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

Action convergence of general hypergraphs and tensors

Giulio Zucal [email protected]
Abstract

Action convergence provides a limit theory for linear bounded operators An:L(Ωn)L1(Ωn)A_{n}:L^{\infty}(\Omega_{n})\longrightarrow L^{1}(\Omega_{n}) where Ωn\Omega_{n} 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 An:L(Ωn)××L(Ωn)L1(Ωn)A_{n}:L^{\infty}(\Omega_{n})\times\ldots\times L^{\infty}(\Omega_{n})\longrightarrow L^{1}(\Omega_{n}). 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 LpL^{p}-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 PP-operators, i.e. linear bounded operators

A:L(Ω)L1(Ω)A:L^{\infty}(\Omega)\longrightarrow L^{1}(\Omega)

where Ω\Omega is a probability space. As a matrix can be naturally interpreted as a PP-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

A:L(Ω)rL1(Ω)A:L^{\infty}(\Omega)^{r}\longrightarrow L^{1}(\Omega)

where Ω\Omega is a probability space and L(Ω)r=L(Ω)××L(Ω)L^{\infty}(\Omega)^{r}=L^{\infty}(\Omega)\times\ldots\times L^{\infty}(\Omega) is the cartesian product of L(Ω)L^{\infty}(\Omega) with itself rr times. We name such operators multi-PP-operators.

This convergence notion comes with an associated pseudo-metric dMd_{M}. We therefore say that two multi-PP-operators AA and BB are isomorphic if dM(A,B)=0d_{M}(A,B)=0. The space of classes of isomorphism of multi-PP-operators equipped with dMd_{M} is a metric space.

We obtain a compactness result for multi-PP-operators analogous to the compactness result for the case of PP-operators: Sequences of multi-PP-operators (An)n(A_{n})_{n} that have a uniform bound C>0C>0 on the quantity

Anp1,,prq=sup0f1,,frL(Ωn)An[f1,,fr]qf1p1frprC\|A_{n}\|_{p_{1},\ldots,p_{r}\rightarrow q}=\sup_{0\neq f_{1},\ldots,f_{r}\in L^{\infty}(\Omega_{n})}\frac{\|A_{n}[f_{1},\ldots,f_{r}]\|_{q}}{\|f_{1}\|_{p_{1}}\ldots\|f_{r}\|_{p_{r}}}\leq C

for all nn\in\mathbb{N} have a convergent subsequence in the space of isomorphism classes of multi-PP-operators equipped with the metric dMd_{M}. Moreover,

Ap1,,prqlimnAnp1,,prqC.\|A\|_{p_{1},\ldots,p_{r}\rightarrow q}\leq\lim_{n\rightarrow\infty}\|A_{n}\|_{p_{1},\ldots,p_{r}\rightarrow q}\leq C.

if the sequence is convergent with limit multi-PP-operator AA.

We focus on using multi-linear action convergence to define meaningful convergence notions for tensors and hypergraphs.

Definition 1.1.

Let r,n2r,n\geq 2. An rr-th order nn-dimensional tensor TT consists of nrn^{r} entries

Ti1,,ir,T_{i_{1},\ldots,i_{r}}\in\mathbb{R},

where i1,,ir[n]i_{1},\dots,i_{r}\in[n].
The tensor TT 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-PP-operators in multiple ways. For example, for a 33-rd order symmetric tensor

Ti1,i2,i3T_{i_{1},i_{2},i_{3}}

we can consider the operator

T1[v,w]=i1,i2=1nTi1,i2,i3vi1wi2T_{1}[\mathrm{v},\mathrm{w}]=\sum^{n}_{i_{1},i_{2}=1}T_{i_{1},i_{2},i_{3}}v_{i_{1}}w_{i_{2}}

where v=(vi)i,w=(wi)in\mathrm{v}=(v_{i})_{i},\mathrm{w}=(w_{i})_{i}\in\mathbb{R}^{n} are vectors or alternatively

T2[f,g]=12(i2=1nTi1,i2,i3fi1,i2gi2,i3+i2=1nTi1,i2,i3fi3,i2gi2,i1)T_{2}[f,g]=\frac{1}{2}(\sum^{n}_{i_{2}=1}T_{i_{1},i_{2},i_{3}}f_{i_{1},i_{2}}g_{i_{2},i_{3}}+\sum^{n}_{i_{2}=1}T_{i_{1},i_{2},i_{3}}f_{i_{3},i_{2}}g_{i_{2},i_{1}})

where f=(fij)ij,g=(gij)ijf=(f_{ij})_{ij},g=(g_{ij})_{ij} are symmetric matrices. These different choices of associating a multi-PP-operator to a tensor give rise in general to different convergence notions for tensors. In the case of 33-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-PP-operators (T)1(T)_{1} and (T)2(T)_{2} associated to the tensor TT 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 H=(V,E)H=(V,E) where V={v1,,vn}V=\{v_{1},\ldots,v_{n}\} is the set of vertices, E={e1,,em}E=\{e_{1},\ldots,e_{m}\} is the set of edges and eV\emptyset\neq e\subseteq V for each eEe\in E.
A hypergraph HH is kk-uniform if |e|=k|e|=k for every eEe\in E.

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 H=(V,E)H=(V,E) be a hypergraph on nn nodes with largest edge cardinality rr. The adjacency tensor of HH is the rr-th order nn-dimensional tensor A=A(H)A=A(H) with entries

Ai1,,ir:={0 if {vi1,,vir}E1 if {vi1,,vir}E.A_{i_{1},\ldots,i_{r}}:=\begin{cases}0&\text{ if }\{v_{i_{1}},\ldots,v_{i_{r}}\}\notin E\\ 1&\text{ if }\{v_{i_{1}},\ldots,v_{i_{r}}\}\in E.\\ \end{cases}

(possibly multiplied by some normalizing constant) and as already explained we can associate a tensor with different multi-PP-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-PP-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-PP-operators are preserved under action convergence, i.e. the limit of an action convergent sequence of symmetric, respectively positivity-preserving, multi-PP-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 (Ω,,)(\Omega,\mathcal{F},\mathbb{P}) a standard probability space where \mathcal{F} is a σ\sigma-algebra and \mathbb{P} is a probability measure on (Ω,)(\Omega,\mathcal{F}). We will denote with 𝒫(Ω,)\mathcal{P}(\Omega,\mathcal{F}) or shortened 𝒫(Ω)\mathcal{P}(\Omega) the set of probability measures on (Ω,)(\Omega,\mathcal{F}). Moreover, we will indicate the expectation of a real-valued measurable function (or in probabilistic language a random variable) ff on (Ω,,)(\Omega,\mathcal{F},\mathbb{P}) with 𝔼[f]\mathbb{E}[f]. We indicate the (possibly infinite) LpL^{p}-norm of a real-valued measurable function ff with

fp=(Ω|f(ω)|p𝑑(ω))1/p=(𝔼[|f|p])1/p.\|f\|_{p}=\left(\int_{\Omega}|f(\omega)|^{p}d\mathbb{P}(\omega)\right)^{1/p}=\left(\mathbb{E}[|f|^{p}]\right)^{1/p}.

If a measurable function ff has finite LpL^{p}-norm we say that ff is pp-integrable (or has finite pp-moment). We denote with Lp(Ω,,,)L^{p}(\Omega,\mathcal{F},\mathbb{P},) the usual Banach space of the real-valued measurable pp-integrable functions (identified if they are equal almost everywhere) on (Ω,,)(\Omega,\mathcal{F},\mathbb{P}) equipped with the usual LpL^{p}-norm or equivalently, in probabilistic language, the space of random variables with finite pp-moment. We will use a lot of times the shortened notations Lp(Ω)L^{p}(\Omega) or LpL^{p} when there is no risk of confusion. For a set SS\subset\mathbb{R} we will also denote with LSp(Ω)L_{S}^{p}(\Omega) the space of the pp-integrable random variables taking values in SS.

For a linear operator

A:Lp(Ω,,)\displaystyle A:L^{p}(\Omega,\mathcal{F},\mathbb{P}) Lq(Ω,,)\displaystyle\longrightarrow L^{q}(\Omega,\mathcal{F},\mathbb{P})
fAf\displaystyle f\mapsto Af

we define the (p,q)(p,q)-operator norm

Apq=supfLp,f0Afqfp.\|A\|_{p\rightarrow q}=\sup_{f\in L^{p},\ f\neq 0}\frac{\|Af\|_{q}}{\|f\|_{p}}.

The linear operator AA is said to be bounded (or equivalently continuous) if the operator norm is finite. We denote with p,q{\mathcal{B}}_{p,q} the Banach space of linear bounded operators from Lp(Ω)L^{p}(\Omega) to Lq(Ω)L^{q}(\Omega) equipped with the (p,q)(p,q)-operator norm.
A kk-dimensional random vector is a measurable function 𝐟\mathbf{f} from a probability space (Ω,,)(\Omega,\mathcal{F},\mathbb{P}) to k\mathbb{R}^{k} and we can naturally represent it as

𝐟=(f1,,fk),\mathbf{f}=(f_{1},\ldots,f_{k}),

where f1,,fkf_{1},\ldots,f_{k} are real-valued random variables on (Ω,,,)(\Omega,\mathcal{F},\mathbb{P},). Therefore, a real-valued random variable is a 11-dimensional random vector. For a kk-dimensional random vector 𝐟\mathbf{f}, we denote with (𝐟)=(f1,,fk)\mathcal{L}(\mathbf{f})=\mathcal{L}(f_{1},\ldots,f_{k}) its distribution (or law), that is the measure on k\mathbb{R}^{k} defined as

(𝐟)(A)=(𝐟1(A))\mathcal{L}(\mathbf{f})(A)=\mathbb{P}(\mathbf{f}^{-1}(A))

where AA is a set in the Borel σ\sigma-algebra of k\mathbb{R}^{k}.
Given nn\in\mathbb{N}, we denote by [n][n] the set {1,,n}\{1,\ldots,n\}. 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 ([n],𝒟,𝒰)([n],\mathcal{D},\mathcal{U}) where 𝒰\mathcal{U} is the uniform probability measure on [n][n] and with 𝒟\mathcal{D} the discrete σ\sigma-algebra on [n][n]. Then for any kk-dimensional random vector

𝐟=(f1,,fk)\mathbf{f}=(f_{1},\ldots,f_{k})

the law (𝐟)\mathcal{L}(\mathbf{f}) is

(𝐟)=1ni=1nδ(f1(i),,fk(i))\mathcal{L}(\mathbf{f})=\frac{1}{n}\sum^{n}_{i=1}\delta_{(f_{1}(i),\ldots,f_{k}(i))}

where δ(x1,,xk)\delta_{(x_{1},\ldots,x_{k})} is the Dirac measure centered in (x1,,xk)k(x_{1},\ldots,x_{k})\in\mathbb{R}^{k}.

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 LpL^{p} 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 PP-operator is a linear bounded operator

A:L(Ω,,)\displaystyle A:L^{\infty}(\Omega,\mathcal{F},\mathbb{P}) L1(Ω,,)\displaystyle\longrightarrow L^{1}(\Omega,\mathcal{F},\mathbb{P})

for any probability space (Ω,,)(\Omega,\mathcal{F},\mathbb{P}).
A PP-operator AA is acting on the probability space (Ω,,)(\Omega,\mathcal{F},\mathbb{P}) if the L1L^{1} and LL^{\infty} spaces are defined on (Ω,,)(\Omega,\mathcal{F},\mathbb{P}). We denote the set of all PP-operators with \mathcal{B} and the set of all PP-operators acting on (Ω,,)(\Omega,\mathcal{F},\mathbb{P}) with (Ω,,)\mathcal{B}(\Omega,\mathcal{F},\mathbb{P}).

We give here an example that will be important in the following.

Example 3.2.

A matrix A=(Ai,j)i,j[n]A=(A_{i,j})_{i,j\in[n]} can be interpreted as a PP-operator acting on the probability space Ω=([n],𝒟,𝒰)\Omega=([n],\mathcal{D},\mathcal{U}) where we denoted with 𝒰\mathcal{U} the uniform probability measure on [n][n] and with 𝒟\mathcal{D} the discrete σ\sigma-algebra on [n][n]. In fact, for v=(vi)i[n]nL(Ω)L1(Ω)\mathrm{v}=(v_{i})_{i\in[n]}\in\mathbb{R}^{n}\cong L^{\infty}(\Omega)\cong L^{1}(\Omega)

A:L(Ω)L1(Ω)A:L^{\infty}(\Omega)\longrightarrow L^{1}(\Omega)
(Av)i=j=1nAijvj.(A\mathrm{v})_{i}=\sum^{n}_{j=1}A_{ij}v_{j}.

In particular, a graph can be associated to its adjacency matrix (or its Laplacian matrix) and therefore it can be interpreted as a PP-operator.

We would now like to introduce a metric on PP-operators possibly acting on different probability spaces. This means that we would like to equip \mathcal{B} with a metric and, therefore, with a natural notion of convergence. In reality, we will equip \mathcal{B} with a pseudo-metric and then quotient over equivalent classes (elements at distance 0) of \mathcal{B} 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 ff of L(Ω,,)L^{\infty}(\Omega,\mathcal{F},\mathbb{P}) is a real-valued bounded random variable on (Ω,,)(\Omega,\mathcal{F},\mathbb{P}). Therefore, for a PP-operator AA acting on (Ω,,)(\Omega,\mathcal{F},\mathbb{P}),

AfL1(Ω,,)Af\in L^{1}(\Omega,\mathcal{F},\mathbb{P})

is, by definition, a real-valued random variable with finite expectation. Therefore, for functions f1,,fkL(Ω)f_{1},\ldots,f_{k}\in L^{\infty}(\Omega) we can consider the 2k2k-dimensional random vector

(f1,Af1,,fk,Afk)(f_{1},Af_{1},\ldots,f_{k},Af_{k})

and in particular its distribution (f1,Af1,,fk,Afk)𝒫(2k)\mathcal{L}(f_{1},Af_{1},\ldots,f_{k},Af_{k})\in\mathcal{P}(\mathbb{R}^{2k}). For a PP-operator AA, if a measure μ𝒫(2k)\mu\in\mathcal{P}(\mathbb{R}^{2k}) is such that

μ=(f1,Af1,,fk,Afk)\mu=\mathcal{L}(f_{1},Af_{1},\ldots,f_{k},Af_{k})

for some functions f1,,fkL(Ω)f_{1},\ldots,f_{k}\in L^{\infty}(\Omega) we say that μ\mu is a measure generated by AA through f1,,fkf_{1},\ldots,f_{k}. We now define the set of measures generated by AA. For reasons that will be clear in the following, it will be convenient to allow in our sets only measures generated by functions in L[1,1](Ω)L_{[-1,1]}^{\infty}(\Omega), i.e. functions taking values between +1+1 and 1-1 almost everywhere. Therefore, we define the kk-profile of AA, 𝒮k(A){\mathcal{S}}_{k}(A), as the set of measures generated by AA through functions in L[1,1](Ω)L_{[-1,1]}^{\infty}(\Omega), i.e.

𝒮k(A)=f1,,fkL[1,1](Ω){(f1,Af1,,fk,Afk)}.{\mathcal{S}}_{k}(A)=\bigcup_{f_{1},\ldots,f_{k}\in L_{[-1,1]}^{\infty}(\Omega)}\{\mathcal{L}(f_{1},Af_{1},\ldots,f_{k},Af_{k})\}. (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 d𝒫d_{\mathcal{LP}} on the space of probability measures 𝒫(k)\mathcal{P}\left(\mathbb{R}^{k}\right) is for η1,η2𝒫(k)\eta_{1},\eta_{2}\in\mathcal{P}\left(\mathbb{R}^{k}\right)

d𝒫(η1,η2)=\displaystyle d_{\mathcal{LP}}\left(\eta_{1},\eta_{2}\right)= inf{ε>0:η1(U)η2(Uε)+ε and \displaystyle\inf\left\{\varepsilon>0:\eta_{1}(U)\leq\eta_{2}\left(U^{\varepsilon}\right)+\varepsilon\text{ and }\right.
η2(U)η1(Uε)+ε for all U𝒰k},\displaystyle\left.\eta_{2}(U)\leq\eta_{1}\left(U^{\varepsilon}\right)+\varepsilon\text{ for all }U\in\mathcal{U}_{k}\right\},

where 𝒰k\mathcal{U}_{k} is the Borel σ\sigma-algebra on k\mathbb{R}^{k}, UεU^{\varepsilon} is the set of points that have Euclidean distance smaller than ε\varepsilon from UU.

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 X,Y𝒫(k)X,Y\subset\mathcal{P}\left(\mathbb{R}^{k}\right), their Hausdorff distance

dH(X,Y):=max{supxXinfyYd𝒫(x,y),supyYinfxXd𝒫(x,y)}.d_{H}(X,Y):=\max\left\{\sup_{x\in X}\inf_{y\in Y}d_{\mathcal{LP}}(x,y),\sup_{y\in Y}\inf_{x\in X}d_{\mathcal{LP}}(x,y)\right\}.

Note that dH(X,Y)=0d_{H}(X,Y)=0 if and only if cl(X)=cl(Y)\operatorname{cl}(X)=\operatorname{cl}(Y), where cl\operatorname{cl} is the closure in d𝒫.d_{\mathcal{LP}}. It follows that dHd_{H} is a pseudometric for all subsets in 𝒫(k)\mathcal{P}\left(\mathbb{R}^{k}\right), and it is a metric for closed sets.

Moreover, observe that by definition, the Lévy-Prokhorov distance between probability measures is upper-bounded by 11 and, therefore, the Hausdorff metric for sets of measures is upper-bounded by 11.

We are now ready to define the pseudo-metric we are interested in. Consider two PP-operators

A:L(Ω1,1,1)L1(Ω1,1,1)A:L^{\infty}(\Omega_{1},\mathcal{F}_{1},\mathbb{P}_{1})\rightarrow L^{1}(\Omega_{1},\mathcal{F}_{1},\mathbb{P}_{1})

and

B:L(Ω2,2,2)L1(Ω2,2,2).B:L^{\infty}(\Omega_{2},\mathcal{F}_{2},\mathbb{P}_{2})\rightarrow L^{1}(\Omega_{2},\mathcal{F}_{2},\mathbb{P}_{2}).
Definition 3.5 (Metrization of action convergence).

For the two PP-operators A,BA,B the action convergence metric is

dM(A,B):=k=12kdH(𝒮k(A),𝒮k(B)).d_{M}(A,B):=\sum_{k=1}^{\infty}2^{-k}d_{H}\left({\mathcal{S}}_{k}(A),{\mathcal{S}}_{k}(B)\right).

Moreover, we will say that a sequence of PP-operators {Ai(Ωi)}i=1\left\{A_{i}\in\mathcal{B}\left(\Omega_{i}\right)\right\}_{i=1}^{\infty} is a Cauchy sequence if the sequence is Cauchy in dMd_{M}.

The metric dMd_{M} has some nice compactness properties. In particular, the following theorem gives us that sets of PP-operators with uniformly bounded (p,q)(p,q)-norm with pp\neq\infty are pre-compact in the action convergence metric.

Theorem 3.6 (Theorem 2.14 in [6]).

Let p[1,)p\in[1,\infty) and q[1,]q\in[1,\infty]. Let {Ai}i=1\{A_{i}\}_{i=1}^{\infty} be a Cauchy sequence of PP-operators with uniformly bounded pq\|\cdot\|_{p\to q} norms. Then there is a PP-operator AA such that limidM(Ai,A)=0\lim_{i\to\infty}d_{M}(A_{i},A)=0, and ApqsupiAipq\|A\|_{p\to q}\leq\sup_{i\in\mathbb{N}}\|A_{i}\|_{p\to q}. Therefore, the sequence {Ai}i=1\{A_{i}\}_{i=1}^{\infty} 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 AnA_{n} of graphs GnG_{n}, and let vnv_{n} be the number of vertices of GnG_{n}. Then,

  • The action convergence of the sequence

    Anvn\frac{A_{n}}{v_{n}}

    coincides with graphon convergence [6, Theorem 8.2 and Lemma 8.3]

  • The action convergence of the sequence

    AnA_{n}

    coincides with local-global convergence [6, Theorem 9.2].

We refer to [6] for more details.

A lot of properties of matrices and graphs can be directly translated into the language of PP-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 (p,q)(p,q)-norm with p,q{1,}p,q\notin\{1,\infty\}. 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 PP-operators is important in graph limits theory.

Definition 3.7.

A positivity-preserving and self-adjoint PP-operator is called a graphop.

A graphop is the natural translation in the language of PP-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 n\mathbb{R}^{n} by x=(x1,,xn)\textbf{x}=(x_{1},\dots,x_{n}). For a set EE we denote with |E||E| the cardinality of EE and with 2E2^{E} the powerset of EE.

Definition 4.1.

The symmetrization of an rr-th order tensor TT is the rr-th order tensor Sym(T)Sym(T) where

(Sym(T))i1,ir=1r!σΣTiσ(1),,iσ(r),(Sym(T))_{i_{1}\ldots,i_{r}}=\frac{1}{r!}\sum_{\sigma\in\Sigma}T_{i_{\sigma(1)},\ldots,i_{\sigma(r)}},

where Σ\Sigma is the set of all permutations of [r][r].

It will be very convenient in the following to consider a tensor as many different possible operators.

Definition 4.2.

For an rr-th order nn-dimensional symmetric tensor TT and for s[r1]s\in[r-1], the ss-action of TT on the ss-th order nn-dimensional symmetric tensors f(1),,f(r1)f^{(1)},\ldots,f^{(r-1)} is the operation

(T[f(1),,f(r1)])i1,,is=\displaystyle(T[f^{(1)},\ldots,f^{(r-1)}])_{i_{1},\ldots,i_{s}}=
Sym(j1,,jrs=1nTj1,,jrs,i1,,is\displaystyle Sym(\sum^{n}_{j_{1},\ldots,j_{r-s}=1}T_{j_{1},\ldots,j_{r-s},i_{1},\ldots,i_{s}} fi2,,is,j1(1)fi3,,is,j1,j2(2)fjr2s+2,,jrs,i1(r2s+1)fjrs,i1,,is1(r1)).\displaystyle f^{(1)}_{i_{2},\ldots,i_{s},j_{1}}f^{(2)}_{i_{3},\ldots,i_{s},j_{1},j_{2}}\ldots f^{(r-2s+1)}_{j_{r-2s+2},\ldots,j_{r-s},i_{1}}\ldots f^{(r-1)}_{j_{r-s},i_{1},\ldots,i_{s-1}}).

The ss-action of TT is an operator that sends r1r-1 ss-th order nn-dimensional symmetric tensors in an ss-th order nn-dimensional symmetric tensor. Therefore, the ss-action is an operator acting on real-valued functions with domain the set of subsets of cardinality ss of [n].[n].

To make the definition more clear, we give here some examples of ss-action of a tensor that we will also use in the following.

Example 4.3.

For an rr-th order nn-dimensional symmetric tensor TT the 11-action of TT on the nn-dimensional vectors (first-order tensors) f1,,fr1nf^{1},\ldots,f^{r-1}\in\mathbb{R}^{n} is the operation

(T[f(1),,f(r1)])i=j2,jr1=1nTi,j2,,jr1fj2(1)fjr1(r1).(T[f^{(1)},\ldots,f^{(r-1)}])_{i}=\sum^{n}_{j_{2},\ldots j_{r-1}=1}T_{i,j_{2},\ldots,j_{r-1}}f^{(1)}_{j_{2}}\ldots f^{(r-1)}_{j_{r-1}}.

In the case of r=2r=2, a second-order tensor TT is a matrix and the 11-action on a vector ff is just the classical matrix multiplication with the vector fnf\in\mathbb{R}^{n}, i.e.

(Tf)i=j=1nTijfj.(Tf)_{i}=\sum^{n}_{j=1}T_{ij}f_{j}.
Example 4.4.

For an rr-th order nn-dimensional symmetric tensor TT and the (r1)(r-1)-action of TT on the ss-th order nn-dimensional symmetric tensors f(1),,f(r1)f^{(1)},\ldots,f^{(r-1)} is the operation

(T[f(1),\displaystyle(T[f^{(1)},\ldots ,f(r1)])i1,,ir1\displaystyle,f^{(r-1)}])_{i_{1},\ldots,i_{r-1}}
=Sym(j=1nTj,i2,,irfj,i1,i2,,is1,i^s(1)fj,i1,i2,,i^pis(p)fj,i^1,i2,i3,is(r1)).\displaystyle=Sym(\sum^{n}_{j=1}T_{j,i_{2},\ldots,i_{r}}f^{(1)}_{j,i_{1},i_{2},\ldots,i_{s-1},\hat{i}_{s}}\ldots f^{(p)}_{j,i_{1},i_{2},\ldots,\hat{i}_{p}\ldots i_{s}}\ldots f^{(r-1)}_{j,\hat{i}_{1},i_{2},i_{3}\ldots,i_{s}}).

In particular, for a third-order nn-dimensional symmetric tensor TT the 22-action of TT on the n×nn\times n symmetric matrices f=(fi,j)i,j[n]f=(f_{i,j})_{i,j\in[n]} and g=(gi,j)i,j[n]g=(g_{i,j})_{i,j\in[n]} is the operation

(T[f,g])i,k=12(j=1nTj,i,kfj,igj,k+j=1nTj,k,ifj,kgj,i).(T[f,g])_{i,k}=\frac{1}{2}(\sum^{n}_{j=1}T_{j,i,k}f_{j,i}g_{j,k}+\sum^{n}_{j=1}T_{j,k,i}f_{j,k}g_{j,i}).
Remark 4.5.

For an rr-th order nn-dimensional (not necessarily symmetric) tensor TT and for s[r1]s\in[r-1], we can also consider the non-symmetrized ss-action of TT on the ss-th order nn-dimensional (not necessarily symmetric) tensors f(1),,f(r1)f^{(1)},\ldots,f^{(r-1)} is the operation

(T[f(1),,f(r1)])i1,,is=\displaystyle(T[f^{(1)},\ldots,f^{(r-1)}])_{i_{1},\ldots,i_{s}}=
j1,,jrs=1nTj1,,jrs,i1,,is\displaystyle\sum^{n}_{j_{1},\ldots,j_{r-s}=1}T_{j_{1},\ldots,j_{r-s},i_{1},\ldots,i_{s}} fi2,,is,j1(1)fi3,,is,j1,j2(2)fjr2s+2,,jrs,i1(r2s+1)fjrs,i1,,is1(r1).\displaystyle f^{(1)}_{i_{2},\ldots,i_{s},j_{1}}f^{(2)}_{i_{3},\ldots,i_{s},j_{1},j_{2}}\ldots f^{(r-2s+1)}_{j_{r-2s+2},\ldots,j_{r-s},i_{1}}\ldots f^{(r-1)}_{j_{r-s},i_{1},\ldots,i_{s-1}}.

We now introduce some notation and notions for hypergraphs.

Given an edge eEe\in E, we recall that we denote its cardinality by |e||e|, and in the following we will denote with rr the maximal edge cardinality, i.e.

r:=maxeE|e|.r:=\max_{e\in E}|e|.

Moreover, we observe that for the adjacency tensor of a hypergraph H=(V,E)H=(V,E) on nn vertices with largest edge cardinality rr its adjacency tensor

Ai1,,ir:={0 if {vi1,,vir}E1 if {vi1,,vir}E.A_{i_{1},\ldots,i_{r}}:=\begin{cases}0&\text{ if }\{v_{i_{1}},\ldots,v_{i_{r}}\}\notin E\\ 1&\text{ if }\{v_{i_{1}},\ldots,v_{i_{r}}\}\in E.\\ \end{cases}

is a standard notion for rr-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 rr-uniform hypergraph on nn vertices is the hypergraph with [n][n] and such that EE is the set of all (nr){n\choose r} subsets of VV with cardinality rr.

Recall that graphs are the 22-uniform hypergraphs. Therefore, a random graph is a 22-uniform random hypergraph. We recall here the Erdös-Renyi random graph model.

Example 4.7 (Erdös-Renyi graph).

Consider the vertex set V=[n]V=[n] and we connect each of the possible (n2){n\choose 2} pairs independently with probability pp, i.e. following the law of independent Bernoulli random variables. This is the Erdös-Renyi random graph and we will denote it with G(n,p)G(n,p).

A very common random uniform hypergraph that we will consider is the following.

Example 4.8 (rr-uniform Erdős–Rényi random hypergraph).

We denote with G(n,p,r)G(n,p,r) the rr-uniform random hypergraph with vertex set V=[n]V=[n] and with edge set EE defined as follows: For every ee, set of vertices of cardinality rr, ee is in EE with probability pp, i.e. every edge of cardinality rr is in EE following independent Bernoulli random variables with parameter p. In the case r=2, G(n,p,2)G(n,p,2) corresponds with the Erdős–Rényi random graph G(n,p)G(n,p).

In the case p=1p=1 the rr-uniform Erdős–Rényi random hypergraph G(n,1,r)G(n,1,r) corresponds with the complete rr-uniform hypergraph.

We give also another example of uniform random hypergraph:

Example 4.9.

We denote with T(n,p)T(n,p) the random 33-uniform hypergraph constructed taking the vertex set V=[n]V=[n] and as edges the triangles of the Erdős–Rényi random graph G(n,p)G(n,p) on the same vertex set V=[n]V=[n].

We can generalize naturally this random hypergraph model

Example 4.10.

We denote with R(n,p1,pr1,r)R(n,p_{1},\ldots p_{r-1},r) the rr-uniform random hypergraph on the vertex set V=[n]V=[n] constructed inductively on rr as follows:

  • for r=2r=2 we define R(n,p,2)=G(n,p,2)R(n,p,2)=G(n,p,2).

  • for r>2r>2 we define R(n,p1,,pr1,r)R(n,p_{1},\ldots,p_{r-1},r) as the rr-uniform hypergraph constructed selecting as edges independently with probability pr1p_{r-1} the sets of rr vertices such that R(n,p1,,pr2,r1)R(n,p_{1},\ldots,p_{r-2},r-1) restricted to these rr vertices is the (r1)(r-1)-uniform complete hypergraph on rr vertices.

We notice that the random 33-uniform hypergraph T(n,p)T(n,p) is the same as the random 33-uniform hypergraph R(n,p,1,3)R(n,p,1,3).

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 rthr-th order multi-PP-operator is a multi-linear operator A:L(Ω)r1A:L^{\infty}(\Omega)^{r-1}\rightarrow L1(Ω)L^{1}(\Omega) such that the 1\infty\rightarrow 1 multi-linear operator norm

A1:=supf(i)L(Ω),f(i)0A[f(1),,f(r1)]1f(1)f(r1)\|A\|_{\infty\rightarrow 1}:=\sup_{f^{(i)}\in L^{\infty}(\Omega),\,f^{(i)}\neq 0}\frac{\|A[f^{(1)},\ldots,f^{(r-1)}]\|_{1}}{\|f^{(1)}\|_{\infty}\cdots\|f^{(r-1)}\|_{\infty}}

is finite. We will say that a multi-PP-operator AA is acting on the probability space (Ω,,)(\Omega,\mathcal{F},\mathbb{P}) if the L1L^{1} and LL^{\infty} spaces are defined on (Ω,,)(\Omega,\mathcal{F},\mathbb{P}). We denote the set of all rr-th order multi-PP-operators with r\mathcal{B}_{r} and the set of all rr-th order multi-PP-operators acting on (Ω,,)(\Omega,\mathcal{F},\mathbb{P}) with r(Ω,,)\mathcal{B}_{r}(\Omega,\mathcal{F},\mathbb{P}).

We can relate multi-PP operators and tensors in multiple ways as in the following

Example 5.2.

We can interpret the ss-action of an rr-th order symmetric tensor as a multi-PP-operator

T~:L([n]s,Sym)r1L1([n]s,Sym),\widetilde{T}:L^{\infty}([n]^{s},Sym)^{r-1}\longrightarrow L^{1}([n]^{s},Sym),

where SymSym is the symmetric σ\sigma-algebra on [n]s[n]^{s} and we consider the uniform probability measure on [n]s[n]^{s}, i.e.

({(j1,,js) s.t. (j1,,js)=(iσ(1),,iσ(s)) where σ is a permutation of [s]})\displaystyle\mathbb{P}(\{(j_{1},\ldots,j_{s})\text{ s.t.\ }(j_{1},\ldots,j_{s})=(i_{\sigma(1)},\ldots,i_{\sigma(s)})\text{ where }\sigma\text{ is a permutation of $[s]$}\})
=|{(j1,,js) s.t. (j1,,js)=(iσ(1),,iσ(s)) where σ𝒫}|ns\displaystyle=\frac{|\{(j_{1},\ldots,j_{s})\text{ s.t.\ }(j_{1},\ldots,j_{s})=(i_{\sigma(1)},\ldots,i_{\sigma(s)})\text{ where }\sigma\in\mathcal{P}\}|}{n^{s}}

for all i1,,is[n]i_{1},\ldots,i_{s}\in[n]. We just have to identify the set of ss-th order symmetric tensors with L([n]s,Sym)L1([n]s,Sym)L^{\infty}([n]^{s},Sym)\cong L^{1}([n]^{s},Sym) in the canonical way.

Remark 5.3.

One also consider the non-symmetrized ss-action as a multi-PP operator. The probability space to consider in that case is just [n]s[n]^{s} with the discrete σ\sigma-algebra and the uniform probability measure on [n]s.[n]^{s}.

For functions f1(1),,f1(r1),,fk(1),,fk(r1)L(Ω)f^{(1)}_{1},\ldots,f^{(r-1)}_{1},\ldots,f^{(1)}_{k},\ldots,f^{(r-1)}_{k}\in L^{\infty}(\Omega) we consider the rkrk- dimensional random vector

(f1(1),,f1(r1),A[f1(1),,f1(r1)],,fk(1),,fk(r1),A[fk(1),,fk(r1)])(f^{(1)}_{1},\ldots,f^{(r-1)}_{1},A[f^{(1)}_{1},\ldots,f^{(r-1)}_{1}],\ldots,f^{(1)}_{k},\ldots,f^{(r-1)}_{k},A[f^{(1)}_{k},\ldots,f^{(r-1)}_{k}])

for a multi-PP-operator AA and we call the distribution of this random vector,

(f1(1),,f1(r1),A[f1(1),f1(r1)],,fk(1),,fk(r1),A[fk(1),,fk(r1)]),\displaystyle\mathcal{L}(f^{(1)}_{1},\ldots,f^{(r-1)}_{1},A[f^{(1)}_{1},\ldots f^{(r-1)}_{1}],\ldots,f^{(1)}_{k},\ldots,f^{(r-1)}_{k},A[f^{(1)}_{k},\ldots,f^{(r-1)}_{k}]), (2)

which is a probability measure in 𝒫(rk),\mathcal{P}(\mathbb{R}^{rk}), the measure generated by the multi-PP-operator AA through the ordered sequence of functions f1(1),f^{(1)}_{1},,\ldots,f1(r1),f^{(r-1)}_{1},,fk(1),,fk(r1)\ldots,f^{(1)}_{k},\ldots,f^{(r-1)}_{k}L(Ω)\in L^{\infty}(\Omega). Sometimes, we will use the abbreviation

𝒟A(f1(1),,f1(r1),fk(1),,fk(r1))\displaystyle{\mathcal{D}}_{A}(f^{(1)}_{1},\ldots,f^{(r-1)}_{1}\ldots,f^{(1)}_{k},\ldots,f^{(r-1)}_{k})
=(f1(1),,f1(r1),A[f1(1),,f1(r1)],,fk(1),,fk(r1),A[fk(1),,fk(r1)]).\displaystyle=\mathcal{L}(f^{(1)}_{1},\ldots,f^{(r-1)}_{1},A[f^{(1)}_{1},\ldots,f^{(r-1)}_{1}],\ldots,f^{(1)}_{k},\ldots,f^{(r-1)}_{k},A[f^{(1)}_{k},\ldots,f^{(r-1)}_{k}]).

We now define the set of measures generated by AA. Similarly to the action convergence in the linear case, it is convenient to allow in our sets only measures generated by functions in L[1,1](Ω)L_{[-1,1]}^{\infty}(\Omega), i.e. functions taking values between 1-1 and +1+1 almost everywhere. Therefore, we define the kk-profile of AA, 𝒮k(A){\mathcal{S}}_{k}(A), as the set of measures generated by AA by functions in L[1,1](Ω)L_{[-1,1]}^{\infty}(\Omega).

𝒮k(A)=f1(1),,f1(r1),,fk(1),,fk(r1)L[1,1](Ω)\displaystyle{\mathcal{S}}_{k}(A)=\bigcup_{f^{(1)}_{1},\ldots,f^{(r-1)}_{1},\ldots,f^{(1)}_{k},\ldots,f^{(r-1)}_{k}\in L_{[-1,1]}^{\infty}(\Omega)} {(f1(1),,f1(r1),A[f1(1),f1(r1)],,\displaystyle\{\mathcal{L}(f^{(1)}_{1},\ldots,f^{(r-1)}_{1},A[f^{(1)}_{1},\ldots f^{(r-1)}_{1}],\ldots,
fk(1),,fk(r1),A[fk(1),,fk(r1)])}\displaystyle f^{(1)}_{k},\ldots,f^{(r-1)}_{k},A[f^{(1)}_{k},\ldots,f^{(r-1)}_{k}])\}
=f1(1),,f1(r1),,fk(1),,fk(r1)L[1,1](Ω)𝒟A(f1(1),,f1(r1),fk(1),,fk(r1)).\quad\quad=\bigcup_{f^{(1)}_{1},\ldots,f^{(r-1)}_{1},\ldots,f^{(1)}_{k},\ldots,f^{(r-1)}_{k}\in L_{[-1,1]}^{\infty}(\Omega)}{\mathcal{D}}_{A}(f^{(1)}_{1},\ldots,f^{(r-1)}_{1}\ldots,f^{(1)}_{k},\ldots,f^{(r-1)}_{k}).

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 𝒫(rk)\mathcal{P}(\mathbb{R}^{rk}) (equipped with the Levy-Prokhorov metric d𝒫d_{\mathcal{LP}} (Definition 3.3)), that we will denote with dH.d_{H}.

Remark 5.4.

This is a generalization of the construction in Section 3, see (1). We use a similar notation and terminology to underline the connection with action convergence [6].

We are now ready to define the pseudo-metric we are interested in. Consider two multi-PP-operators

A:L(Ω1)r1L1(Ω1)A:L^{\infty}(\Omega_{1})^{r-1}\rightarrow L^{1}(\Omega_{1})

and

B:L(Ω2)r1L1(Ω2).B:L^{\infty}(\Omega_{2})^{r-1}\rightarrow L^{1}(\Omega_{2}).
Definition 5.5 (Metrization of action convergence).

For the two rthr-th order multi-PP-operators A,BA,B the action convergence metric is

dM(A,B):=k=12kdH(𝒮k(A),𝒮k(B)).d_{M}(A,B):=\sum_{k=1}^{\infty}2^{-k}d_{H}\left({\mathcal{S}}_{k}(A),{\mathcal{S}}_{k}(B)\right).
Remark 5.6.

This is a generalization of the action convergence metric defined in Section 3. For this reason we use the same notation here. However, this metric can be applied to multi-linear operators differently from the metric defined in Section 3.

As the Hausdorff metric dHd_{H} is bounded by 1, we have that also the action convergence distance is bounded by 1.

We will say that a sequence of PP-operators {Air(Ωi)}i=1\left\{A_{i}\in\mathcal{B}_{r}\left(\Omega_{i}\right)\right\}_{i=1}^{\infty} is a Cauchy sequence if the sequence is Cauchy in dMd_{M}.

We notice that a sequence {Air(Ωi)}i=1\left\{A_{i}\in\mathcal{B}_{r}\left(\Omega_{i}\right)\right\}_{i=1}^{\infty} is a Cauchy sequence if and only if for every kk\in\mathbb{N} the sequence {𝒮k(Ai)}i=1\left\{{\mathcal{S}}_{k}\left(A_{i}\right)\right\}_{i=1}^{\infty} is a Cauchy sequence in dHd_{H}.

Remark 5.7.

The completeness of (𝒫(k),d𝒫)\left(\mathcal{P}\left(\mathbb{R}^{k}\right),d_{\mathcal{LP}}\right) implies that the induced Hausdorff topology is also complete [21]. Therefore, a sequence {Ai}i=1\left\{A_{i}\right\}_{i=1}^{\infty} is a Cauchy sequence if and only if for every kk\in\mathbb{N} there is a closed set of measures XkX_{k} such that limidH(𝒮k(Ai),Xk)=0\lim_{i\rightarrow\infty}d_{H}\left({\mathcal{S}}_{k}\left(A_{i}\right),X_{k}\right)=0.

The following lemma is an equivalent of Lemma 2.6 in [6] for multi-PP-operators and guarantees that a subsequence {𝒮k(Ai)}i=1\{{{\mathcal{S}}_{k}\left(A_{i}\right)}\}^{\infty}_{i=1} converges in dHd_{H} to a closed set of measures XkX_{k} under a uniform bound assumption on the 1\|\cdot\|_{\infty\rightarrow 1} norm.

Lemma 5.8.

Let {Ai}i=1\{A_{i}\}^{\infty}_{i=1} be a sequence of rr-th order multi-PP-operators with uniformly bounded (,1)(\infty,1)-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 μ\mu on k\mathbb{R}^{k} let τ(μ)[0,]\tau(\mu)\in[0,\infty] denote the maximal expectation of the marginals of μ\mu,

τ(μ)=max1ik(x1,x2,,xk)k|xi|𝑑μ.\tau(\mu)=\max_{1\leq i\leq k}\int_{(x_{1},x_{2},\dots,x_{k})\in\mathbb{R}^{k}}|x_{i}|\leavevmode\nobreak\ d\mu. (3)

For c+c\in\mathbb{R}^{+} and kk\in\mathbb{N} let

𝒫c(k):={μ:μ𝒫(k),τ(μ)c}.\mathcal{P}_{c}(\mathbb{R}^{k}):=\{\mu:\mu\in\mathcal{P}(\mathbb{R}^{k}),\tau(\mu)\leq c\}.

Let furthermore 𝒬c(k)\mathcal{Q}_{c}(\mathbb{R}^{k}) denote the set of closed sets in the metric space (𝒫c(k),d𝒫)(\mathcal{P}_{c}(\mathbb{R}^{k}),d_{\mathcal{LP}}).

Lemma 5.9.

The metric spaces (𝒫c(k),d𝒫)(\mathcal{P}_{c}(\mathbb{R}^{k}),d_{\mathcal{LP}}) and (𝒬c(k),dH)(\mathcal{Q}_{c}(\mathbb{R}^{k}),d_{H}) are both compact and complete metric spaces.

Proof.

Markov’s inequality gives uniform tightness in 𝒫c(k)\mathcal{P}_{c}(\mathbb{R}^{k}), which implies the compactness of (𝒫c(k),d𝒫)(\mathcal{P}_{c}(\mathbb{R}^{k}),d_{\mathcal{LP}}) 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 Ar(Ω)A\in\mathcal{B}_{r}(\Omega) and let c:=max(A1,1)c:=\max(\|A\|_{\infty\to 1},1). Then for every kk\in\mathbb{N} the closure of 𝒮k(A){\mathcal{S}}_{k}(A) with respect to d𝒫d_{\mathcal{LP}} is in 𝒬c(rk)\mathcal{Q}_{c}(\mathbb{R}^{rk}).

Proof.

Let {vi(1),,vi(r1)}i=1k\{v^{(1)}_{i},\ldots,v_{i}^{(r-1)}\}_{i=1}^{k} be a sequence of functions in L[1,1](Ω)L^{\infty}_{[-1,1]}(\Omega). We have that vi(j)1vi(j)1\|v^{(j)}_{i}\|_{1}\leq\|v^{(j)}_{i}\|_{\infty}\leq 1 for every j[r1]j\in[r-1] and A[vi(1),,vi(r1)]1A1\|A[v^{(1)}_{i},\ldots,v^{(r-1)}_{i}]\|_{1}\leq\|A\|_{\infty\to 1} holds for 1ik1\leq i\leq k. The result follows as the 11-moments of the absolute values of the coordinates in τ\tau, (3), are given by

{vi(j)1}i=1k\{\|v^{(j)}_{i}\|_{1}\}_{i=1}^{k}

for j[r1]j\in[r-1] and

{A[vi(1),,vi(r1)]1}i=1k.\{\|A[v^{(1)}_{i},\ldots,v^{(r-1)}_{i}]\|_{1}\}_{i=1}^{k}.

As in the linear case, for a sequence of multi-PP-operators, we will not be interested only in the convergence of the sequences of kk-profiles {𝒮k(Ai)}i=1\{{\mathcal{S}}_{k}(A_{i})\}^{\infty}_{i=1} but also in the existence of a multi-PP-operator as limit object. This will actually be the convergence we are interested in.

Definition 5.11 (Action convergence of multi-PP-operators).

We say that the sequence {Air(Ωi)}i=1\left\{A_{i}\in\mathcal{B}_{r}\left(\Omega_{i}\right)\right\}_{i=1}^{\infty} is action convergent to the rr-th order multi-PP-operator Ar(Ω)A\in\mathcal{B}_{r}\left(\Omega\right) if it is a Cauchy sequence and it is such that for every positive integer kk the kk-profile 𝒮k(A){\mathcal{S}}_{k}(A) is the limit of the kk-profiles sequence {𝒮k(Ai)}i\{{\mathcal{S}}_{k}(A_{i})\}^{\infty}_{i} in the Hausdorff metric dH.d_{H}. The multi-PP-operator AA is the limit of the sequence {Ai}i=1\left\{A_{i}\right\}_{i=1}^{\infty}.

Additionally, we will say that a sequence of multi-PP-operators {Air(Ωi)}i=1\left\{A_{i}\in\mathcal{B}_{r}\left(\Omega_{i}\right)\right\}_{i=1}^{\infty} is action convergent if there exists a limit multi-PP-operator.

Remark 5.12.

We will often use the following consequence of the definition of action convergence. For an action convergent sequence of operators {Ai}i=1\left\{A_{i}\right\}_{i=1}^{\infty} to a multi-PP-operator AA and for every v(1),,v(r1)L[1,1](Ω)v^{(1)},\ldots,v^{(r-1)}\in L_{[-1,1]}^{\infty}(\Omega), there are elements vi(1),,vi(r1)L[1,1](Ωi)v^{(1)}_{i},\ldots,v^{(r-1)}_{i}\in L_{[-1,1]}^{\infty}\left(\Omega_{i}\right) such that

(vi(1),,vi(r1),Ai[vi(1),,vi(r1)])\mathcal{L}\left(v^{(1)}_{i},\ldots,v^{(r-1)}_{i},A_{i}[v^{(1)}_{i},\ldots,v^{(r-1)}_{i}]\right)

weakly converges to

(v(1),,v(r1),A[v(1),,v(r1)])\mathcal{L}(v^{(1)},\ldots,v^{(r-1)},A[v^{(1)},\ldots,v^{(r-1)}])

as ii goes to infinity.

We introduce now a multi-PP-operator norm for LpL^{p} spaces that is a natural generalization of the linear operator norm

Definition 5.13 (Multi-linear operator norm).

For an rr-th order multi-PP-operator AA the multi-linear operator (p1,,pr1,q)(p_{1},\ldots,p_{r-1},q)-norm is

Ap1,,pr1q:=supf(i)L(Ω),f(i)0A[f(1),,f(r1)]qf(1)p1f(r1)pr.\|A\|_{p_{1},\ldots,p_{r-1}\rightarrow q}:=\sup_{f^{(i)}\in L^{\infty}(\Omega),\,f^{(i)}\neq 0}\frac{\|A[f^{(1)},\ldots,f^{(r-1)}]\|_{q}}{\|f^{(1)}\|_{p_{1}}\cdots\|f^{(r-1)}\|_{p_{r}}}.

We denote the set of all rr-th order multi-PP-operators with finite (p1,,pr1,q)(p_{1},\ldots,p_{r-1},q)- norm with p1,,pr1,q\mathcal{B}_{p_{1},\ldots,p_{r-1},q} and the set of all rr-th order multi-PP-operators acting on (Ω,,)(\Omega,\mathcal{F},\mathbb{P}) with finite (p1,,pr1,q)(p_{1},\ldots,p_{r-1},q)-norm with p1,,pr1,q(Ω,,)\mathcal{B}_{p_{1},\ldots,p_{r-1},q}(\Omega,\mathcal{F},\mathbb{P}).

Remark 5.14.

With an abuse of notation, we can think of a multi-PP-operator AA with bounded (p1,,pr1,q)(p_{1},\ldots,p_{r-1},q)-norm as a multi-linear bounded operator

A:Lp1(Ω)××Lpr(Ω)Lq(Ω)A:L^{p_{1}}(\Omega)\times\ldots\times L^{p_{r}}(\Omega)\rightarrow L^{q}(\Omega)

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-PP-operators with uniformly bounded (p1,,pr1,q)(p_{1},\ldots,p_{r-1},q)-norm with p1,,pr1p_{1},\ldots,p_{r-1}\neq\infty are pre-compact in the action convergence metric.

Theorem 5.15.

For C>0,C>0, p[1,)p\in[1,\infty) and q[1,]q\in[1,\infty], let {Ai}i=1\{A_{i}\}_{i=1}^{\infty} be a Cauchy sequence of rr-th order multi-PP-operators with uniformly bounded p,,pq\|\cdot\|_{p,\ldots,p\to q} norms. Then there is a multi-PP-operator AA such that limidM(Ai,A)=0\lim_{i\to\infty}d_{M}(A_{i},A)=0, and Ap,,pqsupiAip,,pqC\|A\|_{p,\ldots,p\to q}\leq\sup_{i\in\mathbb{N}}\|A_{i}\|_{p,\ldots,p\to q}\leq C. Therefore, the sequence {Ai}i=1\{A_{i}\}_{i=1}^{\infty} 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 p1,,pr1q\|\cdot\|_{p_{1},\ldots,p_{r-1}\rightarrow q} for p1,,pr1[1,)p_{1},\ldots,p_{r-1}\in[1,\infty) directly implies that we have a uniform bound on the norm p,,pq\|\cdot\|_{p,\ldots,p\to q} for p=(maxi[r1]pi)[1,)p=(\max_{i\in[r-1]}p_{i})\in[1,\infty) as

p1,,pr1qp,,pq.\|\cdot\|_{p_{1},\ldots,p_{r-1}\rightarrow q}\leq\|\cdot\|_{p,\ldots,p\to q}.

6 Construction of the limit object

For an rr-th order multi-PP-operator AA and kk\in\mathbb{N} let cl(𝒮k(A))cl({\mathcal{S}}_{k}(A)) denote the closure of 𝒮k(A){\mathcal{S}}_{k}(A) in the space (𝒫(rk),d𝒫)(\mathcal{P}(\mathbb{R}^{rk}),d_{\mathcal{LP}}).

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 {(Ωi,𝒜i,μi)}i=1\left\{\left(\Omega_{i},\mathcal{A}_{i},\mu_{i}\right)\right\}_{i=1}^{\infty} a sequence of probability spaces and let’s assume that {Ai}i=1\left\{A_{i}\right\}_{i=1}^{\infty} is a Cauchy sequence of PP-operators Aip1,,pr1,q(Ωi)A_{i}\in\mathcal{B}_{p_{1},\ldots,p_{r-1},q}\left(\Omega_{i}\right) with supiAip1,,pr1qc\sup_{i}\left\|A_{i}\right\|_{p_{1},\ldots,p_{r-1}\rightarrow q}\leq c for a fixed c+c\in\mathbb{R}^{+}. For every kk\in\mathbb{N}, we can define

Xk:=limicl(𝒮k(Ai)).X_{k}:=\lim_{i\rightarrow\infty}cl({\mathcal{S}}_{k}\left(A_{i}\right)).

We aim to construct a multi-PP-operator with kk-profile that is the limit of the kk-profiles of the operators in a given convergent sequence of operators for every fixed kk, i.e. we will prove that there is a PP-operator Ap1,,pr1,q(Ω)A\in\mathcal{B}_{p_{1},\ldots,p_{r-1},q}(\Omega) for some probability space (Ω,𝒜,μ)(\Omega,\mathcal{A},\mu) such that for every kk\in\mathbb{N} we have that

limicl(𝒮k(Ai))=cl(𝒮k(A)).\lim_{i\rightarrow\infty}cl({\mathcal{S}}_{k}\left(A_{i}\right))=cl({\mathcal{S}}_{k}(A)).

Before the technical proof, we describe the main idea. For every kk\in\mathbb{N} we consider the limit of the kk-profiles 𝒮k(Ai){\mathcal{S}}_{k}(A_{i}) of the sequence of operators AiA_{i}, 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 kk-profiles of the sequence of operators AiA_{i}. Moreover, every element in the kk-profile of AiA_{i} involves rkrk measurable functions on Ωi\Omega_{i} (in the terminology used before the measure is generated through those functions). In probabilistic language, these functions are random variables, since Ωi\Omega_{i} is a probability space. Very roughly speaking, the main idea is to take, for every kk, enough functions needed to generate enough measures (contained in the kk-profiles of the operators AiA_{i}) to approximate a dense countable subset of the limiting kk-profile. These are countably many functions for each ii. 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 Ω:=\Omega:=\mathbb{R}^{\infty}. Each coordinate function in the probability space on \mathbb{R}^{\infty} corresponds to a function involved in a kk-profile for some kk. Since every measure in the kk-profile comes from (r1)k(r-1)k functions and their kk 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 Ω\Omega 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 σ\sigma-algebra to extend the candidate limit multi-linear operator for the entire function space on Ω\Omega. 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 rr-multi-operators).

Let GG and LL be sets. We denote by F(G,L)F(G,L) the free semigroup with generator set GG and rr-multioperator set LL (freely acting on F(G,L))F(G,L)). More precisely, we have that F(G,L)F(G,L) is the smallest set of abstract words satisfying the following properties. (1) GF(G,L)G\subseteq F(G,L). (2) If w1,w2F(G,L)w_{1},w_{2}\in F(G,L), then w1w2F(G,L)w_{1}w_{2}\in F(G,L). (3) If w1,,wr1F(G,L),lLw_{1},\ldots,w_{r-1}\in F(G,L),l\in L, then l(w1,,wr1)F(G,L)l(w_{1},\ldots,w_{r-1})\in F(G,L). There is a unique length function m:F(G,L)m:F(G,L)\rightarrow\mathbb{N} such that m(g)=1m(g)=1 for gGg\in G, m(w1w2)=m(w1)+m(w2)m\left(w_{1}w_{2}\right)=m\left(w_{1}\right)+m\left(w_{2}\right) and m(l(w1,,wr1))=maxs[r1]m(ws)+1.m(l(w_{1},\ldots,w_{r-1}))=\max_{s\in[r-1]}m(w_{s})+1.

We give an example of a word in F(G,L)F(G,L) with LL set of 22-multioperators:

l3(l1(g1,g2),l2(g2,g2g3))l3(g1,g2),l_{3}\left(l_{1}\left(g_{1},g_{2}\right),l_{2}\left(g_{2},g_{2}g_{3}\right)\right)l_{3}\left(g_{1},g_{2}\right),

where g1,g2,g3Gg_{1},g_{2},g_{3}\in G and l1,l2,l3Ll_{1},l_{2},l_{3}\in L. The length of this word is max{max{1,1}+1,max{1,1+1}+1}+1+max{1,1}+1=6\max\{\max\{1,1\}+1,\max\{1,1+1\}+1\}+1+\max\{1,1\}+1=6. Note that if both GG and LL are countable sets, then also F(G,L)F(G,L) is countable.

In the first technical part of the proof, we construct a function system {vi,fL(Ωi)},\left\{v_{i,f}\in L^{\infty}\left(\Omega_{i}\right)\right\}, i,fFi\in\mathbb{N},f\in F for some countable index set FF. Later, we will construct a probability measure κ𝒫(Fr1×[r])\kappa\in\mathcal{P}\left(\mathbb{R}^{F^{r-1}\times[r]}\right) and an operator Ap1,,pr1,q(Fr1×[r],κ)A\in\mathcal{B}_{p_{1},\ldots,p_{r-1},q}\left(\mathbb{R}^{F^{r-1}\times[r]},\kappa\right) using this function system. In the end, we will show that AA is an appropriate limit object for the sequence {Ai}i=1\left\{A_{i}\right\}_{i=1}^{\infty}.

Construction of a function system: First, we define FF, the countable index set. For every kk\in\mathbb{N}, let’s consider XkXkX_{k}^{\prime}\subseteq X_{k} a dense countable subset in the metric space (Xk,d𝒫)\left(X_{k},d_{\mathcal{LP}}\right), which is separable. Let’s define G:=k=1Xk×[k]×[r1]G:=\bigcup_{k=1}^{\infty}X_{k}^{\prime}\times[k]\times[r-1], the generator set. Therefore, the index set FF will be the free semigroup generated by GG and a set of appropriate nonlinear (r1)(r-1)-multi-operators LL. For any yy\in\mathbb{Q} and z+z\in\mathbb{Q}^{+} let hy,z:[0,1]h_{y,z}:\mathbb{R}\rightarrow[0,1] be the (bounded) continuous function defined by hy,z(x)=0h_{y,z}(x)=0 for x(yz,y+z)x\notin(y-z,y+z) and hy,z(x)=1|xy|/zh_{y,z}(x)=1-|x-y|/z for xx\in (yz,y+z)(y-z,y+z). Finally, for every i,lLi\in\mathbb{N},l\in L and v1,,vr1L(Ωi)v_{1},\ldots,v_{r-1}\in L^{\infty}\left(\Omega_{i}\right) we define l(v1,,vr1):=hy,z(Ai[v1,,vr1])l(v_{1},\ldots,v_{r-1}):=h_{y,z}\circ\left(A_{i}[v_{1},\ldots,v_{r-1}]\right), where ll is indexed by the pair (y,z)×+(y,z)\in\mathbb{Q}\times\mathbb{Q}^{+}. Observe that by definition, l(v1,,vr1)1\|l(v_{1},\ldots,v_{r-1})\|_{\infty}\leq 1. Being these functions indexed by ×+\mathbb{Q}\times\mathbb{Q}^{+}, with an abuse of notation, we will denote L=×+L=\mathbb{Q}\times\mathbb{Q}^{+}. Therefore, we let F:=F(G,L)F:=F(G,L) be as in Definition 6.1 and, thus, FF is countable. Furthermore, we define the functions {vi,g}i,gG\left\{v_{i,g}\right\}_{i\in\mathbb{N},g\in G}. For every i,ki,k\in\mathbb{N}, and tXkt\in X_{k}^{\prime} let {vi,(t,j,s)}j[k],s[r1]\left\{v_{i,(t,j,s)}\right\}_{j\in[k],s\in[r-1]} be random variables in L[1,1](Ωi)L_{[-1,1]}^{\infty}\left(\Omega_{i}\right) such that the joint distribution of

(vi,(t,1,1),,vi,(t,1,(r1)),Ai[vi,(t,1,1),,vi,(t,1,(r1))],vi,(t,2,1),,\displaystyle(v_{i,(t,1,1)},\ldots,v_{i,(t,1,(r-1))},A_{i}[v_{i,(t,1,1)},\ldots,v_{i,(t,1,(r-1))}],v_{i,(t,2,1)},\ldots,
vi,(t,k,1),,vi,(t,k,(r1)),Ai[vi,(t,k,1),,vi,(t,k,(r1))])\displaystyle v_{i,(t,k,1)},\ldots,v_{i,(t,k,(r-1))},A_{i}[v_{i,(t,k,1)},\ldots,v_{i,(t,k,({r-1}))}])

converges to tt as ii goes to \infty.

At this point, we will define the functions {vi,w}i,wF\left\{v_{i,w}\right\}_{i\in\mathbb{N},w\in F} recursively to the length of the words m(w)m(w). The functions have been constructed above for words of length m(w)=1m(w)=1. Assume we have already constructed all the functions vi,wv_{i,w} with m(w)jm(w)\leq j for some jj\in\mathbb{N}. Consider a wFw\in F such that m(w)=j+1m(w)=j+1. If w=w1w2w=w_{1}w_{2} for some w1,w2Fw_{1},w_{2}\in F, then set vi,w:=vi,w1vi,w2v_{i,w}:=v_{i,w_{1}}v_{i,w_{2}}. If w=l(w1,w2,,w(r1))w=l\left(w_{1},w_{2},\ldots,w_{(r-1)}\right), then set vi,w:=l(vi,w1,vi,w2,,vi,wr1).v_{i,w}:=l\left(v_{i,w_{1}},v_{i,w_{2}},\ldots,v_{i,w_{r-1}}\right).

Construction of the probability space: Let ξi:ΩiF(r1)×[r]\xi_{i}:\Omega_{i}\rightarrow\mathbb{R}^{F^{(r-1)}\times[r]} be the map such that for f1,,f(r1)F,e[r]f_{1},\ldots,f_{(r-1)}\in F,e\in[r], and ωiΩi\omega_{i}\in\Omega_{i} the (f1,,f(r1),e)(f_{1},\ldots,f_{(r-1)},e) coordinate of ξi(ωi)\xi_{i}\left(\omega_{i}\right) is equal to

(Aie[vi,f1,,vi,f(r1)])(ωi),\left(A_{i}^{e}[v_{i,f_{1}},\ldots,v_{i,f_{(r-1)}}]\right)\left(\omega_{i}\right),

where AisA_{i}^{s} for s[r1]s\in[r-1] is defined to be the projection on the ss-th variable and Air=AiA_{i}^{r}=A_{i}. For the random variable ξi\xi_{i} we denote its distribution with κi𝒫(F(r1)×[r])\kappa_{i}\in\mathcal{P}(\mathbb{R}^{F^{(r-1)}\times[r]}), i.e. κi\kappa_{i} is the joint distribution of the functions {vi,f1}f1F,,{vi,fr1}fr1F\left\{v_{i,f_{1}}\right\}_{f_{1}\in F},\ldots,\left\{v_{i,f_{r-1}}\right\}_{f_{r-1}\in F} and {Ai[vi,f1,,vi,fr1]}f1,,f(r1)F\left\{A_{i}[v_{i,f_{1}},\ldots,v_{i,f_{r-1}}]\right\}_{f_{1},\ldots,f_{(r-1)}\in F}. Since τ(κi)c\tau\left(\kappa_{i}\right)\leq c holds (we recall the definition of τ\tau, equation (3)), there exists a strictly increasing sequence {ni}i=1\left\{n_{i}\right\}_{i=1}^{\infty} in \mathbb{N} and a probability measure κ𝒫(F(r1)×[r])\kappa\in\mathcal{P}\left(\mathbb{R}^{F^{(r-1)}\times[r]}\right) such that κni\kappa_{n_{i}} is weakly convergent to κ\kappa as ii goes to infinity. We will define Ω:=F(r1)×[r]\Omega:=\mathbb{R}^{F^{(r-1)}\times[r]} and consider Ω\Omega as a topological space, equipped with the product topology. Therefore, we constructed the probability space (Ω,𝒜,κ)(\Omega,\mathcal{A},\kappa), where the σ\sigma-algebra 𝒜\mathcal{A} is its Borel σ\sigma-algebra and κ\kappa the probability measure obtained as weak limit of the sequence κni\kappa_{n_{i}}. We remark that κ\kappa is a probability measure, as it is the weak limit of probability distributions.

Construction of the operator: We now define an operator Ap1,,p(r1),q(Ω)A\in\mathcal{B}_{p_{1},\ldots,p_{(r-1)},q}(\Omega) with the probability space Ω\Omega defined above. For (f1,,fr1,e)F(r1)×[r](f_{1},\ldots,f_{r-1},e)\in F^{(r-1)}\times[r] we denote with π(f1,,f(r1),e):F(r1)×[r]\pi_{(f_{1},\ldots,f_{(r-1)},e)}:\mathbb{R}^{F^{(r-1)}\times[r]}\rightarrow\mathbb{R} the projection to the (f1,,f(r1),e)(f_{1},\ldots,f_{(r-1)},e) coordinate. Observe that

π(f1,,f(r1),e)\displaystyle\pi_{(f_{1},\ldots,f_{(r-1)},e)} ξi\displaystyle\circ\xi_{i} (4)
=Aie[vi,f1,,vi,fr1](i,(f1,,f(r1),e)F(r1)×[r]).\displaystyle=A_{i}^{e}[v_{i,f_{1}},\ldots,v_{i,f_{r-1}}]\quad(i\in\mathbb{N},(f_{1},\ldots,f_{(r-1)},e)\in F^{(r-1)}\times[r]).

Additionally, by the definition of κ\kappa, we also notice that π(f1,,f(r1),e)L[1,1](Ω)\pi_{(f_{1},\ldots,f_{(r-1)},e)}\in L_{[-1,1]}^{\infty}(\Omega) for f1,,f(r1)Ff_{1},\ldots,f_{(r-1)}\in F and e[r1]e\in[r-1]. We want now to prove that there exists a unique (p1,,p(r1),q)(p_{1},\ldots,p_{(r-1)},q)-bounded (r1)(r-1)-th order multi-PP-operator AA from L(Ω)××L(Ω)L^{\infty}(\Omega)\times\ldots\times L^{\infty}(\Omega) to L1(Ω)L^{1}(\Omega) with Ap1,,p(r1)qc\|A\|_{p_{1},\ldots,p_{(r-1)}\rightarrow q}\leq c such that A[π(f1,,fr1,1),,π(f1,,fr1,r1)]A[\pi_{(f_{1},\ldots,f_{r-1},1)},\ldots,\pi_{(f_{1},\ldots,f_{r-1},r-1)}] =π(f1,,fr1,r)=\pi_{(f_{1},\ldots,f_{r-1},r)} holds for every f1,,fr1Ff_{1},\ldots,f_{r-1}\in F.

Lemma 6.2.

For the coordinate functions on F(r1)×[r]\mathbb{R}^{F^{(r-1)}\times[r]} the following properties hold:

  1. 1.

    If e[r1]e\in[r-1] and f1,f2Ff_{1},f_{2}\in F, then π(,,,f1f2,,,,e)=π(,,,f1,,,,e)π(,,,f2,,,,e)\pi_{\left(\cdot,\ldots,\cdot,f_{1}f_{2},\cdot,\ldots,\cdot,e\right)}=\pi_{\left(\cdot,\ldots,\cdot,f_{1},\cdot,\ldots,\cdot,e\right)}\pi_{(\cdot,\ldots,\cdot,f_{2},\cdot,\ldots,\cdot,e)} holds in L(Ω)L^{\infty}(\Omega).

  2. 2.

    If f1,fr1Ff_{1}\ldots,f_{r-1}\in F and l=(y,z)L,l=(y,z)\in L, then π(,,,l(f1,,fr1),,,,e)=hy,zπ(f1,,f(r1),r)\pi_{(\cdot,\ldots,\cdot,l(f_{1},\ldots,f_{r-1}),\cdot,\ldots,\cdot,e)}=h_{y,z}\circ\pi_{(f_{1},\ldots,f_{(r-1)},r)} holds in L(Ω)L^{\infty}(\Omega).

  3. 3.

    If as(1),as(2),,as(ds)F,λs(1),λs(2),,λs(ds)a^{(1)}_{s},a^{(2)}_{s},\ldots,a_{s}^{(d_{s})}\in F,\lambda_{s}^{(1)},\lambda_{s}^{(2)},\ldots,\lambda_{s}^{(d_{s})}\in\mathbb{R}, for every s[r1]s\in[r-1] then

    j1,j(r1)=1d1,,d(r1)λ1(j1)λ2(j2)λ(r1)(j(r1))π(a1(j1),a2(j2),,a(r1)(js),r)q\displaystyle\left\|\sum_{j_{1},\ldots j_{(r-1)}=1}^{d_{1},\ldots,d_{(r-1)}}\lambda_{1}^{(j_{1})}\lambda_{2}^{(j_{2})}\ldots\lambda_{(r-1)}^{(j_{(r-1)})}\pi_{\left(a^{(j_{1})}_{1},a^{(j_{2})}_{2},\ldots,a_{(r-1)}^{(j_{s})},r\right)}\right\|_{q}
    cj1=1d1λ1(j1)π(a1(j1),,,,1)p1jr1=1d(r1)λ(r1)(j(r1))π(,,,a1(jr1),r1)pr1.\displaystyle\leq c\left\|\sum_{j_{1}=1}^{d_{1}}\lambda_{1}^{(j_{1})}\pi_{\left(a^{(j_{1})}_{1},\cdot,\ldots,\cdot,1\right)}\right\|_{p_{1}}\ldots\left\|\sum_{j_{r-1}=1}^{d_{(r-1)}}\lambda_{(r-1)}^{(j_{(r-1)})}\pi_{\left(\cdot,\ldots,\cdot,a^{(j_{r-1})}_{1},r-1\right)}\right\|_{p_{r-1}}.
  4. 4.

    For all e[r1]e\in[r-1], the linear span of the functions {π(,,,f,,,,e)}fF\left\{\pi_{(\cdot,\ldots,\cdot,f,\cdot,\ldots,\cdot,e)}\right\}_{f\in F} is dense in the space Lpe(Ω)L^{p_{e}}(\Omega).

  5. 5.

    Assume that kk\in\mathbb{N} and tXkt\in X_{k}^{\prime}. Then (t,j,s)GF(t,j,s)\in G\subset F holds for 1jk,1\leq j\leq k, s[r1]s\in[r-1] and we have

    (π((t,1,1),,1),π((t,2,1),,1),,π((t,k,1),,1),π(,(t,1,2),,2),π(,(t,2,2),,2),,\displaystyle\mathcal{L}\left(\pi_{((t,1,1),\ldots,1)},\pi_{((t,2,1),\ldots,1)},\ldots,\pi_{((t,k,1),\ldots,1)},\pi_{(\cdot,(t,1,2),\ldots,2)},\pi_{(\cdot,(t,2,2),\ldots,2)},\ldots,\right.
    π(,(t,k,2),,2),,π(,(t,k,r1),r1),,π((t,k,1),,(t,k,r1),r))=t.\displaystyle\left.\pi_{(\cdot,(t,k,2),\ldots,2)},\ldots,\pi_{(\ldots,(t,k,r-1),r-1)},\ldots,\pi_{((t,k,1),\ldots,(t,k,r-1),r)}\right)=t.
Remark 6.3.

When functions on Ω\Omega are treated as functions in Lr(Ω)L^{r}(\Omega) for some r[1,]r\in[1,\infty], 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 μ\mu on {(x,x):x[0,1]}\{(x,x):x\in[0,1]\} which is a Borel measure on 2\mathbb{R}^{2}. The xx-coordinate function (x,y)x(x,y)\mapsto x and the yy-coordinate function (x,y)y(x,y)\mapsto y coincide in the space Lr(2,μ)L^{r}(\mathbb{R}^{2},\mu), as they agree on the support of μ\mu. 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 r[1,)r\in[1,\infty). For every vLr(Ω)v\in L^{r}(\Omega) we have that

limnvj=n2n2(j/n)hj/n,1/nvr=0.\lim_{n\rightarrow\infty}\left\|v-\sum_{j=-n^{2}}^{n^{2}}(j/n)h_{j/n,1/n}\circ v\right\|_{r}=0.

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 r[1,)r\in[1,\infty). Let {viL(Ω)}iI\left\{v_{i}\in L^{\infty}(\Omega)\right\}_{i\in I} be a system of functions for some countable index set II such that for every a,bIa,b\in I there is cIc\in I with vavb=vcv_{a}v_{b}=v_{c}. Let 𝒜0\mathcal{A}_{0} be the σ\sigma-algebra generated by the functions {vi}iI\left\{v_{i}\right\}_{i\in I}. Suppose that the constant 1 function on Ω\Omega can be approximated by a uniformly bounded family of finite linear combinations of {vi}iI\left\{v_{i}\right\}_{i\in I}. Then the LrL^{r}-closure of the linear span of {viL(Ω)}iI\left\{v_{i}\in L^{\infty}(\Omega)\right\}_{i\in I} is Lr(Ω,𝒜0,κ)L^{r}\left(\Omega,\mathcal{A}_{0},\kappa\right).

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 ii\in\mathbb{N} and f1,f2Ff_{1},f_{2}\in F, it holds that vi,f1f2=vi,f1vi,f2v_{i,f_{1}f_{2}}=v_{i,f_{1}}v_{i,f_{2}}. Therefore, by equation (4) and the continuity of π\pi, it follows that each κi\kappa_{i} is supported on the closed set

e[r1]{ω:ωFr1×[r],π(,f1f2,,e)(ω)=π(,f1,,e)(ω)π(,f2,,e)(ω)}.\bigcap_{e\in[r-1]}\left\{\omega:\omega\in\mathbb{R}^{F^{r-1}\times[r]},\pi_{\left(\ldots,f_{1}f_{2},\ldots,e\right)}(\omega)=\pi_{\left(\ldots,f_{1},\ldots,e\right)}(\omega)\pi_{\left(\ldots,f_{2},\ldots,e\right)}(\omega)\right\}.

Therefore, κ\kappa is also supported inside this set and hence the equality π(,f1f2,,e)=π(,f1,,e)π(,f2,,e)\pi_{\left(\ldots,f_{1}f_{2},\ldots,e\right)}=\pi_{\left(\ldots,f_{1},\ldots,e\right)}\pi_{\left(\ldots,f_{2},\ldots,e\right)} holds κ\kappa-almost everywhere for every e[r1]e\in[r-1].

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 ii\in\mathbb{N} and f1,fr1F,l=(y,z)Lf_{1}\ldots,f_{r-1}\in F,l=(y,z)\in L we have vi,l(f1,,fr1)=l(vi,f1,,vi,fr1)=hy,z(Ai[vi,f1,,vi,fr1])v_{i,\left.l(f_{1},\ldots,f_{r-1}\right)}=l\left(v_{i,f_{1}},\ldots,v_{i,f_{r-1}}\right)=h_{y,z}\circ\left(A_{i}[v_{i,f_{1}},\ldots,v_{i,f_{r-1}}]\right). Thus, by the definition of κi\kappa_{i}, equation (4) and the continuity of π\pi we obtain that κi\kappa_{i} is supported inside the closed set

e[r1]{ω:ωFr1×[r],π(,l(f1,,fr1),,e)(ω)=hy,z(π(f1,,fr1,r)(ω))}\bigcap_{e\in[r-1]}\left\{\omega:\omega\in\mathbb{R}^{F^{r-1}\times[r]},\pi_{(\ldots,l(f_{1},\ldots,f_{r-1}),\ldots,e)}(\omega)=h_{y,z}\left(\pi_{(f_{1},\ldots,f_{r-1},r)}(\omega)\right)\right\}

for every ii\in\mathbb{N}. Therefore, for every e[r1]e\in[r-1], the equality π(,l(f1,,fr1),,e)=hy,zπ(f1,,fr1,r)\pi_{(\ldots,l(f_{1},\ldots,f_{r-1}),\ldots,e)}=h_{y,z}\circ\pi_{(f_{1},\ldots,f_{r-1},r)} holds κ\kappa-almost everywhere.

To show the third claim, we recall that Aip1,,pr1qc\left\|A_{i}\right\|_{p_{1},\ldots,p_{r-1}\rightarrow q}\leq c holds for every ii\in\mathbb{N} and thus

j1,,j(r1)=1d1,,d(r1)λ1(j1)λ(r1)(j(r1))\displaystyle\left\|\sum_{j_{1},\ldots,j_{(r-1)}=1}^{d_{1},\ldots,d_{(r-1)}}\lambda_{1}^{(j_{1})}\ldots\lambda_{(r-1)}^{(j_{(r-1)})}\right. Ai[vi,aj1,,vi,aj(r1)]q\displaystyle\left.A_{i}[v_{i,a_{j_{1}}},\ldots,v_{i,a_{j_{(r-1)}}}]\right\|_{q}
cj1=1d1λ1(j1)vi,aj1p1j(r1)=1d(r1)λ(r1)(j(r1))vi,ajr1pr1.\displaystyle\leq c\left\|\sum_{j_{1}=1}^{d_{1}}\lambda_{1}^{(j_{1})}v_{i,a_{j_{1}}}\right\|_{p_{1}}\ldots\left\|\sum_{j_{(r-1)}=1}^{d_{(r-1)}}\lambda_{(r-1)}^{(j_{(r-1)})}v_{i,a_{j_{r-1}}}\right\|_{p_{r-1}}.

The sums in the factors on the right-hand side are functions in L(Ωi)L^{\infty}\left(\Omega_{i}\right) whose values for the respective e[r1]e\in[r-1] are in the compact intervals [λ(e),λ(e)][-\lambda^{(e)},\lambda^{(e)}] for λ(e):=je=1de|λje(e)|\lambda^{(e)}:=\sum_{j_{e}=1}^{d_{e}}\left|\lambda^{(e)}_{j_{e}}\right|, therefore, we obtain that je=1deλe(je)π(,aje,,e)\sum_{j_{e}=1}^{d_{e}}\lambda^{(j_{e})}_{e}\pi_{\left(\ldots,a_{j_{e}},\ldots,e\right)} is a bounded, continuous function on the support of κ\kappa. Therefore, using that κi\kappa_{i} converges to κ\kappa weakly and equation (4) again (in particular, integrating the pp-th power of the absolute values with respect to κi\kappa_{i}), we obtain that

limije=1deλe(je)vi,ajepe=je=1deλe(je)π(,aje,,e)pe.\lim_{i\rightarrow\infty}\left\|\sum_{j_{e}=1}^{d_{e}}\lambda_{e}^{(j_{e})}v_{i,a_{j_{e}}}\right\|_{p_{e}}=\left\|\sum_{j_{e}=1}^{d_{e}}\lambda_{e}^{(j_{e})}\pi_{\left(\ldots,a_{j_{e}},\ldots,e\right)}\right\|_{p_{e}}.

On the other side, weak convergence implies the following inequality:

j1,,j(r1)=1d1,,d(r1)\displaystyle\left\|\sum_{j_{1},\ldots,j_{(r-1)}=1}^{d_{1},\ldots,d_{(r-1)}}\right. λ1(j1)λ(r1)(j(r1))π(a1,,a(r1),r)q\displaystyle\lambda_{1}^{(j_{1})}\ldots\lambda_{(r-1)}^{(j_{(r-1)})}\pi_{(a_{1},\ldots,a_{(r-1)},r)}\Biggr{\|}_{q}
lim supij1,,j(r1)=1d1,,d(r1)λ1(j1)λ(r1)(j(r1))Ai[vi,aj1,,vi,aj(r1)]q.\displaystyle\leq\limsup_{i\rightarrow\infty}\left\|\sum_{j_{1},\ldots,j_{(r-1)}=1}^{d_{1},\ldots,d_{(r-1)}}\lambda_{1}^{(j_{1})}\ldots\lambda_{(r-1)}^{(j_{(r-1)})}A_{i}[v_{i,a_{j_{1}}},\ldots,v_{i,a_{j_{(r-1)}}}]\right\|_{q}.

as |j1,,j(r1)=1d1,,d(r1)λ1(j1)λ(r1)(j(r1))π(a1,,a(r1),r)|q\left|\sum_{j_{1},\ldots,j_{(r-1)}=1}^{d_{1},\ldots,d_{(r-1)}}\lambda_{1}^{(j_{1})}\ldots\lambda_{(r-1)}^{(j_{(r-1)})}\pi_{(a_{1},\ldots,a_{(r-1)},r)}\right|^{q} is a continuous non-negative function. Therefore, putting those inequalities together we obtain the third statement.

To prove the fourth statement, let s(e)\mathcal{H}^{(e)}_{s} be the LsL^{s}-closure of the linear span of the function system {π(f1,,f(r1),e)}f1,,f(r1)F\left\{\pi_{(f_{1},\ldots,f_{(r-1)},e)}\right\}_{f_{1},\ldots,f_{(r-1)}\in F} for e[r1]e\in[r-1] and s[1,)s\in[1,\infty).

First of all we notice that

π(f,,1)=π(,f,,e)=π(,f,r1)\pi(f,\ldots,1)=\pi(\ldots,f,\ldots,e)=\pi(\ldots,f,r-1)

for all fFf\in F and e[r1]e\in[r-1] and, therefore, s(1)==s(e)==s(r1)\mathcal{H}^{(1)}_{s}=\ldots=\mathcal{H}^{(e)}_{s}=\ldots=\mathcal{H}^{(r-1)}_{s}. From now on we will write s=s(e)\mathcal{H}_{s}=\mathcal{H}^{(e)}_{s} as it does not depend on ee.

Now we prove that π(f1,,f(r1),r)q\pi_{(f_{1},\ldots,f_{(r-1)},r)}\in\mathcal{H}_{q} holds for every f1,,f(r1)Ff_{1},\ldots,f_{(r-1)}\in F. From the second statement of the lemma, it follows that the following equality holds

j=n2n2(j/n)hj/n,1/nπ(f1,,f(r1),r)=j=n2n2(j/n)π(,lj(f1,,f(r1)),e),\sum_{j=-n^{2}}^{n^{2}}(j/n)h_{j/n,1/n}\circ\pi_{(f_{1},\ldots,f_{(r-1)},r)}=\sum_{j=-n^{2}}^{n^{2}}(j/n)\pi_{(\ldots,l_{j}(f_{1},\ldots,f_{(r-1)}),e)}, (5)

where ljl_{j} is represented by the pair (j/n,1/n)(j/n,1/n) for n2jn2-n^{2}\leq j\leq n^{2}.

Thus, we notice that the left-hand side is in q\mathcal{H}_{q} as the right-hand side of (5) obviously is in q\mathcal{H}_{q}. Moreover, π(f1,,f(r1),r)Lq(Ω)\pi_{(f_{1},\ldots,f_{(r-1)},r)}\in L^{q}(\Omega) by the third statement of the lemma. Hence, by Lemma 6.2, we have that, the left-hand side of (5) converges to π(f1,,f(r1),r)\pi_{(f_{1},\ldots,f_{(r-1)},r)} in Lq(Ω)L^{q}(\Omega), as nn goes to \infty, and hence π(f1,,fr1,r)q\pi_{(f_{1},\ldots,f_{r-1},r)}\in\mathcal{H}_{q}.

For fixed e[r1]e\in[r-1], let 𝒜0\mathcal{A}_{0} denote the σ\sigma-algebra generated by the functions {π(f1,,fr1,e)}f1,,fr1F\left\{\pi_{(f_{1},\ldots,f_{r-1},e)}\right\}_{f_{1},\ldots,f_{r-1}\in F}. Observe that already in X1X_{1}^{\prime} the constant function 11 can be approximated on Ω\Omega. Therefore, we obtain by the first statement in this lemma and Lemma 6.2 that r=Lr(Ω,𝒜0,κ)\mathcal{H}_{r}=L^{r}\left(\Omega,\mathcal{A}_{0},\kappa\right) holds for every e[r1]e\in[r-1] and rr\in [1,)[1,\infty). Thus, we obtained that for every f1,,fr1Ff_{1},\ldots,f_{r-1}\in F the equality π(f1,,fr1,r)q=Lq(Ω,𝒜0,κ)\pi_{(f_{1},\ldots,f_{r-1},r)}\in\mathcal{H}_{q}=L^{q}\left(\Omega,\mathcal{A}_{0},\kappa\right) holds and, hence, all coordinate functions on Fr1×[r]\mathbb{R}^{F^{r-1}\times[r]} are measurable in 𝒜0\mathcal{A}_{0}. This finally proves that r=Lr(Ω,𝒜0,κ)=Lr(Ω,𝒜,κ)=Lr(Ω)\mathcal{H}_{r}=L^{r}\left(\Omega,\mathcal{A}_{0},\kappa\right)=L^{r}(\Omega,\mathcal{A},\kappa)=L^{r}(\Omega) holds for every r[1,)r\in[1,\infty).

From the definition of the functions {vi,(t,j,e)}i,j[k]\left\{v_{i,(t,j,e)}\right\}_{i\in\mathbb{N},j\in[k]} and the definition of the probability measure κ\kappa, we directly obtain the last statement of the lemma. ∎

We will need also the following lemma to prove the existence of the multi-PP-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 V1,,VrV_{1},\ldots,V_{r} and UU be Banach spaces and W1,,WrW_{1},\ldots,W_{r} where, for every i[r]i\in[r], WiW_{i} is a dense subspace of ViV_{i}. For a multi-linear bounded operator

T0:W1××WrUT_{0}:W_{1}\times\ldots\times W_{r}\longrightarrow U
(x1,,xr)T[x1,,xr](x_{1},\ldots,x_{r})\mapsto T[x_{1},\ldots,x_{r}]

there exists a unique multi-linear bounded operator

T:V1××VrUT:V_{1}\times\ldots\times V_{r}\longrightarrow U

and

T0=T.\|T_{0}\|=\|T\|.
Proof.

For every (x1,,xr)V1××Vr(x_{1},\ldots,x_{r})\in V_{1}\times\ldots\times V_{r} we define

T[x1,,xr]=limnT[x1,n,,xr,n]T[x_{1},\ldots,x_{r}]=\lim_{n\rightarrow\infty}T[x_{1,n},\ldots,x_{r,n}]

where (x1,n,,xr,n)(x1,,xr)(x_{1,n},\ldots,x_{r,n})\rightarrow(x_{1},\ldots,x_{r}) as nn\rightarrow\infty where (x1,n,,xr,n)W1××Wr(x_{1,n},\ldots,x_{r,n})\in W_{1}\times\ldots\times W_{r} for every nn\in\mathbb{N} and the convergence is in the natural norm on V1××VrV_{1}\times\ldots\times V_{r}. We show that this definition is independent of the sequence we choose. We consider two sequences

(x1,n,,xr,n)(x1,,xr)(x_{1,n},\ldots,x_{r,n})\rightarrow(x_{1},\ldots,x_{r})
(y1,n,,yr,n)(x1,,xr)(y_{1,n},\ldots,y_{r,n})\rightarrow(x_{1},\ldots,x_{r})
T[x1,n,,xr,n]T[y1,n,,yr,n]\displaystyle\|T[x_{1,n},\ldots,x_{r,n}]-T[y_{1,n},\ldots,y_{r,n}]\|
T[x1,n,,xr,n]T[y1,n,yr1,n,,xr,n]+\displaystyle\leq\|T[x_{1,n},\ldots,x_{r,n}]-T[y_{1,n},y_{r-1,n},\ldots,x_{r,n}]+\ldots
+T[x1,n,y2,n,,yr,n]T[y1,n,,yr,n]\displaystyle\hskip 56.9055pt+T[x_{1,n},y_{2,n},\ldots,y_{r,n}]-T[y_{1,n},\ldots,y_{r,n}]\|
Ci=1r(j=1i1xj,n)xi,nyi,n(j=i+1ryj,n)\displaystyle\leq C\sum^{r}_{i=1}\left(\prod^{i-1}_{j=1}\|x_{j,n}\|\right)\|x_{i,n}-y_{i,n}\|\left(\prod^{r}_{j=i+1}\|y_{j,n}\|\right)
Ki=1rxi,nyi,n0\displaystyle\leq K\sum^{r}_{i=1}\|x_{i,n}-y_{i,n}\|\rightarrow 0

as n0n\rightarrow 0. Moreover,

T0=supx1W1,,xrWr,x1,,xr0T[x1,,xr]x1xr\displaystyle\|T_{0}\|=\sup_{x_{1}\in W_{1},\ldots,x_{r}\in W_{r},\ x_{1},\ldots,x_{r}\neq 0}\frac{\|T[x_{1},\ldots,x_{r}]\|}{\|x_{1}\|\ldots\|x_{r}\|}
=supx1V1,,xrVr,x1,,xr0T[x1,,xr]x1xr=T\displaystyle=\sup_{x_{1}\in V_{1},\ldots,x_{r}\in V_{r},\ x_{1},\ldots,x_{r}\neq 0}\frac{\|T[x_{1},\ldots,x_{r}]\|}{\|x_{1}\|\ldots\|x_{r}\|}=\|T\|

as the sets WiW_{i} are dense in ViV_{i} and therefore W1××WrW_{1}\times\ldots\times W_{r} is dense in V1××VrV_{1}\times\ldots\times V_{r}.

We now finally define the limit operator Ap1,,pr1,q(Ω)A\in\mathcal{B}_{p_{1},\ldots,p_{r-1},q}(\Omega). For f1,,fr1Ff_{1},\ldots,f_{r-1}\in F, let

A[π(f1,,1),,π(,fe,,e),,π(,fr1,r1)]=π(f1,,fr1,r).A[\pi_{(f_{1},\ldots,1)},\ldots,\pi_{(\ldots,f_{e},\ldots,e)},\ldots,\pi_{(\ldots,f_{r-1},r-1)}]=\pi_{(f_{1},\ldots,f_{r-1},r)}.

This defines a multi-linear operator on the linear span of {π(f1,,fr1,e)}f1,,fr1F\left\{\pi_{(f_{1},\ldots,f_{r-1},e)}\right\}_{f_{1},\ldots,f_{r-1}\in F}. This operator is bounded by the third statement of Lemma 6.2. Thus, there exists a unique continuous multi-linear extension on its Lp1××Lpr1L^{p_{1}}\times\ldots\times L^{p_{r-1}}-closure. In fact, by the fourth statement of Lemma 6.2 and Lemma 6.6, we get that there is a unique operator Ap1,,pr1,q(Ω)A\in\mathcal{B}_{p_{1},\ldots,p_{r-1},q}(\Omega) with Ap1,,pr1qc\|A\|_{p_{1},\ldots,p_{r-1}\rightarrow q}\leq c such that A[π(f1,,1),,π(,fr1,r1)]=π(f1,,fr1,r)A[\pi_{(f_{1},\ldots,1)},\ldots,\pi_{(\ldots,f_{r-1},r-1)}]=\pi_{(f_{1},\ldots,f_{r-1},r)} holds for every f1,,fr1Ff_{1},\ldots,f_{r-1}\in F.
Last part of the proof: From the last statement of Lemma 6.2 together with the equality

A[π((t,j,1),,1),,π(,(t,j,r1),r1)]=π((t,j,1),,(t,j,r1),r)A[\pi_{((t,j,1),\ldots,1)},\ldots,\pi_{(\ldots,(t,j,r-1),r-1)}]=\pi_{((t,j,1),\ldots,(t,j,r-1),r)}

we obtain that for every kk\in\mathbb{N} and tXkt\in X_{k}^{\prime} it holds t𝒮k(A)t\in{\mathcal{S}}_{k}(A). Hence, for every kk\in\mathbb{N} we directly observe that Xkcl(𝒮k(A))X_{k}\subseteq cl({\mathcal{S}}_{k}(A)). We now want to prove that Xk=cl(𝒮k(A))X_{k}=cl({\mathcal{S}}_{k}(A)) for every kk\in\mathbb{N} and thus we still need to show the converse inclusion cl(𝒮k(A))Xkcl({\mathcal{S}}_{k}(A))\subseteq X_{k}. Let kk\in\mathbb{N} and let v1,1,v2,1,,vk,1,,v1,r1,v2,r1,,vk,r1L[1,1](Ω)v_{1,1},v_{2,1},\ldots,v_{k,1},\ldots,v_{1,r-1},v_{2,r-1},\ldots,v_{k,r-1}\in L_{[-1,1]}^{\infty}(\Omega). Hence, we aim to prove that

α:=𝒟A({vj,s}j[k],s[r1])Xk.\alpha:=\mathcal{D}_{A}\left(\left\{v_{j,s}\right\}_{j\in[k],s\in[r-1]}\right)\in X_{k}.

For ε>0\varepsilon>0 arbitrary, it follows by the fourth statement of Lemma 6.2 that for some large enough natural number mm there are elements f1,s,f2,s,,fm,sFf_{1,s},f_{2,s},\ldots,f_{m,s}\in F and real numbers {λa,j,s}a[m],j[k],s[r1]\left\{\lambda_{a,j,s}\right\}_{a\in[m],j\in[k],s\in[r-1]} such that for every j[k]j\in[k] we have wj,svj,spsε\left\|w_{j,s}-v_{j,s}\right\|_{p_{s}}\leq\varepsilon, where wj,s:=w_{j,s}:= a=1m\sum_{a=1}^{m} λa,j,s\lambda_{a,j,s} π(,fa,s,,s)\pi_{\left(\ldots,f_{a,s},\ldots,s\right)} for j[k],s[r1]j\in[k],s\in[r-1].

We recall that only vectors with \infty-norm bounded by 11 are admitted in the profiles. For this reason, we will need to use a truncation function h~\tilde{h}. Let h~:[1,1]\tilde{h}:\mathbb{R}\rightarrow[-1,1] be the continuous function with h~(x)=x\tilde{h}(x)=x for x[1,1],h~(x)=1x\in[-1,1],\tilde{h}(x)=-1 for x(,1]x\in(-\infty,-1] and h~(x)=1\tilde{h}(x)=1 for x[1,)x\in[1,\infty). We notice that |wj,s(ω)vj,s(ω)||h~wj,s(ω)vj,s(ω)|\left|w_{j,s}(\omega)-v_{j,s}(\omega)\right|\geq\left|\tilde{h}\circ w_{j,s}(\omega)-v_{j,s}(\omega)\right| holds almost everywhere as vj,s1\left\|v_{j,s}\right\|_{\infty}\leq 1 for every j[k]j\in[k] and s[r1].s\in[r-1]. By wj,svj,spsε\left\|w_{j,s}-v_{j,s}\right\|_{p_{s}}\leq\varepsilon, we observe that h~wj,svj,spsε\left\|\tilde{h}\circ w_{j,s}-v_{j,s}\right\|_{p_{s}}\leq\varepsilon for j[k]j\in[k]. Therefore, using the triangle inequality we obtain

h~wj,swj,spsh~wj,svj,sps+vj,swj,sps2ε\left\|\tilde{h}\circ w_{j,s}-w_{j,s}\right\|_{p_{s}}\leq\left\|\tilde{h}\circ w_{j,s}-v_{j,s}\right\|_{p_{s}}+\left\|v_{j,s}-w_{j,s}\right\|_{p_{s}}\leq 2\varepsilon (6)

for j[k],s[r1]j\in[k],s\in[r-1]. For i,e[r1]i\in\mathbb{N},\ e\in[r-1] and j[k]j\in[k] let zi,j,s:=a=1mλa,j,svi,fa,j,sz_{i,j,s}:=\sum_{a=1}^{m}\lambda_{a,j,s}v_{i,f_{a,j,s}} and let

βi:=𝒟Ai({zi,j,s}j[k],s[r1]).\beta_{i}:=\mathcal{D}_{A_{i}}\left(\left\{z_{i,j,s}\right\}_{j\in[k],s\in[r-1]}\right).

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 κ\kappa, it follows that

β:=limiβni=𝒟A({wj,s}j[k],s[r1])\beta:=\lim_{i\rightarrow\infty}\beta_{n_{i}}=\mathcal{D}_{A}\left(\left\{w_{j,s}\right\}_{j\in[k],s\in[r-1]}\right)

holds in d𝒫d_{\mathcal{LP}}. Moreover, we have

A[vj,1,,vj,r1]A[wj,1,,wj,r1]1\displaystyle\left\|A[v_{j,1},\ldots,v_{j,r-1}]-A[w_{j,1},\ldots,w_{j,r-1}]\right\|_{1}
A[vj,1,,vj,r1]A[wj,1,,wj,r1]q\displaystyle\leq\left\|A[v_{j,1},\ldots,v_{j,r-1}]-A[w_{j,1},\ldots,w_{j,r-1}]\right\|_{q}
e=1r1c(s=1e1vj,sps)vj,ewj,epe(s=e+1r1wj,sps)\displaystyle\leq\sum^{r-1}_{e=1}c\left(\prod^{e-1}_{s=1}\|v_{j,s}\|_{p_{s}}\right)\|v_{j,e}-w_{j,e}\|_{p_{e}}\left(\prod^{r-1}_{s=e+1}\|w_{j,s}\|_{p_{s}}\right)
c(r1)maxs[r1]{vj,sps+ε}r2ε\displaystyle\leq c(r-1)\max_{s\in[r-1]}\ \{\|v_{j,s}\|_{p_{s}}+\varepsilon\}^{r-2}\varepsilon
c(r1){1+ε}r2εCε\displaystyle\leq c(r-1)\{1+\varepsilon\}^{r-2}\varepsilon\leq C\ \varepsilon

since wj,sL(Ω)w_{j,s}\in L^{\infty}(\Omega) and where the second last inequality follows from vj,s1\|v_{j,s}\|_{\infty}\leq 1. From Lemma 11.4 we have that d𝒫(α,β)(rk)3/4(Cε)1/2d_{\mathcal{LP}}(\alpha,\beta)\leq(rk)^{3/4}\left(C^{\prime}\varepsilon\right)^{1/2}, where C:=C^{\prime}:= max(C,1)\max(C,1). Let

βi:=𝒟Ai({h~zi,j,s}j[k],s[r1]).\beta_{i}^{\prime}:=\mathcal{D}_{A_{i}}\left(\left\{\tilde{h}\circ z_{i,j,s}\right\}_{j\in[k],s\in[r-1]}\right).

Observe that the function

f:f:\mathbb{R}\longrightarrow\mathbb{R}
f(x)=h~(x)xf(x)=\tilde{h}(x)-x

is continuous. Moreover, the functions zi,j,sz_{i,j,s} all take values in the compact interval [mλ~,mλ~][-m\tilde{\lambda},m\tilde{\lambda}] where λ~=maxa[m],j[k],s[r1]|λa,j,s|\tilde{\lambda}=\max_{a\in[m],j\in[k],s\in[r-1]}{|\lambda_{a,j,s}|}. Therefore, it follows that

f(zi,j,s)psf(wj,s)ps2ε\|f(z_{i,j,s})\|_{p_{s}}\rightarrow\|f(w_{j,s})\|_{p_{s}}\leq 2\varepsilon

for ii\rightarrow\infty as ff is continuous, zi,j,sz_{i,j,s} converge in distribution to wj,sw_{j,s}, zi,j,sz_{i,j,s} are uniformly bounded and the inequality in (6). Hence, if ii is large enough, then f(zi,j,s)ps=h~zi,j,szi,j,sps3ε\left\|f(z_{i,j,s})\right\|_{p_{s}}=\left\|\tilde{h}\circ z_{i,j,s}-z_{i,j,s}\right\|_{p_{s}}\leq 3\varepsilon holds for j[k]j\in[k] and therefore d𝒫(βi,βi)d_{\mathcal{LP}}\left(\beta_{i}^{\prime},\beta_{i}\right)\leq (rk)3/4(3Cε)1/2(rk)^{3/4}\left(3C^{\prime}\varepsilon\right)^{1/2} by Lemma 11.4.

We choose now {ni}i=1\left\{n_{i}^{\prime}\right\}_{i=1}^{\infty} to be a subsequence of {ni}i=1\left\{n_{i}\right\}_{i=1}^{\infty} such that β:=limiβni\beta^{\prime}:=\lim_{i\rightarrow\infty}\beta_{n_{i}^{\prime}}^{\prime} exists. Noticing that βXk\beta^{\prime}\in X_{k} and d𝒫(β,β)(rk)3/4(3Cε)1/2d_{\mathcal{LP}}\left(\beta^{\prime},\beta\right)\leq(rk)^{3/4}\left(3C^{\prime}\varepsilon\right)^{1/2}, we get that

d𝒫(Xk,α)d𝒫(β,α)d𝒫(β,β)+d𝒫(β,α)3(rk)3/4(Cε)1/2.d_{\mathcal{LP}}\left(X_{k},\alpha\right)\leq d_{\mathcal{LP}}\left(\beta^{\prime},\alpha\right)\leq d_{\mathcal{LP}}\left(\beta^{\prime},\beta\right)+d_{\mathcal{LP}}(\beta,\alpha)\leq 3(rk)^{3/4}\left(C^{\prime}\varepsilon\right)^{1/2}.

This inequality holds for arbitrary ε>0\varepsilon>0 and, hence, we finally obtain αXk\alpha\in X_{k}.

Remark 6.7.

This proof works generally for any sequence of multi-PP-operators with a uniform bound on their order. However, this proof cannot work for sequences of multi-PP-operators in which the order of the multi-PP-operators is diverging.

7 Properties of limit objects

In this section, we discuss some properties of multi-PP-operators that are preserved under action convergence.

Definition 7.1.

Let Ar(Ω)A\in\mathcal{B}_{r}(\Omega) be a multi-PP-operator.

  • AA is symmetric if

    𝔼[A[v1,,vr1]vr]=𝔼[A[vπ(1),vπ(r1)]vπ(r)]\mathbb{E}[A[v_{1},\ldots,v_{r-1}]v_{r}]=\mathbb{E}[A[v_{\pi(1)}\ldots,v_{\pi(r-1)}]v_{\pi(r)}]

    holds for every v1,,vrL(Ω)v_{1},\ldots,v_{r}\in L^{\infty}(\Omega) and for every π\pi permutation of [r][r].

  • AA is positivity-preserving if for every v1,,vr1L(Ω)v_{1},\ldots,v_{r-1}\in L^{\infty}(\Omega) with v1(x),,v_{1}(x),\ldots, vr1(x)0v_{r-1}(x)\geq 0 for almost every xΩx\in\Omega, we have that (A[v1,,vr1])(x)0(A[v_{1},\ldots,v_{r-1}])(x)\geq 0 holds for almost every xΩx\in\Omega.

  • AA is cc-regular if A[1Ω,,1Ω]=c1ΩA[1_{\Omega},\ldots,1_{\Omega}]=c1_{\Omega} for some cc\in\mathbb{R}.

  • AA is a hypergraphop if it is positivity-preserving and symmetric.

  • AA is atomless if Ω\Omega is atomless.

In particular, we notice that the ss-action of the adjacency tensor of a hypergraph is positivity-preserving and symmetric, i.e. a hypergraphop.

Remark 7.2.

The cc-regularity property of a multi-PP-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 rr-uniform hypergraph H=(V,E)H=(V,E) we define for s[r1]s\in[r-1] the ss-degree as

degs(v1,,vs)={eE:v1,,vse},\deg_{s}(v_{1},\ldots,v_{s})=\{e\in E:\ v_{1},\ldots,v_{s}\in e\},

for v1,,vsVv_{1},\ldots,v_{s}\in V pairwise distinct (compare this degree notion with (17) in the following). We observe that the ss-action of the adjacency tensor of an rr-uniform hypergraph is cc-regular if and only if the hypergraph has constant ss-degrees equal to cc.

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-PP-operators are closed with respect to dMd_{M}.

Proof.

Let’s assume Ar(Ω)A\in\mathcal{B}_{r}(\Omega) and Br(Ω2)B\in\mathcal{B}_{r}(\Omega_{2}) to be two multi-PP-operators with dM(B,A)=dd_{M}(B,A)=d. Additionally, let’s suppose Ar(Ω)A\in\mathcal{B}_{r}(\Omega) to be atomless. Therefore, there exists a random variable vL[1,1](Ω)v\in L^{\infty}_{[-1,1]}(\Omega) such that its distribution is uniform on [1,1][-1,1]. Let’s define α:=𝒟A(v,,v)\alpha:=\mathcal{D}_{A}(v,\ldots,v). As dH(𝒮1(A),𝒮1(B))2dd_{H}({\mathcal{S}}_{1}(A),{\mathcal{S}}_{1}(B))\leq 2d we have that β=𝒟B(w,w(2),,w(r1))𝒮1(B)\beta=\mathcal{D}_{B}(w,w^{(2)},\ldots,w^{(r-1)})\in{\mathcal{S}}_{1}(B) with d𝒫(β,α)3dd_{\mathcal{LP}}(\beta,\alpha)\leq 3d and thus d𝒫(α1,β1)3dd_{\mathcal{LP}}(\alpha_{1},\beta_{1})\leq 3d, where α1=(v)=Unif[1,1]\alpha_{1}=\mathcal{L}(v)=\text{Unif}_{[-1,1]} and β1=(w)\beta_{1}=\mathcal{L}(w) are the marginals of α\alpha and β\beta on the first coordinate. Thus, the distance d𝒫d_{\mathcal{LP}} between β1\beta_{1} and the uniform distribution is at most 3d3d. Therefore, the largest atom in β1\beta_{1} is at most 10d10d as by the definition of Levy-Prokhorov distance

inf{δ:β1({x0})α1(Bδ(x0))+δ}d𝒫(α1,β1)3d\inf\{\delta:\ \beta_{1}(\{x_{0}\})\leq\alpha_{1}(B_{\delta}(x_{0}))+\delta\}\leq d_{\mathcal{LP}}(\alpha_{1},\beta_{1})\leq 3d

and α1(Bδ(x0))=2δ\alpha_{1}(B_{\delta}(x_{0}))=2\delta. Hence the largest atom in Ω2\Omega_{2} has weight at most 10d=10dM(B,A)10d=10d_{M}(B,A). For this reason, if BB is the limit of atomless operators, then BB is atomless. ∎

Under uniform boundedness conditions, positivity and symmetry of multi-PP-operators are preserved under action convergence.

Lemma 7.4.

Let p[1,]p\in[1,\infty] and q(1,)q\in(1,\infty). Let {Air(Ωi)}i=1\{A_{i}\in\mathcal{B}_{r}(\Omega_{i})\}_{i=1}^{\infty} be a sequence of multi-PP-operators with a uniform bound on the (p,,p,q)(p,\ldots,p,q)-norms converging to a multi-PP-operator Ar(Ω)A\in\mathcal{B}_{r}(\Omega). If AiA_{i} is symmetric for every ii, then AA is also symmetric.

Proof.

Let π\pi be a permutation of [r].[r]. To show the statement let v1,,vrL[1,1](Ω)v_{1},\ldots,v_{r}\in L^{\infty}_{[-1,1]}(\Omega) and let μ:=𝒟A(v1,,vr)\mu:=\mathcal{D}_{A}(v_{1},\ldots,v_{r}). By the definition of action convergence, it follows that for every ii\in\mathbb{N} there exist functions vi,1,,vi,rL[1,1](Ωi)v_{i,1},\ldots,v_{i,r}\in L_{[-1,1]}^{\infty}(\Omega_{i}) such that μi:=𝒟Ai(vi,1,,vi,r,vi,π(1),,vi,π(r))\mu_{i}:=\mathcal{D}_{A_{i}}(v_{i,1},\ldots,v_{i,r},v_{i,\pi(1)},\ldots,v_{i,\pi(r)}) weakly converges to μ\mu. By Lemma 11.5, we have that 𝔼[vi,r(Ai[vi,1,,vi,r1])]\mathbb{E}[v_{i,r}(A_{i}[v_{i,1},\ldots,v_{i,r-1}])] goes to 𝔼[vr(A[v1,,vr1])]\mathbb{E}[v_{r}(A[v_{1},\ldots,v_{r-1}])] and 𝔼[vi,π(r)(Ai[vi,π(1),,vi,π(r1)])]\mathbb{E}[v_{i,\pi(r)}(A_{i}[v_{i,\pi(1)},\ldots,v_{i,\pi(r-1)}])] goes to 𝔼[vπ(r)(A[vπ(1),,\mathbb{E}[v_{\pi(r)}(A[v_{\pi(1)},\ldots, vπ(r1)])]v_{\pi(r-1)}])] as ii goes to infinity. But additionally, we notice that

𝔼[vi,r(Ai[vi,1,,vi,r1])]=𝔼[vi,π(r)(Ai[vi,π(1),,vi,π(r1)])]\mathbb{E}[v_{i,r}(A_{i}[v_{i,1},\ldots,v_{i,r-1}])]=\mathbb{E}[v_{i,\pi(r)}(A_{i}[v_{i,\pi(1)},\ldots,v_{i,\pi(r-1)}])]

and therefore

𝔼[vr(A[v1,,vr1])]=𝔼[vπ(r)(A[vπ(1),,vπ(r1)])]\mathbb{E}[v_{r}(A[v_{1},\ldots,v_{r-1}])]=\mathbb{E}[v_{\pi(r)}(A[v_{\pi(1)},\ldots,v_{\pi(r-1)}])]

This concludes the proof. ∎

Remark 7.5.

The ss-action of the adjacency tensor of a hypergraph is positive and symmetric.

Moreover, positivity-preserving and cc-regular multi-PP-operators are also closed under action convergence, under slightly different uniform boundedness conditions.

Lemma 7.6.

Let p[1,),q[1,],cp\in[1,\infty),q\in[1,\infty],c\in\mathbb{R} and let {Air(Ωi)}i=1\{A_{i}\in\mathcal{B}_{r}(\Omega_{i})\}_{i=1}^{\infty} be a sequence of multi-PP-operators with a uniform bound on the (p,,p,q)(p,\ldots,p,q)-norms converging to a PP-operator Ar(Ω)A\in\mathcal{B}_{r}(\Omega). Then we have the following two statements.

  1. 1.

    If AiA_{i} is positivity-preserving for every ii, then AA is also positivity-preserving.

  2. 2.

    If AiA_{i} is cc-regular for every ii, then AA is also cc-regular.

Proof.

To show the first statement, let v1,,vr1L[0,1](Ω)v_{1},\ldots,v_{r-1}\in L^{\infty}_{[0,1]}(\Omega). By the definition of action convergence, there is a sequence {vi,1,,\{v_{i,1},\ldots, vi,r1L[1,1](Ωi)}i=1v_{i,r-1}\in L_{[-1,1]}^{\infty}(\Omega_{i})\}_{i=1}^{\infty} such that 𝒟Ai(vi,1,,vi,r1)\mathcal{D}_{A_{i}}(v_{i,1},\ldots,v_{i,r-1}) weakly converges to 𝒟A(v1,,vr1)\mathcal{D}_{A}(v_{1},\ldots,v_{r-1}) as ii goes to infinity. As (vi,1,,vi,r1)\mathcal{L}(v_{i,1},\ldots,v_{i,r-1}) weakly converges to the non-negative distribution (v1,,vr1)\mathcal{L}(v_{1},\ldots,v_{r-1}) it follows that (vi,1|vi,1|,,vi,r1|vi,r1|)\mathcal{L}(v_{i,1}-|v_{i,{1}}|,\ldots,v_{i,r-1}-|v_{i,{r-1}}|) weakly converges to δ0\delta_{0}. Thus, by Lemma 11.6, we have that

d𝒫(𝒟Ai(vi,1,,vi,r1),𝒟Ai(|vi,1|,,|vi,r1|))0d_{\mathcal{LP}}(\mathcal{D}_{A_{i}}(v_{i,1},\ldots,v_{i,r-1}),\mathcal{D}_{A_{i}}(|v_{i,1}|,\ldots,|v_{i,r-1}|))\rightarrow 0

for ii\rightarrow\infty and, for this reason, 𝒟Ai(|vi,1|,,|vi,r1|)\mathcal{D}_{A_{i}}(|v_{i,1}|,\ldots,|v_{i,r-1}|) weakly converges to the probability measure (v1,,\mathcal{L}(v_{1},\ldots, vr1,A[v1,,vr1])v_{r-1},A[v_{1},\ldots,v_{r-1}]). The fact that Ai[|vi,1|,,|vi,r1|]A_{i}[|v_{i,1}|,\ldots,|v_{i,r-1}|] is non-negative for every ii directly implies that A[v1,,vr1]A[v_{1},\ldots,v_{r-1}] is non-negative.

To show the second statement, let vi,1,,vi,r1L[1,1](Ωi)v_{i,1},\ldots,v_{i,r-1}\in L_{[-1,1]}^{\infty}(\Omega_{i}) be a sequence of functions such that 𝒟Ai(vi,1,,vi,r1)\mathcal{D}_{A_{i}}(v_{i,1},\ldots,v_{i,r-1}) weakly converges to 𝒟A(1Ω,,1Ω)\mathcal{D}_{A}(1_{\Omega},\ldots,1_{\Omega}). We notice that 𝒟(vi,11Ωi,,vi,r11Ωi)\mathcal{D}(v_{i,1}-1_{\Omega_{i}},\ldots,v_{i,r-1}-1_{\Omega_{i}}) weakly converges to δ0\delta_{0} and, for this reason, by Lemma 11.6 we have that

d𝒫(𝒟Ai(1Ωi,,1Ωi),𝒟Ai(vi,1,,vi,r1))0d_{\mathcal{LP}}(\mathcal{D}_{A_{i}}(1_{\Omega_{i}},\ldots,1_{\Omega_{i}}),\mathcal{D}_{A_{i}}(v_{i,1},\ldots,v_{i,r-1}))\rightarrow 0

as ii\rightarrow\infty. Hence, it follows that 𝒟A(1Ω,,1Ω)\mathcal{D}_{A}(1_{\Omega},\ldots,1_{\Omega}) is the weak limit of 𝒟Ai(1Ωi,,1Ωi)\mathcal{D}_{A_{i}}(1_{\Omega_{i}},\ldots,1_{\Omega_{i}}). The result directly follows now by the fact that Ai[1Ωi,,1Ωi]=c1ΩiA_{i}[1_{\Omega_{i}},\ldots,1_{\Omega_{i}}]=c1_{\Omega_{i}}. ∎

Remark 7.7.

The ss-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-PP-operators.

The following two lemmas are generalizations of Lemmas 2.12 and 2.13 in [6].

Lemma 8.1.

Let rr and kk be positive integers and let A,BA,B be rr-th order multi-PP-operators both in r(Ω)\mathcal{B}_{r}(\Omega) for some probability space (Ω,𝒜,μ).(\Omega,\mathcal{A},\mu). Then

dH(𝒮k(A),𝒮k(B))AB11/2(2k)3/4.d_{H}({\mathcal{S}}_{k}(A),{\mathcal{S}}_{k}(B))\leq\|A-B\|_{\infty\to 1}^{1/2}(2k)^{3/4}.
Proof.

Let μ𝒮k(A)\mu\in{\mathcal{S}}_{k}(A) be arbitrary. We have that there exist functions v1,v2,,vkL[1,1](Ω)v_{1},v_{2},\dots,v_{k}\in L^{\infty}_{[-1,1]}(\Omega) such that μ\mu is equal to the probability measure 𝒟A({vi(j)}j[r],i[k])\mathcal{D}_{A}(\{v^{(j)}_{i}\}_{j\in[r],i\in[k]}). Let ν=𝒟B({vi(j)}j[r],i[k])𝒮k(B)\nu=\mathcal{D}_{B}(\{v^{(j)}_{i}\}_{j\in[r],i\in[k]})\in{\mathcal{S}}_{k}(B). Since

A[vi(1),,vi(r1)]B[vi(1),,vi(r1)]1AB1j[r1]vi(j)AB1\|A[v^{(1)}_{i},\ldots,v^{(r-1)}_{i}]-B[v^{(1)}_{i},\ldots,v^{(r-1)}_{i}]\|_{1}\leq\|A-B\|_{\infty\to 1}\prod_{j\in[r-1]}\|v^{(j)}_{i}\|_{\infty}\leq\|A-B\|_{\infty\to 1}

holds for every i[k]i\in[k], we have by Lemma 11.4 that d𝒫(μ,ν)AB11/2(2k)3/4d_{\mathcal{LP}}(\mu,\nu)\leq\|A-B\|_{\infty\to 1}^{1/2}(2k)^{3/4}. We obtained that

supμ𝒮k(A)infν𝒮k(B)d𝒫(μ,ν)AB11/2(2k)3/4.\sup_{\mu\in{\mathcal{S}}_{k}(A)}\inf_{\nu\in{\mathcal{S}}_{k}(B)}d_{\mathcal{LP}}(\mu,\nu)\leq\|A-B\|_{\infty\to 1}^{1/2}(2k)^{3/4}.

By switching the roles of AA and BB and repeating the same argument we get the above inequality with AA and BB switched. This implies the statement of the lemma. ∎

The following lemma is a direct consequence of Lemma 8.1.

Lemma 8.2.

Assume that A,BA,B are rr-th order multi-PP-operators acting on the same space L(Ω)L^{\infty}(\Omega). We have dM(A,B)3AB11/2d_{M}(A,B)\leq 3\|A-B\|_{\infty\to 1}^{1/2}.

Proof.

Using Lemma 8.1 we obtain that

dM(A,B)AB11/2k=12k(2k)3/43AB11/2.d_{M}(A,B)\leq\|A-B\|_{\infty\to 1}^{1/2}\sum_{k=1}^{\infty}2^{-k}(2k)^{3/4}\leq 3\|A-B\|_{\infty\to 1}^{1/2}.

For a multi-PP-operator A(Ω)A\in{\mathcal{B}}(\Omega) we define the multi cut norm as

A,multi=\displaystyle\|A\|_{\square,\text{multi}}= supf(1),,f(r)L[0,1](Ω)|f(r),A[f(1),f(r1)]|.\displaystyle\sup_{f^{(1)},\ldots,f^{(r)}\in L_{[0,1]}^{\infty}(\Omega)}\left|\left\langle f^{(r)},A[f^{(1)},\ldots f^{(r-1)}]\right\rangle\right|.

We obtain that for an rr-th order multi-PP-operator the 1\infty\rightarrow 1 norm is equivalent to the multi cut norm. This is a generalization of Lemma 8.11 in [34] for graphons.

Lemma 8.3.

Let AA be a multi-PP-operator A(Ω).A\in{\mathcal{B}}(\Omega). The following inequality holds:

A,multiA12rA,multi.\|A\|_{\square,\text{multi}}\leq\|A\|_{\infty\rightarrow 1}\leq 2^{r}\|A\|_{\square,\text{multi}}.
Proof.

We get the first inequality by definition:

A1\displaystyle\left\|A\right\|_{\infty\rightarrow 1} =supf(1),,f(r1)L[1,1](Ω)A[f(1),f(r1)]1\displaystyle=\sup_{f^{(1)},\ldots,f^{(r-1)}\in L_{[-1,1]}^{\infty}(\Omega)}\left\|A[f^{(1)},\ldots f^{(r-1)}]\right\|_{1}
=supf(1),,f(r)L[1,1](Ω)f(r),A[f(1),f(r1)]\displaystyle=\sup_{f^{(1)},\ldots,f^{(r)}\in L_{[-1,1]}^{\infty}(\Omega)}\left\langle f^{(r)},A[f^{(1)},\ldots f^{(r-1)}]\right\rangle
=supf(1),,f(r)L[1,1](Ω)|f(r),A[f(1),f(r1)]|\displaystyle=\sup_{f^{(1)},\ldots,f^{(r)}\in L_{[-1,1]}^{\infty}(\Omega)}\left|\left\langle f^{(r)},A[f^{(1)},\ldots f^{(r-1)}]\right\rangle\right|
A,multi.\displaystyle\geq\|A\|_{\square,\text{multi}}.

We now show the second inequality. We observe the following equality:

A1=supf(1),,f(r),g(1),,g(r)L[0,1](Ω)f(r)g(r),A[f(1)g(1),,f(r1)g(r1)].\left\|A\right\|_{\infty\rightarrow 1}=\sup_{f^{(1)},\ldots,f^{(r)},g^{(1)},\ldots,g^{(r)}\in L_{[0,1]}^{\infty}(\Omega)}\left\langle f^{(r)}-g^{(r)},A[f^{(1)}-g^{(1)},\ldots,f^{(r-1)}-g^{(r-1)}]\right\rangle.

Moreover, for any f(1),,f(r),g(1),,g(r)L[0,1](Ω)f^{(1)},\ldots,f^{(r)},g^{(1)},\ldots,g^{(r)}\in L_{[0,1]}^{\infty}(\Omega) we have the following inequality

f(r)g(r),A[f(1)g(1),,f(r1)g(r1)]\displaystyle\left\langle f^{(r)}-g^{(r)},A[f^{(1)}-g^{(1)},\ldots,f^{(r-1)}-g^{(r-1)}]\right\rangle
=f(r),A[f(1)g(1),,f(r1)g(r1)]g(r),A[f(1)g(1),,f(r1)g(r1)]\displaystyle=\left\langle f^{(r)},A[f^{(1)}-g^{(1)},\ldots,f^{(r-1)}-g^{(r-1)}]\right\rangle-\left\langle g^{(r)},A[f^{(1)}-g^{(1)},\ldots,f^{(r-1)}-g^{(r-1)}]\right\rangle
=f(r),A[f(1),f(2)g(2),,f(r1)g(r1)]\displaystyle=\left\langle f^{(r)},A[f^{(1)},f^{(2)}-g^{(2)},\ldots,f^{(r-1)}-g^{(r-1)}]\right\rangle
f(r),A[g(1),f(2)g(2),,f(r1)g(r1)]\displaystyle\hskip 14.22636pt-\left\langle f^{(r)},A[g^{(1)},f^{(2)}-g^{(2)},\ldots,f^{(r-1)}-g^{(r-1)}]\right\rangle
g(r),A[f(1),f(2)g(2),,f(r1)g(r1)]\displaystyle\hskip 14.22636pt-\left\langle g^{(r)},A[f^{(1)},f^{(2)}-g^{(2)},\ldots,f^{(r-1)}-g^{(r-1)}]\right\rangle
+g(r),A[g(1),f(2)g(2),,f(r1)g(r1)]\displaystyle\hskip 14.22636pt+\left\langle g^{(r)},A[g^{(1)},f^{(2)}-g^{(2)},\ldots,f^{(r-1)}-g^{(r-1)}]\right\rangle
2rA,multi.\displaystyle\leq 2^{r}\|A\|_{\square,\text{multi}}.

Therefore, we obtain A12rA,multi.\left\|A\right\|_{\infty\rightarrow 1}\leq 2^{r}\|A\|_{\square,\text{multi}}.

Let φ:ΩΩ\varphi:\Omega\to\Omega be a bijective measure-preserving transformation. We denote with φ1\varphi^{-1} its measure-preserving inverse. The transformation φ\varphi induces a natural, linear action on L1(Ω)L^{1}(\Omega), which we also indicate by φ\varphi, defined by (f)φ(x)=f(φ(x))(f)^{\varphi}(x)=f(\varphi(x)). Furthermore, for A(Ω)A\in\mathcal{B}(\Omega) let AφA^{\varphi} defined as

Aφ[f(1),,f(r1)]=(A[(f(1))φ,,(f(r1))φ])φ1.A^{\varphi}[f^{(1)},\ldots,f^{(r-1)}]=(A[(f^{(1)})^{\varphi},\ldots,(f^{(r-1)})^{\varphi}])^{\varphi^{-1}}.

We observe that if A(Ω)A\in\mathcal{B}(\Omega), then Aφ(Ω)A^{\varphi}\in\mathcal{B}(\Omega) and dM(A,Aφ)=0d_{M}(A,A^{\varphi})=0. Let AA and BB be two rr-th order multi-PP-operators such that A,B(Ω).A,B\in{\mathcal{B}}(\Omega). The multi cut distance between AA and BB is defined as

δ,multi(A,B)=infφ,ψAφBψ,\delta_{\square,\text{multi}}(A,B)=\inf_{\varphi,\psi}\leavevmode\nobreak\ \|A^{\varphi}-B^{\psi}\|_{\square},

where the infimum is taken over all φ,ψ\varphi,\psi invertible measure-preserving transformations from Ω\Omega to Ω.\Omega.

Lemma 8.4.

Assume that A,BA,B are rr-th order multi-PP-operators acting on the same space Ω\Omega. Then dM(A,B)32rδ,multi(A,B)1/2d_{M}(A,B)\leq{\color[rgb]{0,0,0}3\cdot 2^{r}}\delta_{\square,\text{multi}}(A,B)^{1/2}.

Proof.

The proof follows directly from Lemma 8.2 and observing that dM(A,B)=dM(Aϕ,Bψ)d_{M}(A,B)=d_{M}(A^{\phi},B^{\psi}) for any bijective and measure preserving transformations ϕ,ψ\phi,\psi from Ω\Omega 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 33-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 ss-action of a tensor from Section 4. For a 33rd-order tensor T=(Ti,j,k)i,j,k[n]T=(T_{i,j,k})_{i,j,k\in[n]} the 1-action and the 22-action of the tensor are respectively

(T1[f,g])i=j,k=1nTj,i,kfigk(T_{1}[f,g])_{i}=\sum^{n}_{j,k=1}T_{j,i,k}f_{i}g_{k}

and

(T2[f,g])i,k=12(j=1nTj,i,kfj,igj,k+j=1nTj,i,kfj,kgj,i)\displaystyle(T_{2}[f,g])_{i,k}=\frac{1}{2}(\sum^{n}_{j=1}T_{j,i,k}f_{j,i}g_{j,k}+\sum^{n}_{j=1}T_{j,i,k}f_{j,k}g_{j,i})

Therefore, we can interpret the 11-action as an operator

T1:(L([n]))2L1([n])\displaystyle T_{1}:(L^{\infty}([n]))^{2}\longrightarrow L^{1}([n])

and the 22-action as

T2:(L([n]×[n],Sym))2L1([n]×[n],Sym)T_{2}:(L^{\infty}([n]\times[n],Sym))^{2}\longrightarrow L^{1}([n]\times[n],Sym)

where SymSym is the symmetric σ\sigma-algebra on [n]×[n][n]\times[n].

Remark 9.1.

More generally the ss-action of an rr-th order symmetric tensor TT is acting on r1r-1 symmetric ss-th order tensors and gives as an output another symmetric ss-th order tensor. For this reason, this ss-action can be interpreted as an operator

Ts:(L([n]s,Sym))r1L1([n]s,Sym)T_{s}:(L^{\infty}([n]^{s},Sym))^{r-1}\longrightarrow L^{1}([n]^{s},Sym)

where SymSym is the symmetric σ\sigma-algebra on [n]s[n]^{s}.

In such a way, we obtain two notions of convergence for sequences of 33rd-order tensors Tn=(Ti,j,k)i,j,k[n]T_{n}=(T_{i,j,k})_{i,j,k\in[n]}, the one obtained by the action convergence of the sequence of multi-linear operators (T1)n(T_{1})_{n} and the one obtained by the action convergence of the sequence of multi-linear operators (T2)n(T_{2})_{n}.

Remark 9.2.

As already pointed out we can associate to an rr-th order symmetric tensor its ss-action for s[r1]s\in[r-1]. These different actions can be interpreted as r1r-1 different multi-PP-operators. Therefore, for a sequence of rr-th order symmetric tensors, we obtain r1r-1 different notions of convergence.

We will use the results in this section later in this work to connect the metric dMd_{M} with other norms and metrics for hypergraph limits.

9.1 Uniform bounds on sequences of ss-actions

We recall that in the case of graphs, we typically consider the 11-action of normalized adjacency matrices. In particular, for dense graphs, we consider

A~(G):L([n])L1([n])\widetilde{A}(G):L^{\infty}([n])\longrightarrow L^{1}([n])
(A~(G)[f])i=jA(G)i,jnfj.(\widetilde{A}(G)[f])_{i}=\sum_{j}\frac{A(G)_{i,j}}{n}f_{j}.

for a graph GG on the vertex set [n].[n]. These linear bounded operators can be easily extended to linear bounded operators

A~(G):L1([n])L([n])\widetilde{A}(G):L^{1}{([n])}\longrightarrow L^{\infty}([n])

and we have that these operators are uniformly bounded (independently by the cardinality of the vertex set nn) as

A~(G)[f]=maxi|jA(G)i,jnfj|\displaystyle\|\widetilde{A}(G)[f]\|_{\infty}=\max_{i}|\sum_{j}\frac{A(G)_{i,j}}{n}f_{j}|
maxijA(G)i,jn|fj|j|fj|n=f1.\displaystyle\leq\max_{i}\sum_{j}\frac{A(G)_{i,j}}{n}|f_{j}|\leq\sum_{j}\frac{|f_{j}|}{n}=\|f\|_{1}.

We can observe from the following example that for hypergraphs with maximal edge cardinality r>2r>2 this is not true.

Example 9.3.

For example, consider a (dense) hypergraph HH, its adjacency tensor (Ai,j,k)i,j,k[n](A_{i,j,k})_{i,j,k\in[n]} and the associated multi-PP-operator

(A~(H)[f,g])i,k=A~[f,g]i,k=12(j=1nAj,i,knfj,i,gj,k+j=1nAj,i,knfj,k,gj,i)(\widetilde{A}(H)[f,g])_{i,k}=\widetilde{A}[f,g]_{i,k}=\frac{1}{2}(\sum^{n}_{j=1}\frac{A_{j,i,k}}{n}f_{j,i},g_{j,k}+\sum^{n}_{j=1}\frac{A_{j,i,k}}{n}f_{j,k},g_{j,i})

and consider the matrices f,gf,g such

fi,j=gi,j={fi,j=0 if i,j1fi,j=1 if i=1 or j=1.f_{i,j}=g_{i,j}=\begin{cases}f_{i,j}=0&\text{ if }i,j\neq 1\\ f_{i,j}=1&\text{ if }i=1\text{ or }j=1\end{cases}.

However, we can consider a smaller extension.

Lemma 9.4.

The sequence of operators

A~(Gn):L2([n]×[n],Sym)×L2([n]×[n],Sym)L2([n]×[n],Sym)\widetilde{A}(G_{n}):L^{2}{([n]\times[n],Sym)}\times L^{2}{([n]\times[n],Sym)}\longrightarrow L^{2}([n]\times[n],Sym)
(A~(Gn)[f,g])i,k=A~[f,g]i,k=12(j=1nAj,i,knfj,i,gj,k+j=1nAj,i,knfj,k,gj,i)(\widetilde{A}(G_{n})[f,g])_{i,k}=\widetilde{A}[f,g]_{i,k}=\frac{1}{2}(\sum^{n}_{j=1}\frac{A_{j,i,k}}{n}f_{j,i},g_{j,k}+\sum^{n}_{j=1}\frac{A_{j,i,k}}{n}f_{j,k},g_{j,i})

is uniformly bounded in L2L^{2}-operator norm.

Proof.

In these spaces, we have a uniform bound, in fact

|A~[f,g]i,k|12(j=1nAj,i,kn|fj,igj,k|+j=1nAj,i,kn|fj,kgj,i|)\displaystyle|\widetilde{A}[f,g]_{i,k}|\leq\frac{1}{2}(\sum^{n}_{j=1}\frac{A_{j,i,k}}{n}|f_{j,i}g_{j,k}|+\sum^{n}_{j=1}\frac{A_{j,i,k}}{n}|f_{j,k}g_{j,i}|)
12(j=1n1n|fj,igj,k|+j=1n1n|fj,kgj,i|)\displaystyle\leq\frac{1}{2}(\sum^{n}_{j=1}\frac{1}{n}|f_{j,i}g_{j,k}|+\sum^{n}_{j=1}\frac{1}{n}|f_{j,k}g_{j,i}|)
12((j=1n1n|fj,i|2)12(j=1n1n|gj,k|2)12+(j=1n1n|fj,k|2)12(j=1n1n|gj,i|2)12)\displaystyle\leq\frac{1}{2}((\sum^{n}_{j=1}\frac{1}{n}|f_{j,i}|^{2})^{\frac{1}{2}}(\sum^{n}_{j=1}\frac{1}{n}|g_{j,k}|^{2})^{\frac{1}{2}}+(\sum^{n}_{j=1}\frac{1}{n}|f_{j,k}|^{2})^{\frac{1}{2}}(\sum^{n}_{j=1}\frac{1}{n}|g_{j,i}|^{2})^{\frac{1}{2}})

where the last inequality follows by Cauchy-Schwartz inequality. Therefore, we obtain

A~[f,g]2(1n2i,k=1|A~[f,g]i,k|2)12\displaystyle\|\widetilde{A}[f,g]\|_{2}\leq(\frac{1}{n^{2}}\sum_{i,k=1}|\widetilde{A}[f,g]_{i,k}|^{2})^{\frac{1}{2}}
12(1n2i,k=1((j=1n1n|fj,i|2)12(j=1n1n|gj,k|2)12+(j=1n1n|fj,k|2)12(j=1n1n|gj,i|2)12)2)12\displaystyle\leq\frac{1}{2}\left(\frac{1}{n^{2}}\sum_{i,k=1}\left((\sum^{n}_{j=1}\frac{1}{n}|f_{j,i}|^{2})^{\frac{1}{2}}(\sum^{n}_{j=1}\frac{1}{n}|g_{j,k}|^{2})^{\frac{1}{2}}+(\sum^{n}_{j=1}\frac{1}{n}|f_{j,k}|^{2})^{\frac{1}{2}}(\sum^{n}_{j=1}\frac{1}{n}|g_{j,i}|^{2})^{\frac{1}{2}}\right)^{2}\right)^{\frac{1}{2}}
12(1n2i,k=1(j=1n1n|fj,i|2)(1nj=1n|gj,k|2))12+(1n2i,k=1(j=1n1n|fj,k|2)(j=1n1n|gj,i|2))12\displaystyle\leq\frac{1}{2}\left(\frac{1}{n^{2}}\sum_{i,k=1}(\sum^{n}_{j=1}\frac{1}{n}|f_{j,i}|^{2})(\frac{1}{n}\sum^{n}_{j=1}|g_{j,k}|^{2})\right)^{\frac{1}{2}}+\left(\frac{1}{n^{2}}\sum_{i,k=1}(\sum^{n}_{j=1}\frac{1}{n}|f_{j,k}|^{2})(\sum^{n}_{j=1}\frac{1}{n}|g_{j,i}|^{2})\right)^{\frac{1}{2}}
12((1n2j,i=1n|fj,i|2)12(1n2j,k=1n|gj,k|2)12+(1n2j,k=1n|fj,k|2)12(1n2j,i=1n|gj,i|2)12)\displaystyle\leq\frac{1}{2}\left((\frac{1}{n^{2}}\sum^{n}_{j,i=1}|f_{j,i}|^{2})^{\frac{1}{2}}(\frac{1}{n^{2}}\sum^{n}_{j,k=1}|g_{j,k}|^{2})^{\frac{1}{2}}+(\frac{1}{n^{2}}\sum^{n}_{j,k=1}|f_{j,k}|^{2})^{\frac{1}{2}}(\frac{1}{n^{2}}\sum^{n}_{j,i=1}|g_{j,i}|^{2})^{\frac{1}{2}}\right)
=f2g2\displaystyle=\|f\|_{2}\|g\|_{2}

where in the third inequality we used Minkowski inequality.

Remark 9.5.

More in general, for r>2r>2, for a sequence of dense hypergraphs the sequence of (r1)(r-1)-actions of the relative normalized adjacency tensors cannot be extended/interpreted as a uniformly bounded sequence of linear operators from L1××L1L^{1}\times\ldots\times L^{1} to LL^{\infty}. Therefore, one has to consider them as operators from Lp1××Lpr1L^{p_{1}}\times\ldots\times L^{p_{r-1}} to LqL^{q} with p1,,pr11p_{1},\ldots,p_{r-1}\neq 1 and qq\neq\infty. 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 [0,1][0,1]. Instead, for every rr the sequence of 11-actions of the normalized adjacency matrices of dense graphs is a uniformly bounded sequence of linear operators from L1××L1L^{1}\times\ldots\times L^{1} to LL^{\infty}.

Remark 9.6.

The same estimates hold for the non-symmetrized ss-action.

9.2 Properties of ss-actions as PP-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-PP-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-PP-operator.

Lemma 9.7.

For a 33-rd order symmetric tensor T=((T)i,j,k)i,j,k[n]T=((T)_{i,j,k})_{i,j,k\in[n]} the multi-PP-operator (T)2(T)_{2} is symmetric.

Proof.

The result follows from the following equality

𝔼[(T)2[f,g]h]\displaystyle\mathbb{E}[(T)_{2}[f,g]h] =1n2i,k=1n12(j=1nTi,j,kfi,jgj,k+j=1nTi,j,kgi,jfj,k)hi,k\displaystyle=\frac{1}{n^{2}}\sum^{n}_{i,k=1}\frac{1}{2}(\sum^{n}_{j=1}T_{i,j,k}f_{i,j}g_{j,k}+\sum^{n}_{j=1}T_{i,j,k}g_{i,j}f_{j,k})h_{i,k} (7)
=1n2i,j,k=1nTi,j,kfi,jgj,khi,k\displaystyle=\frac{1}{n^{2}}\sum^{n}_{i,j,k=1}T_{i,j,k}f_{i,j}g_{j,k}h_{i,k}
=𝔼[(T)2[h,g]f]\displaystyle=\mathbb{E}[(T)_{2}[h,g]f]
=𝔼[(T)2[f,h]g].\displaystyle=\mathbb{E}[(T)_{2}[f,h]g].

Remark 9.8.

Similarly, the ss-action of a symmetric rr-th order nn-dimensional symmetric tensor TT is symmetric for every s[r1]s\in[r-1] 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 ss-actions

Recall the 11-actions introduced in 4.2 for tensors. In this section, we generalize the notion of 11-action to rr-kernels (see below) and study its properties. We will also present a natural generalization of ss-action, for s[r],s\in[r], to rr-kernels.

Let Ω\Omega be a probability space. We call a measurable function

W:ΩrW:\Omega^{r}\rightarrow\mathbb{R}

such that W1<\|W\|_{1}<\infty an rr-kernel.

We will say that an rr-kernel WW is an rr-graphon if WW takes values in [0,1].[0,1].

Remark 9.9.

This is a trivial generalization of real-valued graphons [34]. In particular, for r=2r=2 we have that the rr-graphons are the real-valued graphons.

Remark 9.10.

An rr-th order nn-dimensional tensor is an rr-kernel where Ω=[n],\Omega=[n], endowed with the uniform measure. One can also naturally represent a tensor with a rr-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 rr-kernel WW with its 11-action, the rr-th order multi-PP-operator

(W)1:L(Ω)r1L1(Ω)(W)_{1}:L^{\infty}(\Omega)^{r-1}\rightarrow L^{1}(\Omega)

defined as

(W)1[f(1),f(r1)]=Ωr1W(x1,,xk)f(1)(x1)f(r1)(xr1)dx1dxr1.(W)_{1}[f^{(1)},\ldots f^{(r-1)}]=\int_{\Omega^{r-1}}W(x_{1},\ldots,x_{k})f^{(1)}(x_{1})\ldots f^{(r-1)}(x_{r-1})\mathrm{d}x_{1}\ldots\mathrm{d}x_{r-1}.

For a kk-kernel WW we can define the 11-cut norm as

W1\displaystyle\|W\|_{\square_{1}} =supf(1),,f(r):Ω[0,1]|ΩrW(x1,,xr)f(1)(x1)f(r)(xr)dx1dxr|\displaystyle=\sup_{f^{(1)},\ldots,f^{(r)}:\Omega\rightarrow[0,1]}\left|\int_{\Omega^{r}}W(x_{1},\ldots,x_{r})f^{(1)}(x_{1})\ldots f^{(r)}(x_{r})\mathrm{d}x_{1}\ldots\mathrm{d}x_{r}\right|
=supf(1),,f(r)L[0,1](Ω)f(r),(W)1[f(1),f(r1)]=(W)1,multi\displaystyle=\sup_{f^{(1)},\ldots,f^{(r)}\in L_{[0,1]}^{\infty}(\Omega)}\left\langle f^{(r)},{\color[rgb]{0,0,0}(W)_{1}}[f^{(1)},\ldots f^{(r-1)}]\right\rangle=\|(W)_{1}\|_{\square,\text{multi}}

Compare also [45].

From Lemma 8.3 we directly obtain that for an rr-kernel WW the 1\infty\rightarrow 1 norm of the associated multi-PP-operator (W)1(W)_{1} is equivalent to the 11-cut norm.

W1(W)112rW1.\|W\|_{\square_{1}}\leq\|(W)_{1}\|_{\infty\rightarrow 1}\leq 2^{r}\|W\|_{\square_{1}}. (8)

This is a generalization of Lemma 8.11 in [34] for graphons.

For a bijective measure-preserving transformation φ:ΩΩ\varphi:\Omega\rightarrow\Omega and an rr-kernel W,W, we denote with WφW^{\varphi} the rr-kernel defined for every x1,,xrΩx_{1},\ldots,x_{r}\in\Omega as

Wφ(x1,,xr)=W(φ(x1),,φ(xr)).W^{\varphi}(x_{1},\ldots,x_{r})=W(\varphi(x_{1}),\ldots,\varphi(x_{r})).

We observe that (W)1φ=(Wφ)1.(W)_{1}^{\varphi}=(W^{\varphi})_{1}. Moreover, for two rr-kernels WW and UU on the same probability space Ω\Omega, we define the 11-cut metric

δ1(U,W)\displaystyle\delta_{\square_{1}}(U,W) =infφ,ψWφUψ1\displaystyle=\inf_{\varphi,\psi}\|W^{\varphi}-U^{\psi}\|_{\square_{1}} (9)
=infφ,ψ(W)1φ(U)1ψ,multi\displaystyle=\inf_{\varphi,\psi}\|(W)_{1}^{\varphi}-(U)_{1}^{\psi}\|_{\square,\text{multi}}
=δ,multi((U)1,(W)1).\displaystyle=\delta_{\square,\text{multi}}((U)_{1},(W)_{1}).

Therefore, from Lemma 8.4, we obtain

dM((W)1,(U)1)32rδ1(W,U)1/2.d_{M}((W)_{1},(U)_{1})\leq{\color[rgb]{0,0,0}3\cdot 2^{r}}\delta_{\square_{1}}(W,U)^{1/2}. (10)

This implies that convergence in the 11-cut metric (or 11-cut norm) of a sequence of rr-kernels implies multi-linear action convergence of the sequence of the 11-actions associated with the rr-kernels.

Remark 9.11.

Similarly, we can consider the ss-action of a symmetric rr-kernel WW as the straightforward generalization of the ss-action of a symmetric tensor to rr-kernels. For brevity, we write down explicitly only the 22-action for a symmetric 33-kernel WW that is the multi-PP-operator

(W)2:L(Ω)2L1(Ω)(W)_{2}:L^{\infty}(\Omega)^{2}\rightarrow L^{1}(\Omega)
(W)2[f,g]=12(Ω2W(x,y,z)f(x,y)g(y,z)dy+Ω2W(x,y,z)f(z,y)g(y,x)dy).(W)_{2}[f,g]=\frac{1}{2}\left(\int_{\Omega^{2}}W(x,y,z)f(x,y)g(y,z)\mathrm{d}y+\int_{\Omega^{2}}W(x,y,z)f(z,y)g(y,x)\mathrm{d}y\right).

Let’s now consider the (too) strong 22-cut norm

W2,TS\displaystyle\|W\|_{\square_{2,\text{TS}}} =supf,g,h:[0,1]2[0,1] symmetric |[0,1]3W(x,y,z)f(x,y)g(x,z)h(y,z)dxdydz|\displaystyle=\sup_{\begin{subarray}{c}f,g,h:\left[0,1]^{2}\rightarrow[0,1]\right.\\ \text{ symmetric }\end{subarray}}\left|\int_{[0,1]^{3}}W(x,y,z)f(x,y)g(x,z)h(y,z)\mathrm{d}x\mathrm{d}y\mathrm{d}z\right|
=(W)2,multi.\displaystyle=\|(W)_{2}\|_{\square,\text{multi}}.

Therefore, we can use the reasoning used for (W)1,(W)_{1}, substituting (W)1(W)_{1} with (W)2(W)_{2} and W1\|W\|_{\square_{1}} with W2,TS,\|W\|_{\square_{2,\text{TS}}}, to obtain

dM((W)2,(U)2)32rδ2,TS(W,U)1/2=32r(infφ,ψWφUψ2,TS)1/2.d_{M}((W)_{2},(U)_{2})\leq{\color[rgb]{0,0,0}3\cdot 2^{r}}\delta_{\square_{2,TS}}(W,U)^{1/2}={\color[rgb]{0,0,0}3\cdot 2^{r}}(\inf_{\varphi,\psi}\|W^{\varphi}-U^{\psi}\|_{\square_{2,\text{TS}}})^{1/2}. (11)

The (too) strong 22-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 dM,d_{M}, 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 33-uniform Erdős–Rényi hypergraph G(n,18,3)G(n,\frac{1}{8},3) from Example 4.8 and the 33-uniform hypergraph T(n,12)T(n,\frac{1}{2}), i.e. the 33-uniform hypergraph with the triangles of an Erdős–Rényi graph G(n,12,2)G(n,\frac{1}{2},2) from Example 4.9 as edges. We now consider the sequence (An)n(A_{n})_{n\in\mathbb{N}} of the normalized adjacency tensors associated with (a realization of) G(n,18,3)G(n,\frac{1}{8},3), i.e.  An=A(G(n,18,3))/nA_{n}=\nicefrac{{A(G(n,\frac{1}{8},3))}}{{n}}, and the sequence (Bn)n(B_{n})_{n\in\mathbb{N}} of the normalized adjacency tensors associated with (a realization of) T(n,12)T(n,\frac{1}{2}), i.e. Bn=A(T(n,12))/nB_{n}=\nicefrac{{A(T(n,\frac{1}{2}))}}{{n}}. 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-PP-operators (An/n)1(A_{n}/n)_{1} and (Bn/n)1(B_{n}/n)_{1} they both action converge to the same limit object, the 11-action of the constant 33-graphon W=1/8W=1/8 defined on [0,1]×[0,1]×[0,1],[0,1]\times[0,1]\times[0,1], 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 11-cut norm 1\|\cdot\|_{\square_{1}} (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 44 vertices. In T(n,12)T(n,\frac{1}{2}) there cannot be exactly three edges but in G(n,18,3)G(n,\frac{1}{8},3) this happens with probability

418181878=71024.4\cdot\frac{1}{8}\cdot\frac{1}{8}\cdot\frac{1}{8}\cdot\frac{7}{8}=\frac{7}{1024}.

Instead, if we now consider the sequences of multi-PP-operators (An)2(A_{n})_{2} and (Bn)2(B_{n})_{2} 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 22-cut norm 2,TS\|\cdot\|_{\square_{2,\text{TS}}} (see Section 3 in [45] again) that the sequence of multi-PP-operators (An)2(A_{n})_{2} converges to the 22-action of the 33-graphon W=1/8W=1/8 defined on [0,1]×[0,1]×[0,1],[0,1]\times[0,1]\times[0,1], where the unit interval is endowed with the Lebesgue measure. However, we cannot use the same method to say something about the sequence (Bn)2(B_{n})_{2} as BnB_{n} is not convergent in 2,TS.\|\cdot\|_{\square_{2,\text{TS}}}.

Lemma 9.13.

The (sub-)sequences of the multi-PP-operators (An)2(A_{n})_{2} and (Bn)2(B_{n})_{2}, as defined in Example 9.12, have different action convergence limits.

Proof.

Let’s denote with EnE_{n} the set of edges of (a realization of) the Erdős–Rényi graph G(n,12)G(n,\frac{1}{2}) from which T(n,12)T(n,\frac{1}{2}) is generated, that is the Erdős–Rényi graph from which the triangles are taken to create the edges of T(n,12)T(n,\frac{1}{2}). Let 𝟙Ωn\mathbbm{1}_{\Omega_{n}} be the n×nn\times n matrix with every entry equal to 11. We can observe that the distribution

(𝟙Ωn,𝟙Ωn,(An)2[𝟙Ωn,𝟙Ωn])\mathcal{L}(\mathbbm{1}_{\Omega_{n}},\mathbbm{1}_{\Omega_{n}},(A_{n})_{2}[\mathbbm{1}_{\Omega_{n}},\mathbbm{1}_{\Omega_{n}}])

of the 33-random vector

(𝟙Ωn,𝟙Ωn,(An)2[𝟙Ωn,𝟙Ωn])(\mathbbm{1}_{\Omega_{n}},\mathbbm{1}_{\Omega_{n}},(A_{n})_{2}[\mathbbm{1}_{\Omega_{n}},\mathbbm{1}_{\Omega_{n}}])

where

(An)2[𝟙Ωn,𝟙Ωn]i,k=j=1n(An)j,i,k(𝟙Ωn)j,i,(𝟙Ωn)j,k=j=1n(An)j,i,k(A_{n})_{2}[\mathbbm{1}_{\Omega_{n}},\mathbbm{1}_{\Omega_{n}}]_{i,k}=\sum^{n}_{j=1}(A_{n})_{j,i,k}(\mathbbm{1}_{\Omega_{n}})_{j,i},(\mathbbm{1}_{\Omega_{n}})_{j,k}=\sum^{n}_{j=1}(A_{n})_{j,i,k}

and the distribution

(𝟙Ωn,𝟙Ωn,(Bn)2[𝟙Ωn,𝟙Ωn])\mathcal{L}(\mathbbm{1}_{\Omega_{n}},\mathbbm{1}_{\Omega_{n}},(B_{n})_{2}[\mathbbm{1}_{\Omega_{n}},\mathbbm{1}_{\Omega_{n}}])

of the 33-random vector

(𝟙Ωn,𝟙Ωn,(Bn)2[𝟙Ωn,𝟙Ωn])(\mathbbm{1}_{\Omega_{n}},\mathbbm{1}_{\Omega_{n}},(B_{n})_{2}[\mathbbm{1}_{\Omega_{n}},\mathbbm{1}_{\Omega_{n}}])

where

(Bn)2[𝟙Ωn,𝟙Ωn]i,k=j=1n(Bn)j,i,k(𝟙Ωn)j,i,(𝟙Ωn)j,k=j=1n(Bn)j,i,k(B_{n})_{2}[\mathbbm{1}_{\Omega_{n}},\mathbbm{1}_{\Omega_{n}}]_{i,k}=\sum^{n}_{j=1}(B_{n})_{j,i,k}(\mathbbm{1}_{\Omega_{n}})_{j,i},(\mathbbm{1}_{\Omega_{n}})_{j,k}=\sum^{n}_{j=1}(B_{n})_{j,i,k}

are very different. In fact, for nn\rightarrow\infty we have that for any (i,k)[n]×[n]{(i,i):i[n]}(i,k)\in[n]\times[n]\setminus\{(i,i):\ i\in[n]\} for any j[n]j\in[n] the probability that {i,j,k}\{i,j,k\} is an edge of G(n,18,3)G(n,\frac{1}{8},3) is 18\frac{1}{8}. Therefore, (An)2[𝟙Ωn,𝟙Ωn]i,k(A_{n})_{2}[\mathbbm{1}_{\Omega_{n}},\mathbbm{1}_{\Omega_{n}}]_{i,k} is a sum of nn Bernoulli (almost) independent random variables with parameter 1/81/8 divided by n.n. Therefore, by (a standard argument using) the law of large numbers we obtain that

(𝟙Ωn,𝟙Ωn,(An)2[𝟙Ωn,𝟙Ωn])δ(1,1,18)\mathcal{L}(\mathbbm{1}_{\Omega_{n}},\mathbbm{1}_{\Omega_{n}},(A_{n})_{2}[\mathbbm{1}_{\Omega_{n}},\mathbbm{1}_{\Omega_{n}}])\rightarrow\delta_{(1,1,\frac{1}{8})}

However, similarly, we obtain that

(𝟙Ωn,𝟙Ωn,(Bn)2[𝟙Ωn,𝟙Ωn])12δ(1,1,0)+12δ(1,1,14)\mathcal{L}(\mathbbm{1}_{\Omega_{n}},\mathbbm{1}_{\Omega_{n}},(B_{n})_{2}[\mathbbm{1}_{\Omega_{n}},\mathbbm{1}_{\Omega_{n}}])\rightarrow\frac{1}{2}\delta_{(1,1,0)}+\frac{1}{2}\delta_{(1,1,\frac{1}{4})}

as if (i,k)En(i,k)\in E_{n} then for any j[n]j\in[n] the probability that {i,j,k}\{i,j,k\} is an edge of T(n,12)T(n,\frac{1}{2}) is 14\frac{1}{4} but if (i,k)En(i,k)\notin E_{n} then there is no j[n]j\in[n] such that {i,j,k}\{i,j,k\} is an edge of T(n,12)T(n,\frac{1}{2}). Therefore, the 11-profiles 𝒮1(A){\mathcal{S}}_{1}(A) and 𝒮1(B){\mathcal{S}}_{1}(B) of the action convergence limits AA and BB (passing to subsequences if it is necessary) of the sequences ((An)2)n((A_{n})_{2})_{n} and ((Bn)2)n((B_{n})_{2})_{n} are at Hausdorff distance bigger than a constant c>0c>0. Let’s suppose by contradiction that there exists f,gL[1,1](Ω)f,g\in L_{[-1,1]}^{\infty}(\Omega) such that for every ε>0\varepsilon>0

d𝒫((f,g,A[f,g]),(𝟙Ω,𝟙Ω,B[𝟙Ω,𝟙Ω]))ε.d_{\mathcal{LP}}(\mathcal{L}(f,g,A[f,g]),\mathcal{L}(\mathbbm{1}_{\Omega},\mathbbm{1}_{\Omega},B[\mathbbm{1}_{\Omega},\mathbbm{1}_{\Omega}]))\leq\varepsilon.

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 11 and 1-1, convergence in probability is equivalent to the convergence of the pp-th moment. Therefore, for any δ>0\delta>0, we can choose ε\varepsilon small enough such that

𝟙Ωf1𝟙Ωfp1<δ and 𝟙Ωg1𝟙Ωgp2<δ\|\mathbbm{1}_{\Omega}-f\|_{1}\leq\|\mathbbm{1}_{\Omega}-f\|_{p_{1}}<\delta\text{ and }\|\mathbbm{1}_{\Omega}-g\|_{1}\leq\|\mathbbm{1}_{\Omega}-g\|_{p_{2}}<\delta

and, therefore,

A[𝟙Ω,𝟙Ω]A[f,g]1A[𝟙Ω,𝟙Ω]A[f,g]q<2Cδ\|A[\mathbbm{1}_{\Omega},\mathbbm{1}_{\Omega}]-A[f,g]\|_{1}\leq\|A[\mathbbm{1}_{\Omega},\mathbbm{1}_{\Omega}]-A[f,g]\|_{q}<2C\delta

Using Lemma 11.3, we obtain that

d𝒫((f,g,A[f,g]),(𝟙Ω,𝟙Ω,A[𝟙Ω,𝟙Ω]))334δ12max{(2C)12,1}d_{\mathcal{LP}}(\mathcal{L}(f,g,A[f,g]),\mathcal{L}(\mathbbm{1}_{\Omega},\mathbbm{1}_{\Omega},A[\mathbbm{1}_{\Omega},\mathbbm{1}_{\Omega}]))\leq 3^{\frac{3}{4}}\delta^{\frac{1}{2}}\max\{(2C)^{\frac{1}{2}},1\}

Therefore, for the triangular inequality we have

d𝒫((f,g,A[f,g]),\displaystyle d_{\mathcal{LP}}(\mathcal{L}(f,g,A[f,g]), (𝟙Ω,𝟙Ω,B[𝟙Ω,𝟙Ω]))\displaystyle\mathcal{L}(\mathbbm{1}_{\Omega},\mathbbm{1}_{\Omega},B[\mathbbm{1}_{\Omega},\mathbbm{1}_{\Omega}]))
|d𝒫((𝟙Ω,𝟙Ω,B[𝟙Ω,𝟙Ω]),(𝟙Ω,𝟙Ω,A[𝟙Ω,𝟙Ω]))\displaystyle\geq\big{|}d_{\mathcal{LP}}(\mathcal{L}(\mathbbm{1}_{\Omega},\mathbbm{1}_{\Omega},B[\mathbbm{1}_{\Omega},\mathbbm{1}_{\Omega}]),\mathcal{L}(\mathbbm{1}_{\Omega},\mathbbm{1}_{\Omega},A[\mathbbm{1}_{\Omega},\mathbbm{1}_{\Omega}]))
d𝒫((f,g,A[f,g]),(𝟙Ω,𝟙Ω,A[𝟙Ω,𝟙Ω]))|\displaystyle\hskip 11.38092pt-d_{\mathcal{LP}}(\mathcal{L}(f,g,A[f,g]),\mathcal{L}(\mathbbm{1}_{\Omega},\mathbbm{1}_{\Omega},A[\mathbbm{1}_{\Omega},\mathbbm{1}_{\Omega}]))\big{|}
K334δ12max{(2C)12,1}c>0\displaystyle\geq K-3^{\frac{3}{4}}\delta^{\frac{1}{2}}\max\{(2C)^{\frac{1}{2}},1\}\geq c>0

where K>0K>0 and δ0\delta\rightarrow 0 as ε0\varepsilon\rightarrow 0. But this is in contradiction with

d𝒫((f,g,A[f,g]),(𝟙Ω,𝟙Ω,B[𝟙Ω,𝟙Ω]))ε.d_{\mathcal{LP}}(\mathcal{L}(f,g,A[f,g]),\mathcal{L}(\mathbbm{1}_{\Omega},\mathbbm{1}_{\Omega},B[\mathbbm{1}_{\Omega},\mathbbm{1}_{\Omega}]))\leq\varepsilon.

Remark 9.14.

We could have deduced directly that the weak limit of

(𝟙Ωn,𝟙Ωn,(An)2[𝟙Ωn,𝟙Ωn])δ(1,1,18)\mathcal{L}(\mathbbm{1}_{\Omega_{n}},\mathbbm{1}_{\Omega_{n}},(A_{n})_{2}[\mathbbm{1}_{\Omega_{n}},\mathbbm{1}_{\Omega_{n}}])\rightarrow\delta_{(1,1,\frac{1}{8})}

as we know the limit constant 33-graphon W=1/8W=1/8 on [0,1]3[0,1]^{3} of (An)2.(A_{n})_{2}. Observe in fact that

(𝟙[0,1],𝟙[0,1],(W)2[𝟙[0,1],𝟙[0,1])=δ(1,1,18).\mathcal{L}(\mathbbm{1}_{[0,1]},\mathbbm{1}_{[0,1]},(W)_{2}[\mathbbm{1}_{[0,1]},\mathbbm{1}_{[0,1]})=\delta_{(1,1,\frac{1}{8})}.

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 kk-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 33-rd order nn-dimensional symmetric tensors T=(Ti,j)i,j[n]T=(T_{i,j})_{i,j\in[n]} and (T~)i,j[n](\widetilde{T})_{i,j\in[n]}, the 22-actions T2T_{2} and T~2\widetilde{T}_{2} are at distance zero in action convergence distance dMd_{M} if and only if there exists a bijective map

ψ:[n][n]\psi:[n]\rightarrow[n]

such that

Ti,j,k=T~ψ(i),ψ(j),ψ(k).T_{i,j,k}=\widetilde{T}_{\psi(i),\psi(j),\psi(k)}.
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

ϕ:([n]×[n],Sym)([n]×[n],Sym)\displaystyle\phi:([n]\times[n],Sym)\longrightarrow([n]\times[n],Sym)
(i,k)ϕ(i,k)=(ϕ1(i,k),ϕ2(i,k))\displaystyle(i,k)\mapsto\phi(i,k)=(\phi_{1}(i,k),\phi_{2}(i,k))

such that

(T2[f,g])ϕ=T~2[fϕ,gϕ])(T_{2}[f,g])^{\phi}=\widetilde{T}_{2}[f^{\phi},g^{\phi}])

for all symmetric matrices f,gf,g on [n]×[n][n]\times[n].

Because, in general, to have

(f,g,T2[f,g])=(fϕ,gϕ,(T~2[fϕ,gϕ])){\mathcal{L}}(f,g,T_{2}[f,g])={\mathcal{L}}(f^{\phi},g^{\phi},(\widetilde{T}_{2}[f^{\phi},g^{\phi}]))

we need

(T2[f,g])ϕ=T~2[fϕ,gϕ]).(T_{2}[f,g])^{\phi}=\widetilde{T}_{2}[f^{\phi},g^{\phi}]).

Therefore, we can compare the two terms

(T2[f,g])i,kϕ=12(j=1nTj,ϕ1(i,k),ϕ2(i,k)fj,ϕ1(i,k)gj,ϕ2(i,k)+j=1nTj,ϕ2(i,k),ϕ1(i,k)fj,ϕ2(i,k)gj,ϕ1(i,k))\displaystyle(T_{2}[f,g])_{i,k}^{\phi}=\frac{1}{2}(\sum^{n}_{j=1}T_{j,\phi_{1}(i,k),\phi_{2}(i,k)}f_{j,\phi_{1}(i,k)}g_{j,\phi_{2}(i,k)}+\sum^{n}_{j=1}T_{j,\phi_{2}(i,k),\phi_{1}(i,k)}f_{j,\phi_{2}(i,k)}g_{j,\phi_{1}(i,k)})

and

T2~[fϕ,gϕ]i,k=12(j=1nT~j,i,kfϕ1(j,i),ϕ2(j,i)gϕ1(j,k),ϕ2(j,k)+j=1nT~j,i,kfϕ1(j,k),ϕ2(j,k)gϕ1(j,i),ϕ2(j,i)).\widetilde{T_{2}}[f^{\phi},g^{\phi}]_{i,k}=\frac{1}{2}(\sum^{n}_{j=1}\widetilde{T}_{j,i,k}f_{\phi_{1}(j,i),\phi_{2}(j,i)}g_{\phi_{1}(j,k),\phi_{2}(j,k)}+\sum^{n}_{j=1}\widetilde{T}_{j,i,k}f_{\phi_{1}(j,k),\phi_{2}(j,k)}g_{\phi_{1}(j,i),\phi_{2}(j,i)}).

Now, we choose f=𝟙{ϕ1(i,k),a}f=\mathbbm{1}_{\{\phi_{1}(i,k),a\}} and g=𝟙{a,ϕ2(i,k)}g=\mathbbm{1}_{\{a,\phi_{2}(i,k)\}} where 𝟙{c,d}\mathbbm{1}_{\{c,d\}} is the indicator function of the set {(c,d),(d,c)}\{(c,d),(d,c)\}. Then we have

(T2[f,g])i,kϕ=12(j=1nTj,ϕ1(i,k),ϕ2(i,k)𝟙{ϕ1(i,k),a}j,ϕ1(i,k)𝟙{a,ϕ2(i,k)}j,ϕ2(i,k)\displaystyle(T_{2}[f,g])_{i,k}^{\phi}=\frac{1}{2}(\sum^{n}_{j=1}T_{j,\phi_{1}(i,k),\phi_{2}(i,k)}{\mathbbm{1}_{\{\phi_{1}(i,k),a\}}}_{j,\phi_{1}(i,k)}{\mathbbm{1}_{\{a,\phi_{2}(i,k)\}}}_{j,\phi_{2}(i,k)}
+j=1nTj,ϕ2(i,k),ϕ1(i,k)𝟙{ϕ1(i,k),a}j,ϕ2(i,k)𝟙{a,ϕ2(i,k)}j,ϕ1(i,k))=\displaystyle+\sum^{n}_{j=1}T_{j,\phi_{2}(i,k),\phi_{1}(i,k)}{\mathbbm{1}_{\{\phi_{1}(i,k),a\}}}_{j,\phi_{2}(i,k)}{\mathbbm{1}_{\{a,\phi_{2}(i,k)\}}}_{j,\phi_{1}(i,k)})=
12Ta,ϕ1(i,k),ϕ2(i,k)\displaystyle\frac{1}{2}T_{a,\phi_{1}(i,k),\phi_{2}(i,k)}

and

T2~[fϕ,gϕ]i,k=12(j=1nT~j,i,k𝟙{ϕ1(i,k),a}ϕ1(j,i),ϕ2(j,i)𝟙{a,ϕ2(i,k)}ϕ1(j,k),ϕ2(j,k)\displaystyle\widetilde{T_{2}}[f^{\phi},g^{\phi}]_{i,k}=\frac{1}{2}(\sum^{n}_{j=1}\widetilde{T}_{j,i,k}{\mathbbm{1}_{\{\phi_{1}(i,k),a\}}}_{\phi_{1}(j,i),\phi_{2}(j,i)}{\mathbbm{1}_{\{a,\phi_{2}(i,k)\}}}_{\phi_{1}(j,k),\phi_{2}(j,k)}
+j=1nT~j,i,k𝟙{ϕ1(i,k),a}ϕ1(j,k),ϕ2(j,k)𝟙{a,ϕ2(i,k)}ϕ1(j,i),ϕ2(j,i)).\displaystyle+\sum^{n}_{j=1}\widetilde{T}_{j,i,k}{\mathbbm{1}_{\{\phi_{1}(i,k),a\}}}_{\phi_{1}(j,k),\phi_{2}(j,k)}{\mathbbm{1}_{\{a,\phi_{2}(i,k)\}}}_{\phi_{1}(j,i),\phi_{2}(j,i)}).

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:

ϕ1(i,k)=ϕ1(d,i)\displaystyle\phi_{1}(i,k)=\phi_{1}(d,i) (12)
a=ϕ2(d,i)=ϕ2(d,k)\displaystyle a=\phi_{2}(d,i)=\phi_{2}(d,k)
ϕ2(i,k)=ϕ1(d,k)\displaystyle\phi_{2}(i,k)=\phi_{1}(d,k)
ϕ1(i,k)=ϕ2(d,i)\displaystyle\phi_{1}(i,k)=\phi_{2}(d,i) (13)
a=ϕ1(d,i)=ϕ2(d,k)\displaystyle a=\phi_{1}(d,i)=\phi_{2}(d,k)
ϕ2(i,k)=ϕ1(d,k)\displaystyle\phi_{2}(i,k)=\phi_{1}(d,k)
ϕ1(i,k)=ϕ1(d,i)\displaystyle\phi_{1}(i,k)=\phi_{1}(d,i) (14)
a=ϕ2(d,i)=ϕ1(d,k)\displaystyle a=\phi_{2}(d,i)=\phi_{1}(d,k)
ϕ2(i,k)=ϕ2(d,k)\displaystyle\phi_{2}(i,k)=\phi_{2}(d,k)
ϕ1(i,k)=ϕ2(d,i)\displaystyle\phi_{1}(i,k)=\phi_{2}(d,i) (15)
a=ϕ1(d,i)=ϕ1(d,k)\displaystyle a=\phi_{1}(d,i)=\phi_{1}(d,k)
ϕ2(i,k)=ϕ2(d,k)\displaystyle\phi_{2}(i,k)=\phi_{2}(d,k)

We observe that varying aa we accordingly vary dd as ϕ\phi is bijective. In fact, for all conditions (12), (13), (14) and (15) if there would be two distinct dd and d~\tilde{d} in [n][n] corresponding to the same aa then ϕ\phi would fail to be bijective. For this reason, we obtain from the conditions (12),(13),(14) and (15) that ϕ1\phi_{1} (respectively ϕ2\phi_{2}) depend only on the second variable. Moreover, we notice that a necessary condition to be bijective and measure-preserving (measurable) for ϕ\phi is

ϕ1(i,k)=ϕ2(k,i).\displaystyle\phi_{1}(i,k)=\phi_{2}(k,i). (16)

Therefore, we notice that conditions (13) and (14) would contradict condition (16). In conclusion, we can only have from (12) and (16) that

ϕ1(i,j)=ψ(j)\phi_{1}(i,j)=\psi(j)
ϕ2(i,j)=ψ(i)\phi_{2}(i,j)=\psi(i)

or from (15) and (16) that

ϕ1(i,j)=ψ(i)\phi_{1}(i,j)=\psi(i)
ϕ2(i,j)=ψ(j).\phi_{2}(i,j)=\psi(j).

where ψ\psi is a permutation of [n].[n]. Therefore, substituting and requiring that

(T2[f,g])ϕ=T2~[fϕ,gϕ](T_{2}[f,g])^{\phi}=\widetilde{T_{2}}[f^{\phi},g^{\phi}]

we obtain that

Tψ(d),ψ(i),ψ(k)=Ta,ψ(i),ψ(k)=Ta,ϕ1(i,k),ϕ2(i,k)=2(T2[f,g])i,kϕ=2T2~[fϕ,gϕ]i,k=T~d,i,k.T_{\psi(d),\psi(i),\psi(k)}=T_{a,\psi(i),\psi(k)}=T_{a,\phi_{1}(i,k),\phi_{2}(i,k)}=2(T_{2}[f,g])_{i,k}^{\phi}=2\widetilde{T_{2}}[f^{\phi},g^{\phi}]_{i,k}=\widetilde{T}_{d,i,k}.

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 r1r-1-actions of two rr-th order symmetric tensors T=(Ti1,,ir)i1,,ir[n]T=(T_{i_{1},\ldots,i_{r}})_{i_{1},\ldots,i_{r}\in[n]} and T~=(T~i1,,ir)i1,,ir[n]\widetilde{T}=(\widetilde{T}_{i_{1},\ldots,i_{r}})_{i_{1},\ldots,i_{r}\in[n]} are completely determined by the action convergence distance, i.e. their (r1)(r-1)-actions are at action convergence distance dMd_{M} zero if and only if

Tψ(i1),,ψ(ir)=T~i1,,ir.T_{\psi(i_{1}),\ldots,\psi(i_{r})}=\widetilde{T}_{i_{1},\ldots,i_{r}}.

In fact, similarly to the case r=3r=3, there must exist a bijective and measure-preserving transformation

ϕ=(ϕ1,,ϕr):([n]r1,Sym)([n]r1,Sym)\phi=(\phi_{1},\ldots,\phi_{r}):([n]^{r-1},Sym)\longrightarrow([n]^{r-1},Sym)

such that

((T)r1[f1,,fr1])ϕ=(T~)r1[f1ϕ,,fr1ϕ]((T)_{r-1}[f_{1},\ldots,f_{r-1}])^{\phi}=(\widetilde{T})_{r-1}[f^{\phi}_{1},\ldots,f^{\phi}_{r-1}]

for all f1,,fr1f_{1},\ldots,f_{r-1} symmetric (r1)(r-1)-th order tensors and where for a symmetric (r1)(r-1)-th order tensor ff we define

fϕ(i1,,ir1)=f(ϕ1(i1),,ϕr1(ir1)).f^{\phi}(i_{1},\ldots,i_{r-1})=f(\phi_{1}(i_{1}),\ldots,\phi_{r-1}(i_{r-1})).

Moreover, using the test functions fs=𝟙{a,ϕ1(i1,,ir1),,ϕ^s(i1,,ϕr1),,ϕr1(i1,,,ir1)}f_{s}=\mathbbm{1}_{\{a,\phi_{1}(i_{1},\ldots,i_{r-1}),\ldots,\hat{\phi}_{s}(i_{1},\ldots,\phi_{r-1}),\ldots,\phi_{r-1}(i_{1},\ldots,\ldots,i_{r-1})\}}, where 𝟙{a1,,ar}\mathbbm{1}_{\{a_{1},\ldots,a_{r}\}} represents the indicator function of the set

{(aσ(1),,aσ(r))[n]r1:σ is a permutation of [r1]},\{(a_{\sigma(1)},\ldots,a_{\sigma(r)})\in[n]^{r-1}:\sigma\text{ is a permutation of }[r-1]\},

the conditions on the ϕi\phi_{i} imposed by the fact that ϕ\phi is measure-preserving and bijective we obtain that for a permutation σ\sigma of [r1][r-1] we have

ϕ(i1,,ir1)=(ψ(iσ(1)),,ψ(iσ(r1)))\phi(i_{1},\ldots,i_{r-1})=(\psi(i_{\sigma(1)}),\ldots,\psi(i_{\sigma(r-1)}))

where ψ\psi is a permutation of [n].[n].

The previous theorem has the following direct important corollary:

Corollary 9.16.1.

For two hypergraphs H1H_{1} and H2H_{2} with maximal edge cardinality rr the (r1)(r-1)-actions of their adjacency tensors A(H1)A(H_{1}) and A(H2)A(H_{2}) (that are rr-th order tensors) are identified by the action convergence metric dMd_{M} if and only if the hypergraphs H1H_{1} and H2H_{2} are isomorphic.

We expect that the previous theorem and remark can be generalized to any ss-action (s[r1]s\in[r-1]) of an rr-th order tensor. The 11-action case is trivial and we showed in the previous theorem and remark the (r1)(r-1)-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 22-action for 33-uniform hypergraphs might not be the best choice for sparser hypergraphs and the 11-action might be sometimes more appropriate as the following example shows.

Example 10.1.

Consider the 33-uniform hypergraph T(n,sn)T(n,s_{n}) given by the triangles of the sparse Erdős–Rényi random graph G(n,sn)G(n,s_{n}) where sn0s_{n}\rightarrow 0 and snns_{n}n\rightarrow\infty. For every nn we consider a realization HnH_{n} of T(n,sn)T(n,s_{n}) and the related graph GnG_{n} on the same vertex set with the hyperedges of HnH_{n} as triangles. Let’s denote with En[n]×[n]E_{n}\subset[n]\times[n] the (symmetric) set of edges of GnG_{n} and recall that we denote with A(Hn)A(H_{n}) the adjacency tensor of HnH_{n}. In this case, for every fn,gnf_{n},g_{n} (sequences of) symmetric matrices, ((A(Hn)/sn)2[fn,gn])\mathcal{L}\left(\left({A(H_{n})}/{s_{n}}\right)_{2}[f_{n},g_{n}]\right) weakly converges to δ0,\delta_{0}, the Dirac function centered in 0.0. In fact, if we consider the sequence of multi-PP-operators

(A(Hn)sn)2:(L([n]×[n],Sym,n))2L1([n]×[n],Sym,n)\left(\frac{A(H_{n})}{s_{n}}\right)_{2}:(L^{\infty}([n]\times[n],Sym,\mathbb{P}_{n}))^{2}\longrightarrow L^{1}([n]\times[n],Sym,\mathbb{P}_{n})
(A(Hn)sn)2[f,g]i,k=12(j=1nAj,i,ksnfj,i,gj,k+j=1nAj,i,ksnfj,k,gj,i)\left(\frac{A(H_{n})}{s_{n}}\right)_{2}[f,g]_{i,k}=\frac{1}{2}(\sum^{n}_{j=1}\frac{A_{j,i,k}}{s_{n}}f_{j,i},g_{j,k}+\sum^{n}_{j=1}\frac{A_{j,i,k}}{s_{n}}f_{j,k},g_{j,i})

where n\mathbb{P}_{n} is the uniform measure on [n]×[n][n]\times[n], (A(Hn)/sn)2[f,g]i,k0\left(A(H_{n})/{s_{n}}\right)_{2}[f,g]_{i,k}\neq 0 if and only if (i,k)En(i,k)\in E_{n}. But as nn\rightarrow\infty, n(En)0\mathbb{P}_{n}(E_{n})\rightarrow 0. For this reason, it might be appropriate to consider the 11-action or change the probability measures n\mathbb{P}_{n} in such a way that n\mathbb{P}_{n} converges to some positive constant (for example choose n\mathbb{P}_{n} as the uniform probability measures on EnE_{n}).

Now, we present some possible choices to adapt action convergence to non-uniform hypergraphs.

In fact, considering the 22-action (Definition 4.2) associated with a (normalized) adjacency tensor of a hypergraph HH

A~:(L([n]×[n],Sym,n))2L1([n]×[n],Sym,n)\widetilde{A}:(L^{\infty}([n]\times[n],Sym,\mathbb{P}_{n}))^{2}\longrightarrow L^{1}([n]\times[n],Sym,\mathbb{P}_{n})
A~[f,g]i,k=12(j=1nAj,i,knfj,i,gj,k+j=1nAj,i,knfj,k,gj,i)\widetilde{A}[f,g]_{i,k}=\frac{1}{2}(\sum^{n}_{j=1}\frac{A_{j,i,k}}{n}f_{j,i},g_{j,k}+\sum^{n}_{j=1}\frac{A_{j,i,k}}{n}f_{j,k},g_{j,i})

we notice that considering the probability space [n]×[n][n]\times[n] with uniform probability n\mathbb{P}_{n} (and the symmetric σ\sigma-algebra) the diagonal, i.e. the set

Dn={(i,i):i[n]}[n]×[n]D_{n}=\{(i,i):i\in[n]\}\subset[n]\times[n]

has probability n(Dn)=nn2=1n\mathbb{P}_{n}(D_{n})=\frac{n}{n^{2}}=\frac{1}{n}. Therefore, in the limit nn\rightarrow\infty we have that the edges of cardinality 22 do not play any role in the profile measures of the multi-linear operator. However, we can choose other probability measures n\mathbb{P}_{n} different from the uniform distribution so that the information from the edges with lower cardinality is not lost. A natural choice for n\mathbb{P}_{n} is the discrete measure defined by n({(i,i)})=12n\mathbb{P}_{n}(\{(i,i)\})=\frac{1}{2n} and n({(i,j),(j,i)})=12n(n1)\mathbb{P}_{n}(\{(i,j),(j,i)\})=\frac{1}{2n(n-1)}. This obviously characterizes uniquely the discrete probability measure n\mathbb{P}_{n}. In this case, n(Dn)=12\mathbb{P}_{n}(D_{n})=\frac{1}{2} 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 k>3k>3 where Ω=[n]k\Omega=[n]^{k} with the symmetric σ\sigma-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 n\mathbb{P}_{n}. 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

deg(i1,,ik1)=|{eE s.t. i1,,ik1e\displaystyle deg(i_{1},\ldots,i_{k-1})=|\{e\in E\ \text{ s.t. }\ i_{1},\ldots,i_{k-1}\in e (17)
and |e|=|{i1,,ik1}|+1\displaystyle\text{ and }|e|=|\{i_{1},\ldots,i_{k-1}\}|+1

in the following way

A~i1,,ik=Ai1,,ikdeg(i1,,ik1).\widetilde{A}_{i_{1},\ldots,i_{k}}=\frac{A_{i_{1},\ldots,i_{k}}}{deg(i_{1},\ldots,i_{k-1})}.

It is easy to notice that

deg(i1,,ik1)|V|k+1|V|deg(i_{1},\ldots,i_{k-1})\leq|V|-k+1\leq|V|

In the particular case k=3k=3 we have

A~i,j,k=Ai,j,kdeg(i,k)\widetilde{A}_{i,j,k}=\frac{A_{i,j,k}}{deg(i,k)}

This is interesting for inhomogeneous hypergraphs and for hypergraphs with different edge cardinality. In fact, we can define on Ω=[n]×[n]\Omega=[n]\times[n] the probability measure

n({(i,j)})=deg(i,j)2i,j=1,ijndeg(i,j)\mathbb{P}_{n}(\{(i,j)\})=\frac{deg(i,j)}{2\sum^{n}_{i,j=1,\ i\neq j}deg(i,j)}

if iji\neq j and

n({(i,i)})=deg(i,i)2i=1ndeg(i,i).\mathbb{P}_{n}(\{(i,i)\})=\frac{deg(i,i)}{2\sum^{n}_{i=1}deg(i,i)}.

These operators are also symmetric with respect to the right probability measure.

Lemma 10.3.

The operator (A~)2(\widetilde{A})_{2} is symmetric with respect to the probability measure n\mathbb{P}_{n}.

Proof.

The lemma follows from the following equality

𝔼[(A~)2[f,g]h]=\displaystyle\mathbb{E}[(\widetilde{A})_{2}[f,g]h]=
12(i,k=1,ikn(j=1nAi,j,kdeg(i,k)fi,jgj,k+j=1nAi,j,kdeg(i,k)gi,jfj,k))hi,kdeg(i,k)2i,k=1,ikndeg(i,k)+\displaystyle\frac{1}{2}(\sum^{n}_{i,k=1,\ i\neq k}(\sum^{n}_{j=1}\frac{A_{i,j,k}}{deg(i,k)}f_{i,j}g_{j,k}+\sum^{n}_{j=1}\frac{A_{i,j,k}}{deg(i,k)}g_{i,j}f_{j,k}))h_{i,k}\frac{deg(i,k)}{2\sum^{n}_{i,k=1,\ i\neq k}deg(i,k)}+
12(in(j=1nAi,j,ideg(i,i)fi,jgj,i+j=1nAi,j,ideg(i,i)gi,jfj,i))hi,ideg(i,i)2i=1ndeg(i,i))=\displaystyle\frac{1}{2}(\sum^{n}_{i}(\sum^{n}_{j=1}\frac{A_{i,j,i}}{deg(i,i)}f_{i,j}g_{j,i}+\sum^{n}_{j=1}\frac{A_{i,j,i}}{deg(i,i)}g_{i,j}f_{j,i}))h_{i,i}\frac{deg(i,i)}{2\sum^{n}_{i=1}deg(i,i)})=
12(i,k=1,ikn(j=1nAi,j,kfi,jgj,k+j=1nAi,j,kgi,jfj,k))hi,k12i,k=1,ikndeg(i,k)+\displaystyle\frac{1}{2}(\sum^{n}_{i,k=1,\ i\neq k}(\sum^{n}_{j=1}A_{i,j,k}f_{i,j}g_{j,k}+\sum^{n}_{j=1}A_{i,j,k}g_{i,j}f_{j,k}))h_{i,k}\frac{1}{2\sum^{n}_{i,k=1,\ i\neq k}deg(i,k)}+
12(in(j=1nAi,j,ifi,jgj,i+j=1nAi,j,igi,jfj,i))hi,i12i=1ndeg(i,i))=\displaystyle\frac{1}{2}(\sum^{n}_{i}(\sum^{n}_{j=1}A_{i,j,i}f_{i,j}g_{j,i}+\sum^{n}_{j=1}A_{i,j,i}g_{i,j}f_{j,i}))h_{i,i}\frac{1}{2\sum^{n}_{i=1}deg(i,i)})=
𝔼[(A~)2[f,h]g]=𝔼[(A~)2[h,g]f].\displaystyle\mathbb{E}[(\widetilde{A})_{2}[f,h]g]=\mathbb{E}[(\widetilde{A})_{2}[h,g]f].

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 k>3.k>3.

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 δ,\delta_{\square,\mathbb{R}}) is equivalent to the action convergence of the sequence of the normalized adjacency matrices

A(Gn)|V(Gn)|.\frac{A(G_{n})}{|V(G_{n})|}.

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 rr-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 A[n]A\subset[n], define r(A)r(A) to be the collection of all nonempty subsets of AA, and r<(A)r_{<}(A) to be the collection of all nonempty proper subsets of AA. More generally, let r(A,m)r(A,m) denote the collection of all nonempty subsets of AA of size at most mm. So for instance, r<([k])=r([k],k1)=r([k]{k})r_{<}([k])=r([k],k-1)=r([k]\setminus\{k\}). We will also use the shorthand r[k]r[k] and r<[k]r_{<}[k] to mean r([k])r([k]) and r<([k])r_{<}([k]) respectively.

Any permutation σ\sigma of a set AA induces a permutation on r(A,m)r(A,m). For a set A={v1,vt}[k]A=\{v_{1},\ldots v_{t}\}\subset[k] of cardinality tt where v1<<vtv_{1}<\ldots<v_{t}, we indicate with xA=(xv1,,xvt,xv1v2,xv1vt)\mathrm{x}_{A}=(x_{v_{1}},\ldots,x_{v_{t}},x_{v_{1}v_{2}}\ldots,x_{v_{1}\ldots v_{t}}).

The limit object of a sequence of rr-uniform hypergraphs, i.e. an rr-hypergraphon, is a symmetric measurable function

W:[0,1]2r2[0,1].W:[0,1]^{2^{r}-2}\longrightarrow[0,1].
W(xr[r])=W(x1,,xr,x12,,x(r1)r,x12r1,,x2r)W(\mathrm{x}_{r[r]})=W(x_{1},\ldots,x_{r},x_{12},\ldots,x_{(r-1)r},\ldots x_{12\ldots r-1},\ldots,x_{2\ldots r})

where symmetric means that

W(x1,,xr,x12,,x(r1)r,x12(r1),,x2r)=\displaystyle W(x_{1},\ldots,x_{r},x_{12},\ldots,x_{(r-1)r},\ldots x_{12\ldots(r-1)},\ldots,x_{2\ldots r})=
W(xσ(1),,xσ(r),xσ(1)σ(2),,xσ(r1)σ(r),xσ(1)σ(2)σ(r1),,xσ(2)σ(r))\displaystyle W(x_{\sigma(1)},\ldots,x_{\sigma{(r)}},x_{\sigma{(1)}\sigma{(2)}},\ldots,x_{\sigma(r-1)\sigma(r)},\ldots x_{\sigma(1)\sigma(2)\ldots\sigma(r-1)},\ldots,x_{\sigma(2)\ldots\sigma(r)})

for every permutation σ\sigma of [r][r]. This might be surprising because, differently from the case of graphs (r=2r=2), for r>2r>2 the dimensionality of the rr-th order adjacency tensor associated to an rr-uniform hypergraph, rr, does not coincide with the dimensionality of the rr-hypergraphon, 2r22^{r}-2.

The need for the additional coordinates, representing all proper subsets of [r][r], 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 rr-uniform hypergraphs for r>2r>2 [44].

This is also intuitively related to the various multi-PP-operators associated with a tensor through its ss-actions. In fact for r=3r=3 the additional coordinates are again needed, for example, to differentiate the limits of the sequence of the Erdős–Rényi 33-uniform hypergraphs G(n,18,2)G(n,\frac{1}{8},2) (Example 4.8) and the sequence of the 33-uniform hypergraphs T(n,12)T(n,\frac{1}{2}) given by the triangles of the Erdős–Rényi graph (Example 4.9).

We notice that similarly to how we associated graphons to PP-operators we can associate hypergraphons to multi-PP-operators:

W^:L([0,1]2r12,Sym)××L([0,1]2r12,Sym)L1([0,1]2r12,Sym)\widehat{W}:L^{\infty}([0,1]^{2^{r-1}-2},Sym)\times\ldots\times L^{\infty}([0,1]^{2^{r-1}-2},Sym)\longrightarrow L^{1}([0,1]^{2^{r-1}-2},Sym)
W^[g1,,g(r1)]\displaystyle\widehat{W}[g_{1},\ldots,g_{(r-1)}] (xr([r]{r}))\displaystyle(\mathrm{x}_{r([r]\setminus\{r\})}) (18)
=1(r1)!σ[0,1]2r2r1+1W(xr[r])i=1r1gσ(i)(xr[r]{i})dxA(r)\displaystyle=\frac{1}{(r-1)!}\sum_{\sigma}\int_{[0,1]^{2^{r}-2^{r-1}+1}}W(\mathrm{x}_{r[r]})\prod^{r-1}_{i=1}g_{\sigma(i)}(\mathrm{x}_{r_{[r]\setminus\{i\}}})\mathrm{d}\mathrm{x}_{A(r)}

where σ\sigma here is a permutation of [r1],[r-1], A(r)A(r) is the set of all the proper subsets of [r][r] containing r,r, and SymSym is the symmetric σ\sigma-algebra (i.e. the σ\sigma-algebra generated by the subsets of [0,1]2r12[0,1]^{2^{r-1}-2} that are invariant under the action of all permutations of [r1][r-1]). In particular, for r=3,r=3, we have

W^[g(1),g(2)](x1,x2,x12)\displaystyle\widehat{W}[g^{(1)},g^{(2)}](x_{1},x_{2},x_{12})
=12[0,1]4W(x1,x2,x3,x12,x13,x23)g(1)(x1,x3,x13)g(2)(x2,x3,x23)dx3dx13dx23\displaystyle=\frac{1}{2}\int_{[0,1]^{4}}W(x_{1},x_{2},x_{3},x_{12},x_{13},x_{23})g^{(1)}(x_{1},x_{3},x_{13})g^{(2)}(x_{2},x_{3},x_{23})\mathrm{d}x_{3}\mathrm{d}x_{13}\mathrm{d}x_{23}
+12[0,1]4W(x1,x2,x3,x12,x13,x23)g(2)(x1,x3,x13)g(1)(x2,x3,x23)dx3dx13dx23\displaystyle+\frac{1}{2}\int_{[0,1]^{4}}W(x_{1},x_{2},x_{3},x_{12},x_{13},x_{23})g^{(2)}(x_{1},x_{3},x_{13})g^{(1)}(x_{2},x_{3},x_{23})\mathrm{d}x_{3}\mathrm{d}x_{13}\mathrm{d}x_{23}

We observe that there are promising similarities between the action convergence of hypergraphons and the action convergence of the (r1)(r-1)-action of the adjacency tensor.

Let’s consider for example the hypergraphon,

W(x1,x2,x3,x12,x13,x23)={1 if 0x12,x13,x23120 elseW(x_{1},x_{2},x_{3},x_{12},x_{13},x_{23})=\begin{cases}1\ \text{ if }0\leq x_{12},x_{13},x_{23}\leq\frac{1}{2}\\ 0\ \text{ else}\end{cases}

that is the limit of the sequence of hypergraphs T(n,12)T(n,\frac{1}{2}) given by the triangles of the Erdős–Rényi random graph (see [45] for example) and the action convergence limit of the 22-action (Bn)2(B_{n})_{2} of the sequence of tensors (Bn)n(B_{n})_{n} obtained normalizing the adjacency tensors of the same hypergraphs, i.e. Bn=A(T(n,12))nB_{n}=\frac{A(T(n,\frac{1}{2}))}{n} (recall Example 9.12), we have, for example, that

(𝟙Ωn,𝟙Ωn,(Bn)2[𝟙Ωn,𝟙Ωn])12δ(1,1,0)+12δ(1,1,14)=(𝟙Ω,𝟙Ω,W^[𝟙Ω,𝟙Ω])\mathcal{L}(\mathbbm{1}_{\Omega_{n}},\mathbbm{1}_{\Omega_{n}},(B_{n})_{2}[\mathbbm{1}_{\Omega_{n}},\mathbbm{1}_{\Omega_{n}}])\rightarrow\frac{1}{2}\delta_{(1,1,0)}+\frac{1}{2}\delta_{(1,1,\frac{1}{4})}=\mathcal{L}(\mathbbm{1}_{\Omega},\mathbbm{1}_{\Omega},\widehat{W}[\mathbbm{1}_{\Omega},\mathbbm{1}_{\Omega}])

also if we take the set SnS_{n} to be the (symmetric) set ofpairs that correspond to edges of G(n,12)G(n,\frac{1}{2}). Then also

(𝟙Sn,𝟙Sn,(Bn)2[𝟙Sn,𝟙Sn])(𝟙[0,1]2×[0,12],𝟙[0,1]2×[0,12],W^[𝟙[0,1]2×[0,12],𝟙[0,1]2×[0,12]])\mathcal{L}(\mathbbm{1}_{S_{n}},\mathbbm{1}_{S_{n}},(B_{n})_{2}[\mathbbm{1}_{S_{n}},\mathbbm{1}_{S_{n}}])\rightarrow\mathcal{L}(\mathbbm{1}_{[0,1]^{2}\times[0,\frac{1}{2}]},\mathbbm{1}_{[0,1]^{2}\times[0,\frac{1}{2}]},\widehat{W}[\mathbbm{1}_{[0,1]^{2}\times[0,\frac{1}{2}]},\mathbbm{1}_{[0,1]^{2}\times[0,\frac{1}{2}]}])

and, similarly,

(𝟙Snc,𝟙Snc,(Bn)2[𝟙Snc,𝟙Snc])(𝟙[0,1]2×[12,1],𝟙[0,1]2×[12,1],W^[𝟙[0,1]2×[12,1],𝟙[0,1]2×[12,1]])\mathcal{L}(\mathbbm{1}_{S^{c}_{n}},\mathbbm{1}_{S^{c}_{n}},(B_{n})_{2}[\mathbbm{1}_{S^{c}_{n}},\mathbbm{1}_{S^{c}_{n}}])\rightarrow\mathcal{L}(\mathbbm{1}_{[0,1]^{2}\times[\frac{1}{2},1]},\mathbbm{1}_{[0,1]^{2}\times[\frac{1}{2},1]},\widehat{W}[\mathbbm{1}_{[0,1]^{2}\times[\frac{1}{2},1]},\mathbbm{1}_{[0,1]^{2}\times[\frac{1}{2},1]}])

and

(𝟙Sn,𝟙Snc,(Bn)2[𝟙Sn,𝟙Snc])(𝟙[0,1]2×[0,12],𝟙[0,1]2×[12,1],W^[𝟙[0,1]2×[0,12],𝟙[0,1]2×[12,1]]).\mathcal{L}(\mathbbm{1}_{S_{n}},\mathbbm{1}_{S^{c}_{n}},(B_{n})_{2}[\mathbbm{1}_{S_{n}},\mathbbm{1}_{S^{c}_{n}}])\rightarrow\mathcal{L}(\mathbbm{1}_{[0,1]^{2}\times[0,\frac{1}{2}]},\mathbbm{1}_{[0,1]^{2}\times[\frac{1}{2},1]},\widehat{W}[\mathbbm{1}_{[0,1]^{2}\times[0,\frac{1}{2}]},\mathbbm{1}_{[0,1]^{2}\times[\frac{1}{2},1]}]).

Moreover, for any two 33-hypergraphons WW and UU we can consider the multi-action convergence metric dMd_{M} between the associated multi-PP-operators W^\widehat{W} and U^\widehat{U} defined in equation (18). In particular, in this case, for the multi-PP-operators W^\widehat{W}, equation (2) in the construction of the action convergence metric is

(g1(1),,gk(1),g1(2),,gk(2),W^[g1(1),g1(2)],,W^[gk(1),gk(2)]).\displaystyle\mathcal{L}(g^{(1)}_{1},\ldots,g^{(1)}_{k},g^{(2)}_{1},\ldots,g^{(2)}_{k},\widehat{W}[g_{1}^{(1)},g_{1}^{(2)}],\ldots,\widehat{W}[g_{k}^{(1)},g_{k}^{(2)}]).

where, for j[k],j\in[k], we consider gj(1),gj(2)L[1,1](Ω1×Ω1×Ω2).g^{(1)}_{j},g^{(2)}_{j}\in L_{[-1,1]}^{\infty}(\Omega_{1}\times\Omega_{1}\times\Omega_{2}).

From Lemma 8.2 and Lemma 8.3 we also obtain the following estimate.

Lemma 11.1.

For any two 33-hypergraphons WW and UU and the associated multi-PP-operators W^\widehat{W} and U^\widehat{U} defined in equation (18) we have the following inequality

dM(W^,U^)\displaystyle d_{M}(\widehat{W},\widehat{U}) 12(WU2)1/2\displaystyle\leq{\color[rgb]{0,0,0}12}(\|W-U\|_{\square_{2}})^{1/2}

where for a linear combination of 33-hypergraphons VV

V2=supg1,g2,g3\displaystyle\|V\|_{\square_{2}}=\sup_{g_{1},g_{2},g_{3}} |[0,1]6V(x1,x2,x3,x12,x13,x23)g1(x1,x2,x12)g2(x2,x3,x23)g3(x1,x3,x13)dx1\displaystyle\left|\int_{[0,1]^{6}}V(x_{1},x_{2},x_{3},x_{12},x_{13},x_{23})g_{1}(x_{1},x_{2},x_{12})g_{2}(x_{2},x_{3},x_{23})g_{3}(x_{1},x_{3},x_{13})\mathrm{d}x_{1}\right.
dx2dx3dx12dx13dx23|.\displaystyle\left.\mathrm{d}x_{2}\mathrm{d}x_{3}\mathrm{d}x_{12}\mathrm{d}x_{13}\mathrm{d}x_{23}\right|.

where the supremum is taken over measurable gi:[0,1]3[0,1]g_{i}:[0,1]^{3}\rightarrow[0,1] for every i[3]i\in[3] such that gi(x1,x2,x12)=gi(x2,x1,x12).g_{i}(x_{1},x_{2},x_{12})=g_{i}(x_{2},x_{1},x_{12}).

Remark 11.2.

More generally for two rr-hypergraphons WW and UU we have the following bound for the multi-action convergence distance dMd_{M} between the multi-PP-operators W^\widehat{W} and U^\widehat{U} defined in equation (18):

dM(W^,U^)32r1(WUr1)1/2d_{M}(\widehat{W},\widehat{U})\leq{\color[rgb]{0,0,0}3\cdot 2^{r-1}}(\|W-U\|_{\square_{r-1}})^{1/2}

where r1\|\cdot\|_{\square_{r-1}} is the (r1)(r-1)-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.

Hypergraphon convergence in the sense of Definition 6.6 of [45] (Partitionable convergence) implies action convergence of hypergraphons (interpreted as multi-PP-operators as in (18)). Moreover, the limits have to be compatible.

We anticipate a deeper connection between multi-action convergence, PP-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 Ω1=Ω2=[0,1]\Omega_{1}=\Omega_{2}=[0,1] for every i[6].i\in[6]. Let WW be a hypergraphon and W^\widehat{W} its multi-PP-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 YY from [0,1]3=Ω1×Ω1×Ω2[0,1]^{3}=\Omega_{1}\times\Omega_{1}\times\Omega_{2} to 7k\mathbb{R}^{7k}

Y(x1,x2,x12)\displaystyle Y(x_{1},x_{2},x_{12})
=(f1(1)(x1),f1(1)(x2),,fk(1)(x1),fk(1)(x2),g1(1)(x1,x2,x12),,gk(1)(x1,x2,x12)\displaystyle=(f^{(1)}_{1}(x_{1}),f^{(1)}_{1}(x_{2}),\ldots,f^{(1)}_{k}(x_{1}),f^{(1)}_{k}(x_{2}),g^{(1)}_{1}(x_{1},x_{2},x_{12}),\ldots,g^{(1)}_{k}(x_{1},x_{2},x_{12})
f1(2)(x1),f1(2)(x2),,fk(2)(x1),fk(2)(x2),g1(2)(x1,x2,x12),,gk(2)(x1,x2,x12)\displaystyle\hskip 11.38092ptf^{(2)}_{1}(x_{1}),f^{(2)}_{1}(x_{2}),\ldots,f^{(2)}_{k}(x_{1}),f^{(2)}_{k}(x_{2}),g^{(2)}_{1}(x_{1},x_{2},x_{12}),\ldots,g^{(2)}_{k}(x_{1},x_{2},x_{12})
W^[g1(1),g1(2)](x1,x2,x12),,W^[gk(1),gk(2)](x1,x2,x12)).\displaystyle\hskip 11.38092pt\widehat{W}[g_{1}^{(1)},g_{1}^{(2)}](x_{1},x_{2},x_{12}),\ldots,\widehat{W}[g_{k}^{(1)},g_{k}^{(2)}](x_{1},x_{2},x_{12})).

where, for j[k],j\in[k], we consider gj(1),gj(2)L[1,1](Ω1×Ω1×Ω2)g^{(1)}_{j},g^{(2)}_{j}\in L_{[-1,1]}^{\infty}(\Omega_{1}\times\Omega_{1}\times\Omega_{2}) as before and we additionally consider fj(1),fj(2)L[1,1](Ω1).f^{(1)}_{j},f^{(2)}_{j}\in L^{\infty}_{[-1,1]}(\Omega_{1}).

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 PP-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 PP-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 X,YX,Y be two jointly distributed k\mathbb{R}^{k}-valued random variables. Then

d𝒫((X),(Y))τ(XY)1/2k3/4,d_{\mathcal{LP}}(\mathcal{L}(X),\mathcal{L}(Y))\leq\tau(X-Y)^{1/2}k^{3/4},

where τ\tau 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 v1,v2,,vkv_{1},v_{2},\dots,v_{k} and w1,w2,,wkw_{1},w_{2},\dots,w_{k} be in L1(Ω)L^{1}(\Omega) for some probability space Ω\Omega. Let m:=maxi[k]viwi1m:=\max_{i\in[k]}\|v_{i}-w_{i}\|_{1}. Then

d𝒫((v1,v2,,vk),(w1,w2,,wk))m1/2k3/4.d_{\mathcal{LP}}(\mathcal{L}(v_{1},v_{2},\dots,v_{k}),\mathcal{L}(w_{1},w_{2},\dots,w_{k}))\leq m^{1/2}k^{3/4}.

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 q(1,)q\in(1,\infty). Let {(Xi,Yi)}i=1\{(X_{i},Y_{i})\}_{i=1}^{\infty} be a sequence of pairs of jointly distributed real-valued random variables such that Xi[1,1]X_{i}\in[-1,1] and 𝔼[|Yi|q]c<\mathbb{E}[|Y_{i}|^{q}]\leq c<\infty for some c+c\in\mathbb{R}^{+}. Assume that the distributions of (Xi,Yi)(X_{i},Y_{i}) weakly converge to some probability distribution (X,Y)(X,Y) as ii goes to infinity. Then 𝔼[|Y|q]c\mathbb{E}[|Y|^{q}]\leq c and

limi𝔼[XiYi]=𝔼[XY].\lim_{i\to\infty}\mathbb{E}[X_{i}Y_{i}]=\mathbb{E}[XY].

We give a last technical upper bound for the Lévy–Prokhorov distance of measures generated by a PP-operator through specific random variables. This is a minor modification of Lemma 13.6 in [6].

Lemma 11.6.

Let p[1,)p\in[1,\infty) and let Ar(Ω)A\in\mathcal{B}_{r}(\Omega) be a multi-PP-operator. Let viv_{i} and wiw_{i} be functions in L[1,1](Ω)L_{[-1,1]}^{\infty}(\Omega) for every i[k]i\in[k]. Then we have

d𝒫(𝒟A({vi}i=1k),𝒟A({wi}i=1k))m1/2((2d)p+2p+1d)1/(2p)(2k)3/4,d_{\mathcal{LP}}(\mathcal{D}_{A}(\{v_{i}\}_{i=1}^{k}),\mathcal{D}_{A}(\{w_{i}\}_{i=1}^{k}))\leq m^{1/2}((2d)^{p}+2^{p+1}d)^{1/(2p)}(2k)^{3/4},

where m=max{1,(r1)Ap,p1}m=\max\{1,(r-1)\|A\|_{p\ldots,p\to 1}\} and d=maxi[k]{d𝒫(𝒟(viwi),δ0)}d=\max_{i\in[k]}\{d_{\mathcal{LP}}(\mathcal{D}(v_{i}-w_{i}),\delta_{0})\}.

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 LpL^{p} 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 LpL^{p} 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 (p,q)(p,q)-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].