
The lemma follows from the following equality
∎
Therefore, the limit of a sequence of such operators will be also symmetric and positivity-preserving by Lemma 7.4 and Lemma 7.6.
Remark 10.4.
The previous lemma can be easily generalized for the case
11 Multi-action convergence, hypergraphons and P-variables
From Theorem 8.2 and Lemma 8.3 in [6] we have that dense simple graph sequences convergence (convergence in real-valued cut distance ) is equivalent to the action convergence of the sequence of the normalized adjacency matrices
and to the action convergence of real-valued graphons.
In this section, we present some ideas on the connection of multi-action convergence and other hypergraph limits for dense hypergraph sequences.
The theory of dense uniform hypergraph limits (hypergraphons) has been developed in [20] using techniques from model theory (ultralimits, ultraproducts) and successively translated in a more standard graph limit language in [45]. A good presentation of the model-theoretic approach is given in [44]. We briefly present here the theory of dense hypergraph limits, highlighting the similarities with action convergence, following the analytic presentation in [45].
We start with some notation. For any subset , define to be the collection of all nonempty subsets of , and to be the collection of all nonempty proper subsets of . More generally, let denote the collection of all nonempty subsets of of size at most . So for instance, . We will also use the shorthand and to mean and respectively.
Any permutation of a set induces a permutation on . For a set of cardinality where , we indicate with .
The limit object of a sequence of uniform hypergraphs, i.e. an hypergraphon, is a symmetric measurable function
where symmetric means that
for every permutation of . This might be surprising because, differently from the case of graphs (), for the dimensionality of the th order adjacency tensor associated to an uniform hypergraph, , does not coincide with the dimensionality of the hypergraphon, .
The need for the additional coordinates, representing all proper subsets of , is related to the need for suitable regularity partitions for hypergraphs [22, 42, 40] and it is moreover related to the hierarchy of notions of quasi-randomness in the case of uniform hypergraphs for [44].
This is also intuitively related to the various multi--operators associated with a tensor through its actions. In fact for the additional coordinates are again needed, for example, to differentiate the limits of the sequence of the Erdős–Rényi uniform hypergraphs (Example 4.8) and the sequence of the uniform hypergraphs given by the triangles of the Erdős–Rényi graph (Example 4.9).
We notice that similarly to how we associated graphons to operators we can associate hypergraphons to multi-operators:
(18) | ||||
where here is a permutation of is the set of all the proper subsets of containing and is the symmetric algebra (i.e. the algebra generated by the subsets of that are invariant under the action of all permutations of ). In particular, for we have
We observe that there are promising similarities between the action convergence of hypergraphons and the action convergence of the action of the adjacency tensor.
Let’s consider for example the hypergraphon,
that is the limit of the sequence of hypergraphs given by the triangles of the Erdős–Rényi random graph (see [45] for example) and the action convergence limit of the action of the sequence of tensors obtained normalizing the adjacency tensors of the same hypergraphs, i.e. (recall Example 9.12), we have, for example, that
also if we take the set to be the (symmetric) set ofpairs that correspond to edges of . Then also
and, similarly,
and
Moreover, for any two hypergraphons and we can consider the multi-action convergence metric between the associated multi-operators and defined in equation (18). In particular, in this case, for the multi-operators , equation (2) in the construction of the action convergence metric is
where, for we consider
Lemma 11.1.
For any two hypergraphons and and the associated multi-operators and defined in equation (18) we have the following inequality
where for a linear combination of hypergraphons
where the supremum is taken over measurable for every such that
Remark 11.2.
More generally for two hypergraphons and we have the following bound for the multi-action convergence distance between the multi-operators and defined in equation (18):
where is the cut norm from Definition 4.3 in [45].
In particular, we obtain the following corollary from the previous lemma and remark.
Corollary 11.2.1.
We anticipate a deeper connection between multi-action convergence, variables convergence (see Section 9.4 in [47]) and convergence of hypergraphons (Definition 6.6 in [45]) that we will explore in future work.
We briefly sketch some motivating ideas here.
Let’s denote for every Let be a hypergraphon and its multi-operator representation. Observe, in particular, that we can construct also sets of measures, similarly to as done in Section 5 (see equation (2)), constructing this time probability measures out of the random vectors from to
where, for we consider as before and we additionally consider
Therefore, one can also define a metric for hypergraphons considering the Hausdorff metric on the space of measures as in Section 5, recall Definition 3.5. We observe that this metric works well only for dense hypergraph sequences. We expect this convergence to be equivalent to hypergraphon convergence (as defined in Definition 6.6 in [45]). Notably, this sketched convergence trivially implies multi-action convergence for hypergraphons. Specifically, if the action convergence limits of two sequences of hypergraphons differ, the limits under this modified convergence will also differ. Therefore, we expect action convergence to serve as a useful benchmark for understanding hypergraphon convergence. We have demonstrated many desirable properties for action convergence, which suggests (in some cases directly implies) that these properties also apply to the alternative convergence described above.
Moreover, the convergence just outlined can be viewed as a contraction of the extension of -variables to hypergraphs (as discussed in Section 9.4 in [47]). Recall that in the case of real-valued graphons, action convergence is equivalent to convergence in the real-valued cut distance, which can be considered a contraction of the -variable metric, see Definition 4.19, Corollary 7.9.1 and Lemma 7.9 in [47] (or equivalently, the unlabelled cut distance for probability graphons, see also [1] and [46]).
As already said, we will compare these convergence notions in detail in future work. We expect/conjecture the equivalence of the convergence formulated by Yufei Zaho (Definition 6.6 in [45]) and the modified version of action convergence sketched above for hypergraphons.
Appendix (technical lemmas)
For completeness, we collect here a series of lemmas proven in [6] that we used extensively throughout our work.
We start with an upper-bound on the Lévy–Prokhorov distance of the distribution of two random variables
Lemma 11.3 (Lemma 13.1 in [6]).
Let be two jointly distributed -valued random variables. Then
where is defined as in (3).
A direct consequence of the previous statement is the following Lemma.
Lemma 11.4 (Lemma 13.2 in [6]).
Let and be in for some probability space . Let . Then
The next lemma is a general probabilistic result about limits of random variables, products and expectations.
Lemma 11.5 (Lemma 13.4 in [6]).
Let . Let be a sequence of pairs of jointly distributed real-valued random variables such that and for some . Assume that the distributions of weakly converge to some probability distribution as goes to infinity. Then and
We give a last technical upper bound for the Lévy–Prokhorov distance of measures generated by a operator through specific random variables. This is a minor modification of Lemma 13.6 in [6].
Lemma 11.6.
Let and let be a multi--operator. Let and be functions in for every . Then we have
where and .
Proof.
The proof is identical to the proof of Lemma 13.6 in [6], except that we use the properties of the multi-linear norm here. ∎
Acknowledgements: The author thanks Ágnes Backhausz, Tobias Böhle, Christian Kühn, Raffaella Mulas, Florentin Münch, Balázs Szegedy, Sjoerd van der Niet and Chuang Xu for useful discussions. This work is part (in a slightly different form) of the author’s PhD thesis.
References
References
- [1] R. Abraham, J-F. Delmas, and J. Weibel. Probability-graphons: Limits of large dense weighted graphs, 2023.
- [2] E. Aigner-Horev, D. Conlon, H. Hàn, Y. Person, and M. Schacht. Quasirandomness in hypergraphs. The Electronic Journal of Combinatorics, 25(3), 2018.
- [3] D.J. Aldous. Representations for partially exchangeable arrays of random variables. Journal of Multivariate Analysis, 11(4):581–598, 1981.
- [4] D.J. Aldous. Exchangeability and continuum limits of discrete random structures. In Proceedings of the International Congress of Mathematicians 2010 (ICM 2010) (In 4 Volumes) Vol. I: Plenary Lectures and Ceremonies Vols. II–IV: Invited Lectures, pages 141–153. World Scientific, 2010.
- [5] T. Austin. On exchangeable random variables and the statistics of large graphs and hypergraphs. Probability Surveys, 5:80–145, 2008.
- [6] Á. Backhausz and B. Szegedy. Action convergence of operators and graphs. Canadian Journal of Mathematics, 74(1):72–121, 2022.
- [7] F. Battiston, G. Cencetti, I. Iacopini, V. Latora, M. Lucas, A. Patania, J. Young, and G. Petri. Networks beyond pairwise interactions: Structure and dynamics. Physics Reports, 874:1–92, 2020.
- [8] F. Battiston and G. Petri. Higher-Order Systems. Understanding Complex Systems. Springer Cham, 2022.
- [9] I. Benjamini and O. Schramm. Recurrence of distributional limits of finite planar graphs. Electronic Journal of Probability, 6:1 – 13, 2001.
- [10] O. Bobrowski and D. Krioukov. Random Simplicial Complexes: Models and Phenomena, pages 59–96. Springer International Publishing, Cham, 2022.
- [11] B. Bollobás and O. Riordan. Sparse graphs: Metrics and random models. Random Struct. Algorithms, 39(1):1–38, 2011.
- [12] C. Borgs, J. Chayes, L. Lovász, V.T. Sós, and K. Vesztergombi. Convergent sequences of dense graphs I: Subgraph frequencies, metric properties and testing. Advances in Mathematics, 219(6):1801–1851, 2008.
- [13] C. Borgs, J.T. Chayes, H. Cohn, and Y. Zhao. An theory of sparse graph convergence II: LD convergence, quotients and right convergence. The Annals of Probability, 46(1):337–396, 2018.
- [14] C. Borgs, J.T. Chayes, H. Cohn, and Y. Zhao. An theory of sparse graph convergence I: Limits, sparse random graph models, and power law distributions. Transactions of the American Mathematical Society, 2019.
- [15] T. Böhle, C. Kuehn, R. Mulas, and J. Jost. Coupled hypergraph maps and chaotic cluster synchronization. Europhysics Letters, 136(4):40005, 2022.
- [16] T. Carletti, D. Fanelli, and S. Nicoletti. Dynamical systems on hypergraphs. Journal of Physics: Complexity, 1(3):035006, 2020.
- [17] P. Diaconis and S. Janson. Graph limits and exchangeable random graphs. Rendiconti di Matematica e delle sue Applicazioni, Serie VII, 28:33–61, 2008.
- [18] B. K. Driver. Analysis tools with examples. https://mathweb.ucsd.edu/~bdriver/DRIVER/Book/anal.pdf, 2004.
- [19] G. Elek and B. Szegedy. Limits of hypergraphs, removal and regularity lemmas. a non-standard approach, 2007, arXiv:0705.2179 [math.CO].
- [20] G. Elek and B. Szegedy. A measure-theoretic approach to the theory of dense hypergraphs. Advances in Mathematics, 231(3):1731–1772, 2012.
- [21] R. A. Gordon. Real analysis: A first course. Addison Wesley Higher Mathematics, Reading, MA. Pearson, 2001.
- [22] W. T. Gowers. Hypergraph regularity and the multidimensional szemerédi theorem. Annals of Mathematics, 166(3):897–946, 2007.
- [23] H. Hatami, L. Lovász, and B. Szegedy. Limits of locally–globally convergent graph sequences. Geometric and Functional Analysis, 24:269–296, 2014.
- [24] D. N. Hoover. Relations on probability spaces and arrays of random variables. Institute for Advanced Study, 1979.
- [25] A. Hrušková. Limits of action convergent graph sequences with unbounded -norms, 2022, arXiv:2210.10720 [math.CO].
- [26] J. Jost and R. Mulas. Hypergraph laplace operators for chemical reaction networks. Advances in Mathematics, 351:870–896, 2019.
- [27] J. Jost, R. Mulas, and D. Zhang. Spectra of Discrete Structures. Under review, 2023.
- [28] O. Kallenberg. Symmetries on random arrays and set-indexed processes. Journal of Theoretical Probability, 5(4):727–765, 1992.
- [29] Y. Kohayakawa, V. Rödl, and J. Skokan. Hypergraphs, quasi-randomness, and conditions for regularity. Journal of Combinatorial Theory, Series A, 97(2):307–352, 2002.
- [30] D. Kunszenti-Kovács, L. Lovász, and B. Szegedy. Measures on the square as sparse graph limits. Journal of Combinatorial Theory, Series B, 138:1–40, 2019.
- [31] D. Kunszenti-Kovács, L. Lovász, and B. Szegedy. Multigraph limits, unbounded kernels, and banach space decorated graphs. Journal of Functional Analysis, 282(2):109284, 2022.
- [32] D. Kunszenti-Kovács, L. Lovász, and B. Szegedy. Subgraph densities in markov spaces. Advances in Mathematics, 437:109414, 2024.
- [33] L. Lovász and B. Szegedy. Szemerédi’s lemma for the analyst. GAFA Geometric And Functional Analysis, 17:252–270, 2007.
- [34] L. Lovász. Large Networks and Graph Limits., volume 60 of Colloquium Publications. American Mathematical Society, 2012.
- [35] L. Lovász and B. Szegedy. Limits of dense graph sequences. Journal of Combinatorial Theory, Series B, 96(6):933–957, 2006.
- [36] S. Majhi, M. Perc, and D. Ghosh. Dynamics on higher-order networks: A review. Journal of the Royal Society Interface, 19(188):20220043, 2022.
- [37] R. Mulas, D. Horak, and J. Jost. Graphs, simplicial complexes and hypergraphs: Spectral theory and topology. In F. Battiston and G. Petri, editors, Higher order systems. Springer, 2022.
- [38] R. Mulas, C. Kuehn, and J. Jost. Coupled dynamics on hypergraphs: Master stability of steady states and synchronization. Phys. Rev. E, 101:062313, 2020.
- [39] R. Mulas and G. Zucal. A measure-theoretic representation of graphs. Periodica Mathematica Hungarica, 88:8–24, 2024.
- [40] B. Nagle, V. Rödl, and M. Schacht. The counting lemma for regular k-uniform hypergraphs. Random Structures and Algorithms, 28:113–179, 2006.
- [41] T. M. Roddenberry and S. Segarra. Limits of dense simplicial complexes. Journal of Machine Learning Research, 24(225):1–42, 2023.
- [42] V. Rödl and J. Skokan. Regularity lemma for k-uniform hypergraphs. Random Struct. Algorithms, 25(1):1–42, 2004.
- [43] H. Towsner. An analytic approach to sparse hypergraphs: Hypergraph removal. Discrete Analysis, 3, 04 2012.
- [44] H Towsner. Randomess in the limit, 2022.
- [45] Y. Zhao. Hypergraph limits: A regularity approach. Random Structures and Algorithms, 47, 03 2014.
- [46] G. Zucal. Probability graphons: the right convergence point of view, 2024, arxiv:2407.05998v2 [math.PR].
- [47] G. Zucal. Probability graphons and P-variables: two equivalent viewpoints for dense weighted graph limits, 2024, arxiv:2408.07572 [math.PR].