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

Geometric Characterization of the HH-property for Step-graphons

Mohamed-Ali Belabbas111M.-A. Belabbas is with the Department of Electrical and Computer Engineering and the Coordinated Science Laboratory, University of Illinois, Urbana-Champaign. Email: [email protected]. and Xudong Chen222X. Chen is with the Department of Electrical, Computer, and Energy Engineering, University of Colorado Boulder. Email: [email protected].
Abstract
22footnotetext: M.-A. Belabbas and X. Chen contributed equally to the manuscript in all categories.

In a recent paper [1], we have exhibited a set of conditions that are necessary for the HH-property to hold for the class of step-graphons. In this paper, we prove that these conditions are essentially sufficient.

1 Introduction and Main Result

In [1], we introduced the so-called HH-property for a graphon WW — roughly speaking, it is the property that a graph GG sampled from WW admits a Hamiltonian decomposition a.s.. The presence of a Hamiltonian decomposition in a graph underlies a number of important properties pertinent to structural system theory, such as structural controllability [2] and structural stability [3, 4]. See [1] for details. In that same paper, we have exhibited a set of conditions that were necessary for the HH-property to hold for the class of step-graphons. We show in this paper that these conditions are also essentially (in a sense made precise below) sufficient.

HH-property: We start by recalling the definitions of a graphon and its sampling procedure. A graphon is a symmetric, measurable function W:[0,1]2[0,1]W:[0,1]^{2}\to[0,1]. Step-graphons, along with their partitions, are defined below:

Definition 1 (Step-graphon and its partition).

A graphon WW is a step-graphon if there exists an increasing sequence 0=σ0<σ1<<σq=10=\sigma_{0}<\sigma_{1}<\cdots<\sigma_{q}=1 such that WW is constant over each rectangle [σi,σi+1)×[σj,σj+1)[\sigma_{i},\sigma_{i+1})\times[\sigma_{j},\sigma_{j+1}) for all 0i,jq10\leq i,j\leq q-1. We call σ=(σ0,σ1,,σq)\sigma=(\sigma_{0},\sigma_{1},\ldots,\sigma_{q}) a partition for WW.

Graphons can be used to sample undirected graphs. Other uses of graphons in system theory as limits of adjacency matrices can be found in [5, 6, 7]. In this paper, we denote by GnWG_{n}\sim W graphs GnG_{n} on nn nodes sampled from a graphon WW. The sampling procedure was introduced in [8, 9] and is reproduced below: Let Uni[0,1]\mathrm{Uni}[0,1] be the uniform distribution on [0,1][0,1]. Given a graphon WW, a graph Gn=(V,E)WG_{n}=(V,E)\sim W on nn nodes is obtained as follows:

  1. 1.

    Sample y1,,ynUni[0,1]y_{1},\ldots,y_{n}\sim\mathrm{Uni}[0,1] independently. We call yiy_{i} the coordinate of node viVv_{i}\in V.

  2. 2.

    For any two distinct nodes viv_{i} and vjv_{j}, place an edge (vi,vj)E(v_{i},v_{j})\in E with probability W(yi,yj)W(y_{i},y_{j}).

It should be clear that if 0p10\leq p\leq 1 is a constant and W(s,t)=pW(s,t)=p for all (s,t)[0,1]2(s,t)\in[0,1]^{2}, then GnWG_{n}\sim W is an Erdős-Rényi random graph with parameter pp. Consequently, graphons can be seen as a way to introduce inhomogeneity in the edge densities between different pairs of nodes.

Let WW be a graphon and GnWG_{n}\sim W. In the sequel, we use the notation Gn=(V,E)\vec{G}_{n}=(V,\vec{E}) to denote the directed version of GnG_{n}, defined by the edge set

E:={vivj,vjvi(vi,vj)E}.\vec{E}:=\{v_{i}v_{j},v_{j}v_{i}\mid(v_{i},v_{j})\in E\}. (1)

In words, we replace an undirected edge (vi,vj)(v_{i},v_{j}) with two directed edges vivjv_{i}v_{j} and vjviv_{j}v_{i}. The directed graph Gn\vec{G}_{n} is said to have a Hamiltonian decomposition if it contains a subgraph HH, with the same node set of G\vec{G}, such that HH is a node disjoint union of directed cycles. See Fig. 1 for illustration.

Refer to caption
(a)
Refer to caption
(b)
Figure 1: Left: An undirected graph GG on 44 nodes. Right: The directed graph G\vec{G} obtained from GG by replacing every undirected with two oppositely oriented edges. The cycle D1=v1v2v3v4v1D_{1}=v_{1}v_{2}v_{3}v_{4}v_{1} forms a Hamiltonian decomposition of G\vec{G}. The two node disjoint cycles D2=v1v2v1D_{2}=v_{1}v_{2}v_{1} and D3=v3v4v2D_{3}=v_{3}v_{4}v_{2} also form a Hamiltonian decomposition of G\vec{G}.

We now have the following definition:

Definition 2 (HH-property).

Let WW be a graphon and GnWG_{n}\sim W. Then, WW has the HH-property if

limn(Gn has a Hamiltonian decomposition)=1.\lim_{n\to\infty}\mathbb{P}(\vec{G}_{n}\mbox{ has a Hamiltonian decomposition})=1. (2)

We will see below that the HH-property is essentially a “zero-one” property in a sense that the probability on the left hand side of (2) converges to either 0 or 11.

Key objects: We present three key objects associated with a step-graphon, namely, its concentration vector, its skeleton graph, and its associated edge polytope, all of which were introduced in [1].

Definition 3 (Concentration vector).

Let WW be a step-graphon with partition σ=(σ0,,σq)\sigma=(\sigma_{0},\ldots,\sigma_{q}). The associated concentration vector x=(x1,,xq)x^{*}=(x^{*}_{1},\ldots,x^{*}_{q}) has entries defined as follows: xi:=σiσi1x^{*}_{i}:=\sigma_{i}-\sigma_{i-1}, for all i=1,,qi=1,\ldots,q.

It should be clear from the sampling procedure above that the concentration vector describes the proportion of sampled nodes in each interval [σi,σi+1)[\sigma_{i},\sigma_{i+1}).

Given a step-graphon, its support can be described by a graph, which we call skeleton graph:

Definition 4 (Skeleton graph).

To a step-graphon WW with a partition σ=(σ0,,σq)\sigma=(\sigma_{0},\ldots,\sigma_{q}), we assign the undirected graph S=(U,F)S=(U,F) on qq nodes, with U={u1,,uq}U=\{u_{1},\ldots,u_{q}\} and edge set FF defined as follows: there is an edge between uiu_{i} and uju_{j} if and only if WW is non-zero over [σi1,σi)×[σj1,σj)[\sigma_{i-1},\sigma_{i})\times[\sigma_{j-1},\sigma_{j}). We call SS the skeleton graph of WW for σ\sigma.

We illustrate the relationship between a step-graphon and its skeleton graph in Figure 2.

Refer to caption
(a)
Refer to caption
(b)
Figure 2: Left: A step-graphon WW with the partition σ=(0,0.25,0.5,0.75,1)\sigma=(0,0.25,0.5,0.75,1), with the value coded by the grayscale on the left. Right: The associated skeleton graph SS.

Without loss of generality and for ease of presentation, we will consider throughout this paper step-graphons WW whose skeleton graphs are connected. Though there is no unique skeleton graph associated to a step-graphon (since there are infinitely many different partitions for WW), we show in Proposition 1 that if one such skeleton graph is connected, then so are all the others. For SS not connected, it is not too hard to see that the corresponding step-graphon is block-diagonal. Our results apply naturally to every connected component of SS.

We decompose the edge set of SS as F=F0F1F=F_{0}\cup F_{1}, where elements of F0F_{0} are self-loops, and elements of F1F_{1} are edges between distinct nodes. We also introduce the subset F2F1F_{2}\subseteq F_{1} of edges that are not incident to two nodes with self-loops.

Let :={1,,|F|}\mathcal{I}:=\{1,\ldots,|F|\} be an index set for FF (so that the edges are now ordered). We decompose \mathcal{I} similarly: let 0\mathcal{I}_{0}, 1\mathcal{I}_{1}, and 2\mathcal{I}_{2} index F0F_{0}, F1F_{1}, and F2F_{2} respectively.

To introduce the edge-polytope of SS, we recall that the incidence matrix Z=[zij]Z=[z_{ij}] of SS is an |U|×|F||U|\times|F| matrix with its entries defined as follows:

zij:=12{2,if fjF0 is a loop on node ui,1,if node ui is incident to fjF1,0,otherwise.z_{ij}:=\frac{1}{2}\begin{cases}2,&\text{if }f_{j}\in F_{0}\text{ is a loop on node }u_{i},\\ 1,&\text{if node }u_{i}\text{ is incident to }f_{j}\in F_{1},\\ 0,&\text{otherwise}.\end{cases} (3)

Owing to the factor 12\frac{1}{2} in (3), all columns of ZZ are probability vectors, i.e., all entries are nonnegative and sum to one. The edge polytope of SS was introduced in [10] and the definition is reproduced below (with a slight difference in inclusion of the factor 12\frac{1}{2} of the generators zjz_{j}):

Definition 5 (Edge polytope).

Let S=(U,F)S=(U,F) be a skeleton graph and ZZ be the associated incidence matrix. Let zjz_{j}, for 1j|F|1\leq j\leq|F|, be the columns of ZZ. The edge polytope of SS, denoted by 𝒳(S)\mathcal{X}(S), is the convex hull of the vectors zjz_{j}:

𝒳(S):=conv{zjj=1,,|F|}.\mathcal{X}(S):=\operatorname{conv}\{z_{j}\mid j=1,\ldots,|F|\}. (4)

A point x𝒳(S)x\in\mathcal{X}(S) is said to be in the relative interior of 𝒳(S)\mathcal{X}(S), denoted by int𝒳(S)\operatorname{int}\mathcal{X}(S), if there exists an open neighborhood UU of xx in q\mathbb{R}^{q} (with q=|U|q=|U|) such that U𝒳(S)𝒳(S)U\cap\mathcal{X}(S)\subseteq\mathcal{X}(S). If xx is not an interior point, then it is called a boundary point and we write x𝒳(S)x\in\partial\mathcal{X}(S).

Main result: Let WW be a step-graphon. For a given partition σ\sigma for WW, let xx^{*} and SS be the associated concentration vector and the skeleton graph (which is assumed to be connected). We say that a cycle in SS is odd if it contains an odd number of distinct nodes (or edges); with this definition, self-loops are odd cycles. Given these, we state the following two conditions:

Condition A: The graph SS has an odd cycle.

Condition B: The vector xx^{*} belongs to int𝒳(S)\operatorname{int}\mathcal{X}(S).

The two conditions are stated in terms of a partition σ\sigma and its induced skeleton graph and edge-polytope. As mentioned earlier, there exist infinitely many partitions for a given step-graphon. However, the following proposition states that the two above conditions are invariant under changes of a partition.

Proposition 1.

Let WW be a step-graphon. For any two partitions σ\sigma and σ\sigma^{\prime} for WW, let xx^{*}, xx^{\prime*} be the corresponding concentration vectors and let SS, SS^{\prime} be the corresponding skeleton graphs. Then, the following hold:

  1. 1.

    SS is connected if and only if SS^{\prime} is;

  2. 2.

    SS has an odd cycle if and only if SS^{\prime} does;

  3. 3.

    x𝒳(S)x^{*}\in\mathcal{X}(S) (resp. xint𝒳(S)x^{*}\in\operatorname{int}\mathcal{X}(S)) if and only if x𝒳(S)x^{\prime*}\in\mathcal{X}(S^{\prime}) (resp. x𝒳(S)x^{\prime*}\in\mathcal{X}(S^{\prime})).

We refer the reader to Appendix A for a proof of the proposition.

We are now in a position to state the main result:

Main Theorem.

Let WW be a step-graphon. If it satisfies Conditions A and B for a given (and, hence, any) partition σ\sigma, then it has the HH-property.

Remark 1.

In our earlier work [1], we have shown that if a step-graphon WW has the HH-property, then it is necessary that Condition A and the following hold:

Condition B’: The vector xx^{*} belongs to 𝒳(S)\mathcal{X}(S).

Note that condition B’ is weaker than Condition B: Specifically, Condition B leaves out the set of step-graphons for which x𝒳(S)x^{*}\in\partial\mathcal{X}(S), which is a set of measure zero. For step-graphons satisfying Conditions A and B’, but not B, it is possible that

limn(GnW has a Hamiltonian decomposition)(0,1).\lim_{n\to\infty}\mathbb{P}(\vec{G}_{n}\sim W\mbox{ has a Hamiltonian decomposition})\in(0,1).

We have produced explicit examples of such step-graphon in [11, 1].  

Outline of proof: Given a step-graphon WW with skeleton graph SS, and GnWG_{n}\sim W, the sampling procedure induces a natural graph homomorphism π:GnS\pi:G_{n}\to S, whereby all nodes vjv_{j} of GnG_{n} whose coordinates yjy_{j} belong to [σi1,σi)[\sigma_{i-1},\sigma_{i}) are mapped to uiu_{i}. With a slight abuse of notation, we will use the same letter π\pi to denote the homomorphism π:GnS\pi:\vec{G}_{n}\to\vec{S}.

Let ni(Gn):=|π1(ui)|n_{i}(G_{n}):=|\pi^{-1}(u_{i})| be the number of nodes whose coordinates belong to [σi1,σi)[\sigma_{i-1},\sigma_{i}). We call the following vector the empirical concentration vector of GnG_{n}:

x(Gn):=1n(n1(Gn),,nq(Gn)).x(G_{n}):=\frac{1}{n}(n_{1}(G_{n}),\ldots,n_{q}(G_{n})). (5)

The proof of the Main Theorem contains three steps, outlined below, among which step 2 contains the bulk of the proof.

Step 1: The proof starts by showing how conditions AA and BB imply that the empirical concentration vector eventually belongs to the edge polytope. First, it should be clear that the edge polytope 𝒳(S)\mathcal{X}(S) is a subset of the standard simplex Δq1\Delta^{q-1} in q\mathbb{R}^{q}; thus, dim𝒳(S)(q1)\dim\mathcal{X}(S)\leq(q-1). Condition A, owing to [10], is both necessary and sufficient for the equality to hold. Next, note that nx(Gn)=(n1(Gn),,nq(Gn))nx(G_{n})=(n_{1}(G_{n}),\ldots,n_{q}(G_{n})) is a multinomial random variable with nn trials and qq outcomes with probabilities xix^{*}_{i}, for 1iq1\leq i\leq q. Then, Condition B guarantees, via Chebyshev’s inequality, that x(Gn)x(G_{n}) belongs to int𝒳(S)\operatorname{int}\mathcal{X}(S) a.s. — in this paper, a.s. stands for almost surely as nn\to\infty.

The next two steps are then dedicated to establishing the following fact:

x(Gn)int𝒳(S)Gn admits a Hamiltonian decomposition a.s..x(G_{n})\in\operatorname{int}\mathcal{X}(S)\\ \Rightarrow G_{n}\mbox{ admits a Hamiltonian decomposition {\em a.s.}.} (6)

Step 2: We start by working under the assumption that WW is a binary step-graphon, i.e., W(s,t){0,1}W(s,t)\in\{0,1\} for almost all (s,t)[0,1]2(s,t)\in[0,1]^{2}. In this case, we will see that a sampled graph GnWG_{n}\sim W is completely determined by its empirical concentration vector x(Gn)x(G_{n}). Consequently, our task (establishing (6)) is reduced to establishing the following:

x(Gn)int𝒳(S) and n is sufficiently largeGn admits a Hamiltonian decomposition surely.x(G_{n})\in\operatorname{int}\mathcal{X}(S)\mbox{ and $n$ is sufficiently large}\\ \Rightarrow G_{n}\mbox{ admits a Hamiltonian decomposition surely.} (7)

The proof of (7) is constructive. An important object that will arise therefrom is what we call the AA-matrix assigned to every Hamiltonian decomposition HH for Gn\vec{G}_{n}, written as a map ρ(H)=A\rho(H)=A.

Specifically, the matrix AA is a qq-by-qq matrix whose ijijth entry tallies the number of edges of HH that go from a node in π1(ui)\pi^{-1}(u_{i}) to a node in π1(uj)\pi^{-1}(u_{j}) (A precise definition is in Subsection 2.1 and see Figure 3 for an illustration). Any such matrix is then shown to satisfy a number of enviable properties, among which we have ρ(H)𝟏=x(Gn)\rho(H){\bf 1}=x(G_{n}).

In a nutshell, we have just created the following sequence of maps:

Hρ(H)ρ(H)𝟏=x(Gn),H\xmapsto{}\rho(H)\xmapsto{}\rho(H){\bf 1}=x(G_{n}),

with the domain being all Hamiltonian decompositions in Gn\vec{G}_{n}, for any GnG_{n} sampled from a given binary graphon.

Now, the effort in establishing (7) is to create appropriate right-inverses (at least locally) of the maps in the above sequence, i.e., we aim to create maps xA(x)x\mapsto A(x) and ρ~:A(x)H\tilde{\rho}:A(x)\mapsto H with the property that ρρ~\rho\cdot\tilde{\rho} is the identity map and A(x)𝟏=xA(x){\bf 1}=x. The map xA(x)x\mapsto A(x) is created in Proposition 2, Subsection 2.3, and the map ρ~\tilde{\rho} is created in Proposition 3, Subsection 2.4. From these two subsections, it will be clear that by introducing the AA-matrix as an intermediate object, we can decouple the analytic part of the proof, contained in the creation of the map xA(x)x\mapsto A(x), from the graph-theoretic part, contained in the creation of ρ~\tilde{\rho}. This will conclude the proof of (7).

Step 3: To close gap between binary step-graphons and general ones, we introduce here an equivalence relation on the class of step-graphons, namely, two step-graphons W1W_{1} and W2W_{2} are equivalent if their supports are the same. Or, said otherwise, W1W_{1} and W2W_{2} are equivalent if they share the same concentration vector and skeleton graph. Note, in particular, that each equivalence class [W][W] contains a unique representative which is a binary step-graphon, denoted by WsW^{s}. We then establish (6) by showing that WW has the HH-property if and only if WsW^{s} does. In essence, we show that the HH-property is decided completely by the concentration vector and the skeleton graph of a step-graphon WW.

The proof of this statement builds upon several classical results from random graph theory, and is presented in Subsection 2.5.

Notation: We gather here key notations and conventions.

Graph theory: Let G=(V,E)G=(V,E) be an undirected graph. Graphs in this paper do not have multiple edges, but may have self-loops. We denote edges by (vi,vj)E(v_{i},v_{j})\in E; if vi=vjv_{i}=v_{j}, then we call the edge a self-loop. For a given node viv_{i}, let N(vi):={vjV(vi,vj)E}N(v_{i}):=\{v_{j}\in V\mid(v_{i},v_{j})\in E\} be the neighbor set of viv_{i}. The degree of viv_{i}, denoted by deg(vi)\deg(v_{i}), is the cardinality of N(vi)N(v_{i}).

We will also deal with directed graphs (digraphs) in this paper. Whether a graph is directed or undirected will be clear from the context and/or notation. We denote by vivjv_{i}v_{j} the directed edge from viv_{i} to vjv_{j}; we call vjv_{j} an out-neighbor of viv_{i} and viv_{i} an in-neighbor of vjv_{j}. Similarly, we define N+(vi)N_{+}(v_{i}) and N(vi)N_{-}(v_{i}) the sets of in-neighbors and out-neighbors of viv_{i}, respectively.

Recall that for a given undirected graph G=(V,E)G=(V,E), possibly with self-loops, we let G=(V,E)\vec{G}=(V,\vec{E}) be the directed graph defined as in (1). Self-loops in G\vec{G} are the same as the ones in GG, i.e., they are not duplicated.

A closed walk in a graph (or digraph) is an ordered sequence of nodes v1v2vkv1v_{1}v_{2}\cdots v_{k}v_{1} in GG (resp. G\vec{G}) so that all consecutive nodes are ends of some edges (resp. directed edges). A cycle is a closed walk without repetition of nodes in the sequence except the starting- and the ending-nodes. For clarity of the presentation, we use letter CC to denote cycles in undirected graphs and the letter DD for cycles in directed graphs.

Miscellaneous: We use 𝟏\mathbf{1} to denote a column vector of all 11, whose dimension will be clear within the context. We write xyx\leq y for vectors x,yqx,y\in\mathbb{R}^{q} if the inequality holds entrywise. For a given vector xqx\in\mathbb{R}^{q}, we denote its 1\ell_{1} normalization by x¯\bar{x}, i.e., x¯:=xx1\bar{x}:=\frac{x}{\|x\|_{1}}, with the convention that 0¯=0\bar{0}=0. Further, given the vector xx, we denote by [x][x] the vector whose entry [x]i[x]_{i} is a closest integer to xix_{i} for 1iq1\leq i\leq q where for the case xi=k+12x_{i}=k+\frac{1}{2}, with kk an integer, we set [x]i:=k[x]_{i}:=k. We denote the standard simplex in q\mathbb{R}^{q} by Δq1:={xqx0 and x𝟏=1}\Delta^{q-1}:=\{x\in\mathbb{R}^{q}\mid x\geq 0\mbox{ and }x^{\top}{\bf 1}=1\}. Finally, given a q×qq\times q matrix AA, we denote by suppA\operatorname{supp}A its support, i.e., the set of indices corresponding to its non-zero entries.

2 Analysis and Proof of the Main Theorem

Throughout the proof, WW is a step-graphon, σ\sigma its associated partition, and S=(U,F)S=(U,F) its skeleton graph on qq nodes, which has an odd cycle. Let F0F_{0} and F1F_{1} be given as after Definition 4. We can naturally associate to them the subgraphs

S0:=(U,F0)andS1:=(U,F1).S_{0}:=(U,F_{0})\quad\mbox{and}\quad S_{1}:=(U,F_{1}). (8)

Note that SS has an odd cycle if and only if S0S_{0} is edgewise non-empty or S1S_{1} has an odd cycle. The lemma below states that we can consider, without loss of generality, only the latter case of S1S_{1} containing an odd cycle.

Lemma 1.

Let WW be a step-graphon. If WW admits a partition σ\sigma with skeleton graph SS containing an odd cycle, then WW admits a partition σ\sigma^{\prime} with skeleton graph SS^{\prime} so that the subgraph S1S^{\prime}_{1} has an odd cycle.

The proof of the lemma can be established by using the notion of “one-step refinement” introduced in Appendix A for the partition σ\sigma: If S1S_{1} already has an odd cycle, then there is nothing to prove. Otherwise, consider a one-step refinement on a node with self-loop in SS, which will yield a cycle of length 33 in SS^{\prime}.

2.1 On the edge polytope 𝒳(S)\mathcal{X}(S)

Rank of 𝒳(S)\mathcal{X}(S): Recall that 𝒳(S)\mathcal{X}(S) is the edge-polytope of SS. Similarly, we let 𝒳(Si)\mathcal{X}(S_{i}), for i=1,2i=1,2, be the edge polytope (see Definition 5) of SiS_{i}, i.e., 𝒳(Si)\mathcal{X}(S_{i}) is the convex hull of the zjz_{j}’s, with jij\in{\cal I}_{i}, where i{\cal I}_{i} indexes edges in FiF_{i}.

We call xx an extremal point of a polytope 𝒳\mathcal{X} if there is no line segment in 𝒳\mathcal{X} that contains xx in its interior. The maximal set of extremal points is called the set of extremal generators for 𝒳\mathcal{X}. The following result characterizes the extremal generators of 𝒳(S0)\mathcal{X}(S_{0}), 𝒳(S1)\mathcal{X}(S_{1}), and of 𝒳(S)\mathcal{X}(S):

Lemma 2.

The set of extremal generators of 𝒳(Si)\mathcal{X}(S_{i}), for i=1,2i=1,2, is {zjji}\{z_{j}\mid j\in\mathcal{I}_{i}\}. The set of extremal generators of 𝒳(S)\mathcal{X}(S) is {zii02}\{z_{i}\mid i\in\mathcal{I}_{0}\cup\mathcal{I}_{2}\}. .

Proof.

The statement for 𝒳(S0)\mathcal{X}(S_{0}) is obvious from the definition of the corresponding ziz_{i}. For 𝒳(S1)\mathcal{X}(S_{1}), it suffices to see that the vectors ziz_{i}, for i1i\in{\cal I}_{1}, have exactly two non-negative entries, and the supports of these vectors are pairwise distinct. Hence, if zi=j1cizjz_{i}=\sum_{j\in{\cal I}_{1}}c_{i}z_{j} with cj0c_{j}\geq 0, we necessarily have cj=0c_{j}=0 for jij\neq i and ci=1c_{i}=1. For 𝒳(S)\mathcal{X}(S), we refer the reader to [1, Proposition 1] for a proof.  

The rank of a polytope 𝒳\mathcal{X} is the dimension of its relative interior. It is known [10] that if SS has qq nodes, then

rank𝒳(S)={q1if S has an odd cycle,q2otherwise.\operatorname{rank}\mathcal{X}(S)=\begin{cases}q-1&\mbox{if $S$ has an odd cycle},\\ q-2&\mbox{otherwise.}\end{cases} (9)

Equivalently, we have the following result [12] on the rank of the incidence matrix ZSZ_{S} of SS:

rankZS={qif S has an odd cycle,q1otherwise.\operatorname{rank}Z_{S}=\begin{cases}q&\mbox{if $S$ has an odd cycle},\\ q-1&\mbox{otherwise.}\end{cases} (10)

The AA-matrix: Let GnWG_{n}\sim W and suppose that Gn\vec{G}_{n} has a Hamiltonian decomposition, denoted by HH.

Recall that π:GnS\pi:\vec{G}_{n}\to\vec{S} is the graph homomorphism introduced above (5). Let nij(H)n_{ij}(H) be the number of (directed) edges of HH from a node in π1(ui)\pi^{-1}(u_{i}) to a node in π1(uj)\pi^{-1}(u_{j}). It is not hard to see that (see [1, Lemma 1] for a proof) for all uiUu_{i}\in U,

ni(Gn)=ujN(ui)nij(H)=ujN(ui)nji(H).n_{i}(G_{n})=\sum_{u_{j}\in N(u_{i})}n_{ij}(H)=\sum_{u_{j}\in N(u_{i})}n_{ji}(H). (11)

We now assign to the skeleton graph SS a convex set that will be instrumental in establishing the main result:

Definition 6 (AA-matrices and their set).

Let S=(U,F)S=(U,F) be an undirected graph on qq nodes. We define 𝒜(S)\mathcal{A}(S) as the set of q×qq\times q nonnegative matrices A=[aij]A=[a_{ij}] satisfying the following two conditions:

  1. 1.

    If (ui,uj)F(u_{i},u_{j})\notin F, then aij=0a_{ij}=0;

  2. 2.

    A𝟏=A𝟏A{\bf 1}=A^{\top}{\bf 1}, and 𝟏A𝟏=1{\bf 1}^{\top}A{\bf 1}=1.

Because every defining condition for 𝒜(S){\cal A}(S) is affine, the set 𝒜(S)\mathcal{A}(S) is a convex set.

Now, to each Hamiltonian decomposition HH of Gn\vec{G}_{n}, we assign the following q×qq\times q matrix:

ρ(H):=1n[nij(H)]1i,jq.\rho(H):=\frac{1}{n}\left[n_{ij}(H)\right]_{1\leq i,j\leq q}. (12)

It follows from (11) that ρ(H)𝒜(S)\rho(H)\in\mathcal{A}(S) and ρ(H)𝟏=x(Gn)\rho(H){\bf 1}=x(G_{n}). Furthermore, we have established in [1, Proposition 4] the following connection between the set 𝒜(S)\mathcal{A}(S) and the edge polytope 𝒳(S)\mathcal{X}(S):

𝒳(S)={xqx=A𝟏 for some A𝒜(S)}.\mathcal{X}(S)=\{x\in\mathbb{R}^{q}\mid x=A{\bf 1}\mbox{ for some }A\in\mathcal{A}(S)\}. (13)

We refer the reader to Figure 3 for illustration.

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Figure 3: (a): A step-graphon WW. (b): Its skeleton graph SS. (c): A digraph Gn\vec{G}_{n}, with GnWG_{n}\sim W, for n=5n=5. The edges in black form the Hamiltonian decomposition HH. (d): The matrix ρ(H)\rho(H) defined in (12).

2.2 Local coordinate systems on 𝒳(S)\mathcal{X}(S) and 𝒳(S1)\mathcal{X}(S_{1})

This section establishes the groundwork for the construction of the map xA(x)x\mapsto A(x) described in the proof outline. To this end, we first show how to express any point in a neighborhood 𝒰{\cal U} of x𝒳(S)x^{*}\in\mathcal{X}(S) as a positive combination of the columns of the incidence matrix ZSZ_{S}. This amounts to solving the linear equations ZSϕ(x)=xZ_{S}\phi(x)=x, for x𝒰x\in{\cal U}, with ϕ(x)\phi(x) being continuous in xx and positive. We will solve a similar problem for y𝒳(S1)y^{*}\in\mathcal{X}(S_{1}) and with ZSZ_{S} replaced by ZS1Z_{S_{1}}, and we call the corresponding solution θ(y)\theta(y). These two maps will be put to use in the next subsection.

Construction of the map ϕ\phi: We start with the following lemma:

Lemma 3.

Suppose that S=(U,F)S=(U,F) has an odd cycle. Then, for any xint𝒳(S)x^{*}\in\operatorname{int}\mathcal{X}(S), there exist a closed neighborhood 𝒰\mathcal{U} of xx^{*} in the simplex and a continuous map ϕ:𝒰intΔ|F|1\phi:\mathcal{U}\to\operatorname{int}\Delta^{|F|-1} such that ZSϕ(x)=xZ_{S}\phi(x)=x for any x𝒰x\in\mathcal{U}.

Proof.

Because 𝒳(S)\mathcal{X}(S) is finitely generated by the columns of ZSZ_{S}, i.e. the ziz_{i}’s for ii\in{\cal I}, and because xint𝒳(S)x^{*}\in\operatorname{int}\mathcal{X}(S), there exists a positive probability vector c:=(c1,,c|F|)c:=(c_{1},\ldots,c_{|F|}) such that x=ZScx^{*}=Z_{S}c. Let ϵ:=12mini{ci}\epsilon:=\frac{1}{2}\min_{i\in{\cal I}}\{c_{i}\}. Because SS has at least one odd cycle, we know from (10) that ZSZ_{S} is full rank. Thus, we can pick qq columns, say z1,,zqz_{1},\ldots,z_{q} of ZSZ_{S}, that form a basis of q\mathbb{R}^{q}. Let B|F|B\subset\mathbb{R}^{|F|} be a closed ball centered at 0 with radius ϵ\epsilon, and let

B0:={yB𝟏y=0 and yi=0, for all i>q}.B_{0}:=\{y\in B\mid{\bf 1}^{\top}y=0\mbox{ and }y_{i}=0,\mbox{ for all }i>q\}.

The dimension of B0B_{0} is (q1)(q-1). We now define the map

ψ:B0q:yx+ZSy=ZS(y+c).\psi:B_{0}\to\mathbb{R}^{q}:y\mapsto x^{*}+Z_{S}y=Z_{S}(y+c).

It should be clear that ψ\psi is a linear bijection between B0B_{0} and its image ψ(B0)\psi(B_{0}). By the construction of BB and B0B_{0}, all the entries of (y+c)(y+c), for yB0y\in B_{0}, are positive and, moreover, 𝟏(y+c)=1{\bf 1}^{\top}(y+c)=1. It then follows that the image of ψ\psi is a closed neighborhood of xx^{*} inside 𝒳(S)\mathcal{X}(S). We now set ϕ:=ψ1\phi:=\psi^{-1}. It remains to show that ϕ(x)Δ|F|1\phi(x)\in\Delta^{|F|-1}. This holds because every column of ZSZ_{S} belongs to Δq1\Delta^{q-1} and so does xx. Thus, from ZSϕ(x)=xZ_{S}\phi(x)=x, we have that xx is a convex combination of the columns of ZSZ_{S}, which implies that ϕ(x)Δ|F|1\phi(x)\in\Delta^{|F|-1}.  

Let xint𝒳(S)x^{*}\in\operatorname{int}\mathcal{X}(S) and ϕ\phi be the map given in Lemma 3. For an edge fif_{i} of SS, we let ϕi\phi_{i} be the corresponding entry of ϕ\phi. We next define two functions τi:𝒰q\tau_{i}:\mathcal{U}\to\mathbb{R}^{q}, for i=1,2i=1,2, as follows:

τi(x):xjiϕj(x)zj.\tau_{i}(x):x\mapsto\sum_{j\in{\cal I}_{i}}\phi_{j}(x)z_{j}. (14)

If SS has no self-loops, then τ0\tau_{0} is set to be the zero map. We can thus decompose x𝒰x\in\mathcal{U} as

x=τ0(x)+τ1(x).x=\tau_{0}(x)+\tau_{1}(x).

We record the following simple observation for later use:

Lemma 4.

Let 𝒰{\cal U} be as in Lemma 3. For every x𝒰x\in{\cal U}, the set of indices of nonzero (positive) entries of τ0(x)\tau_{0}(x) is {iui has a self-loop}\{i\mid u_{i}\mbox{ has a self-loop}\} and, moreover, every entry of τ1(x)\tau_{1}(x) is positive.

Proof.

The statement for τ0(x)\tau_{0}(x) is trivial. The statement for τ1(x)\tau_{1}(x) follows from the fact that ϕ(x)\phi(x) has positive entries and no row of ZS1Z_{S_{1}} is identically 0.  

Construction of the map θ\theta: For any x𝒰x\in{\cal U}, let τ¯i(x)\bar{\tau}_{i}(x), for i=0,1i=0,1, be defined as follows:

τ¯i(x):={τi(x)/τi(x)1if τi(x)0,0otherwise.\bar{\tau}_{i}(x):=\begin{cases}\tau_{i}(x)/\|\tau_{i}(x)\|_{1}&\mbox{if }\tau_{i}(x)\neq 0,\\ 0&\mbox{otherwise}.\end{cases}

Since SS has odd cycle, recall that we can assume by Lemma 1 that S1S_{1} has an odd cycle. Thus, by (9), the rank of 𝒳(S1)\mathcal{X}(S_{1}) is (q1)(q-1). In particular, it implies that the relative interior of 𝒳(S1)\mathcal{X}(S_{1}) is open in Δq1\Delta^{q-1}. Further, by Lemma 4, if x𝒰x\in{\cal U}, then τ¯1(x)int𝒳(S1)\bar{\tau}_{1}(x)\in\operatorname{int}\mathcal{X}(S_{1}).

The map θ\theta we introduce below is akin to the map ϕ\phi introduced in Lemma 3, but defined on a closed neighborhood of

x¯1:=τ¯1(x)int𝒳(S1).\bar{x}_{1}^{*}:=\bar{\tau}_{1}(x^{*})\in\operatorname{int}\mathcal{X}(S_{1}).
Lemma 5.

Suppose that SS (and, hence, S1S_{1}) has an odd cycle; then, for the given x¯1int𝒳(S1)\bar{x}^{*}_{1}\in\operatorname{int}\mathcal{X}(S_{1}), there exist a closed neighborhood 𝒱{\cal V} of x¯1\bar{x}_{1}^{*} in Δq1\Delta^{q-1} and a continuous map θ:𝒱intΔ|F1|1\theta:{\cal V}\to\operatorname{int}\Delta^{|F_{1}|-1} such that ZS1θ(y)=yZ_{S_{1}}\theta(y)=y for any y𝒱y\in{\cal V}.

The proof is entirely similar to the one of Lemma 3, and is thus omitted.

Because ϕ\phi and θ\theta are both positive, continuous maps over closed, bounded domains, there exists an α(0,1)\alpha\in(0,1) so that

ϕ(x)α𝟏 for all x𝒰,\displaystyle\phi(x)\geq\alpha{\bf 1}\mbox{ for all }x\in{\cal U}, (15)
θ(x)α𝟏 for all x𝒱.\displaystyle\theta(x)\geq\alpha{\bf 1}\mbox{ for all }x\in{\cal V}.

On the image of τ¯1\bar{\tau}_{1}: For a given xint𝒳(S)x^{*}\in\operatorname{int}\mathcal{X}(S), the domains of ϕ\phi and θ\theta are closed neighborhoods 𝒰{\cal U} and 𝒱{\cal V} of xx^{*} and x¯1\bar{x}_{1}^{*}, respectively. Later in the analysis, we will pick an arbitrary x𝒰x\in{\cal U} and apply θ\theta to τ¯1(x)\bar{\tau}_{1}(x). For this, we need that τ¯1(x)\bar{\tau}_{1}(x) belongs to 𝒱{\cal V}. To this end, we will shrink 𝒰{\cal U} so that τ¯1(𝒰)𝒱\bar{\tau}_{1}({\cal U})\subseteq{\cal V} and thus the composition θτ¯1\theta\bar{\tau}_{1} is well defined. In fact, we have the stronger statement:

Lemma 6.

Let α>0\alpha>0 be given as in (15). There exist a closed neighborhood 𝒰𝒰{\cal U}^{\prime}\subseteq{\cal U} of xx^{*} and a positive ϵ<14α\epsilon<\frac{1}{4}\alpha, such that

τ1(x)+η¯=τ1(x)+ητ1(x)+η1𝒱,\overline{\tau_{1}(x)+\eta}=\frac{\tau_{1}(x)+\eta}{\|\tau_{1}(x)+\eta\|_{1}}\in{\cal V},

for any x𝒰x\in{\cal U}^{\prime} and for any ηq\eta\in\mathbb{R}^{q} with ηϵ\|\eta\|_{\infty}\leq\epsilon.

Proof.

Let 𝒱{\cal V}^{\prime} be a closed ball centered at x¯1\bar{x}_{1}^{*} and contained in the interior of 𝒱{\cal V}. Then, it is known [13, Theorem 4.6] that there exists an ϵ>0\epsilon^{\prime}>0 such that the ϵ\epsilon^{\prime}-neighborhood of 𝒱{\cal V}^{\prime}, with respect to the infinity norm, is contained in the interior of 𝒱{\cal V}. Let 𝒰:=τ¯11(𝒱){\cal U}^{\prime}:=\bar{\tau}_{1}^{-1}({\cal V}^{\prime}) and ϵ\epsilon be sufficiently small so that

(8+4α)ϵqα2<min{ϵ,14α}.\frac{(8+4\alpha)\epsilon}{q\alpha^{2}}<\min\left\{\epsilon^{\prime},\frac{1}{4}\alpha\right\}. (16)

We claim that the above-defined 𝒰{\cal U}^{\prime} and ϵ\epsilon are as desired.

Since τ¯1\bar{\tau}_{1} is continuous and since 𝒱{\cal V}^{\prime} is a closed ball centered at x¯1\bar{x}^{*}_{1}, 𝒰{\cal U}^{\prime} is a closed neighborhood of xx^{*}. Now, pick an arbitrary x𝒰x\in{\cal U}^{\prime}. For ease of notation, we set x1:=τ1(x)x_{1}:=\tau_{1}(x) and x¯1=τ¯1(x)\bar{x}_{1}=\bar{\tau}_{1}(x) for the remainder of this proof. Then,

x1+η¯x¯1\displaystyle\left\|\overline{x_{1}+\eta}-\bar{x}_{1}\right\|_{\infty} =x1+ηx1+η1x1x11\displaystyle=\left\|\frac{x_{1}+\eta}{\|x_{1}+\eta\|_{1}}-\frac{x_{1}}{\|x_{1}\|_{1}}\right\|_{\infty}
=(x11x1+η1)x1+x11ηx1+η1x11\displaystyle=\left\|\frac{(\|x_{1}\|_{1}-\|x_{1}+\eta\|_{1})x_{1}+\|x_{1}\|_{1}\eta}{\|x_{1}+\eta\|_{1}\|x_{1}\|_{1}}\right\|_{\infty}
|x11x1+η1|x1+η1x11x1+ηx1+η1\displaystyle\leq\frac{|\|x_{1}\|_{1}-\|x_{1}+\eta\|_{1}|}{\|x_{1}+\eta\|_{1}\|x_{1}\|_{1}}\|x_{1}\|_{\infty}+\frac{\|\eta\|_{\infty}}{\|x_{1}+\eta\|_{1}}
|x11x1+η1|x1+η1x11+ηx1+η1\displaystyle\leq\frac{|\|x_{1}\|_{1}-\|x_{1}+\eta\|_{1}|}{\|x_{1}+\eta\|_{1}\|x_{1}\|_{1}}+\frac{\|\eta\|_{\infty}}{\|x_{1}+\eta\|_{1}} (17)

where we used the fact that x11\|x_{1}\|_{\infty}\leq 1 to obtain the last inequality. To further evaluate (2.2), we first note that

|x1x1+η1|η1qηqϵ.|\|x\|_{1}-\|x_{1}+\eta\|_{1}|\leq\|\eta\|_{1}\leq q\|\eta\|_{\infty}\leq q\epsilon.

Next, by (3), (14) and (15), every entry of x1x_{1} is greater than 12α\frac{1}{2}\alpha, so x1112qα\|x_{1}\|_{1}\geq\frac{1}{2}q\alpha. Moreover, since ϵ<14α\epsilon<\frac{1}{4}\alpha,

x1+η1x11η112qαqϵ14qα.\|x_{1}+\eta\|_{1}\geq\|x_{1}\|_{1}-\|\eta\|_{1}\geq\frac{1}{2}q\alpha-q\epsilon\geq\frac{1}{4}q\alpha.

Finally, using (16), we can proceed from (2.2) and obtain that

x1+η¯x¯18ϵqα2+4ϵqα=(8+4α)ϵqα2<ϵ,\left\|\overline{x_{1}+\eta}-\bar{x}_{1}\right\|_{\infty}\leq\frac{8\epsilon}{q\alpha^{2}}+\frac{4\epsilon}{q\alpha}=\frac{(8+4\alpha)\epsilon}{q\alpha^{2}}<\epsilon^{\prime},

which implies that x1+η¯\overline{x_{1}+\eta} belongs to the ϵ\epsilon^{\prime}-neighborhood of 𝒱{\cal V}^{\prime} and, hence, to 𝒱{\cal V}.  

Remark 2.

From now on, to simplify the notation, we denote by 𝒰{\cal U} the set 𝒰{\cal U}^{\prime} of Lemma 6.  

2.3 Construction of the map xA(x)x\mapsto A(x)

To construct the map, we first specify its domain, which will be a subset of 𝒰{\cal U}. If xqx\in\mathbb{R}^{q} is an empirical concentration vector of some GnWG_{n}\sim W for a step-graphon WW, then nxnx necessarily has integer entries. Define a subset of 𝒰{\cal U} as follows:

𝒰:={x𝒰nx+q for some n+},{\cal U}^{*}:=\left\{x\in{\cal U}\mid nx\in\mathbb{Z}^{q}_{+}\mbox{ for some }n\in\mathbb{Z}_{+}\right\}, (18)

where +\mathbb{Z}_{+} is the set of positive integers.

Since the analysis for the HH-property will be carried out in the asymptotic regime nn\to\infty, the relevant empirical concentration vectors are those of GnG_{n} for nn large. To this end, let α(0,1)\alpha\in(0,1) be such that (15) is satisfied and ϵ\epsilon be as in Lemma 6. We have the following definition:

Definition 7.

Given an x𝒰x\in{\cal U}^{*}, n+n\in\mathbb{Z}_{+} is paired with xx if n>8αϵn>\frac{8}{\alpha\epsilon} and nxnx is integer valued.

With the above, we now state the main result of this subsection:

Proposition 2.

Let S=(U,F)S=(U,F) be a connected undirected graph, with at least one odd cycle, and S=(U,F)\vec{S}=(U,\vec{F}) be its directed version. Then, there exist a map A:𝒰𝒜(S)A:{\cal U}^{*}\mapsto\mathcal{A}(S) and a positive number a¯\underline{a} such that for any x𝒰x\in{\cal U}^{*} and for any nn paired with xx, the following hold:

  1. 1.

    A(x)𝟏=xA(x){\bf 1}=x;

  2. 2.

    nA(x)nA(x) is integer-valued and ndiagA(x)n\operatorname{diag}A(x) has even entries;

  3. 3.

    ndiagA(x)τ0(x)1n\|\operatorname{diag}A(x)-\tau_{0}(x)\|_{\infty}\leq 1, where τ0\tau_{0} is defined in (14);

  4. 4.

    n|aij(x)aji(x)|1n|a_{ij}(x)-a_{ji}(x)|\leq 1 for all 1i,jq1\leq i,j\leq q;

  5. 5.

    For any uiujFu_{i}u_{j}\in\vec{F}, aij(x)>a¯a_{ij}(x)>\underline{a}.

Note that item 5 and the fact that A(x)𝒜(S)A(x)\in\mathcal{A}(S) imply suppA(x)=S\operatorname{supp}A(x)=\vec{S}, i.e., aij0a_{ij}\neq 0 if and only if uiujFu_{i}u_{j}\in\vec{F}.

The proof of Proposition 2 is constructive. It will rely on a few technical facts we establish here. Let x𝒰x\in{\cal U}^{*} and nn be paired with xx. Define

τ0(x):=2n[n2τ0(x)]andτ1(x):=xτ0(x),\tau^{\prime}_{0}(x):=\frac{2}{n}\left[\frac{n}{2}\tau_{0}(x)\right]\quad\mbox{and}\quad\tau^{\prime}_{1}(x):=x-\tau^{\prime}_{0}(x), (19)

where we recall that the operation []\left[\,\cdot\,\right] is the integer-rounding operation, introduced in the notation of Section 1. The vector nτ0(x)n\tau_{0}^{\prime}(x) is then the vector with even entries closest to the entries of nτ0(x)n\tau_{0}(x). Next, we define

n0:=nτ0(x)1 and n1:=nτ1(x)1n^{\prime}_{0}:=n\|\tau^{\prime}_{0}(x)\|_{1}\mbox{ and }n^{\prime}_{1}:=n\|\tau^{\prime}_{1}(x)\|_{1} (20)

Recall that τ¯0\bar{\tau}^{\prime}_{0} and τ¯1\bar{\tau}^{\prime}_{1} are the 1\ell_{1} normalization of τ0\tau^{\prime}_{0} and τ1\tau^{\prime}_{1}, respectively. It should be clear from the construction that n0+n1=nn^{\prime}_{0}+n_{1}^{\prime}=n and

τ¯i=nniτi, for i=1,2.\bar{\tau}^{\prime}_{i}=\frac{n}{n^{\prime}_{i}}\tau^{\prime}_{i},\mbox{ for }i=1,2.

For a given x𝒰x\in{\cal U}^{*}, there obviously exist infinitely many positive integers nn that are paired with xx. However, the ratios n0/nn^{\prime}_{0}/n and n1/nn^{\prime}_{1}/n are independent of nn and determined completely by xx.

We also need the following lemma:

Lemma 7.

The following items hold:

  1. 1.

    For i=1,2i=1,2, suppτi(x)=suppτi(x)\operatorname{supp}\tau^{\prime}_{i}(x)=\operatorname{supp}\tau_{i}(x) and, moreover, the nonzero entries of τi(x)\tau^{\prime}_{i}(x) are uniformly bounded below by 12(1ϵ/4)α\frac{1}{2}(1-\epsilon/4)\alpha.

  2. 2.

    The ratio n0/nn^{\prime}_{0}/n is bounded below by (1ϵ/8)α|F0|(1-\epsilon/8)\alpha|F_{0}|. The ratio n1/nn_{1}^{\prime}/n is bounded below by 3qα/83q\alpha/8.

  3. 3.

    Let 𝒱{\cal V} be defined as in Lemma 5. Then, τ¯1(x)𝒱\bar{\tau}_{1}(x)\in{\cal V}.

We provide a proof of the lemma in Appendix B.

With the lemma above, we now establish Proposition 2:

Proof of Proposition 2.

We start by defining two matrix-valued functions A0(x)A_{0}(x) and A1(x)A_{1}(x) so that for any x𝒰x\in{\cal U}^{*}, Ai(x)𝒜(S)A_{i}(x)\in\mathcal{A}(S) and Ai(x)𝟏=τ¯i(x){A_{i}}(x){\bf 1}=\bar{\tau}^{\prime}_{i}(x). We will then let A(x)A(x) be the convex combination of these two matrices given by

A(x)=n0nA0(x)+n1nA1(x).A(x)=\frac{n^{\prime}_{0}}{n}A_{0}(x)+\frac{n_{1}^{\prime}}{n}A_{1}(x). (21)

Since 𝒜(S)\mathcal{A}(S) is convex, it will then follow that A(x)𝒜(S)A(x)\in\mathcal{A}(S).

Construction of A0A_{0}: The matrix A0(x)A_{0}(x) is simply given by

A0(x):=diagτ¯0(x).A_{0}(x):=\operatorname{diag}\bar{\tau}^{\prime}_{0}(x). (22)

By Lemma 4, suppτ0(x)\operatorname{supp}\tau_{0}(x) is constant over 𝒰{\cal U}. By the first item of Lemma 7, suppτ0(x)=suppτ0(x)\operatorname{supp}\tau_{0}(x)=\operatorname{supp}\tau^{\prime}_{0}(x) for all x𝒰x\in{\cal U}^{*}. It then follows that suppA0(x)\operatorname{supp}A_{0}(x) is also constant over 𝒰{\cal U}^{*}. By the same item, the nonzero entries of A0(x)A_{0}(x) are uniformly bounded below by a positive constant.

Construction of A1A_{1}: The construction is more involved than the one of A0A_{0}, and requires to first define the intermediate matrix A1A^{\prime}_{1}. To this end, recall that ZS1Z_{S_{1}} is the edge-incidence matrix of S1=(U,F1)S_{1}=(U,F_{1}), obtained by removing the self-loops of SS, and that θ\theta is the map given in Lemma 5, i.e., ZS1θ(y)=yZ_{S_{1}}\theta(y)=y for all y𝒱y\in{\cal V}. Given an edge f=(ui,uj)S1f=(u_{i},u_{j})\in S_{1}, we denote by θf(y)\theta_{f}(y) the corresponding entry of θ(y)\theta(y). By item 3 of Lemma 7, τ¯1(x)\bar{\tau}^{\prime}_{1}(x) belongs to 𝒱{\cal V}, which is the domain of θ\theta. Now, we define the symmetric matrix A1(x)=[a1,ij(x)]q×qA^{\prime}_{1}(x)=[a^{\prime}_{1,ij}(x)]\in\mathbb{R}^{q\times q} as follows:

a1,ij(x):={12θf(τ¯1(x))if f=(ui,uj)F1,0otherwise.a^{\prime}_{1,ij}(x):=\begin{cases}\frac{1}{2}\theta_{f}(\bar{\tau}^{\prime}_{1}(x))&\mbox{if }f=(u_{i},u_{j})\in F_{1},\\ 0&\mbox{otherwise}.\end{cases} (23)

In particular, the diagonal of A1A^{\prime}_{1} is 0, and so will be the diagonal of A1A_{1} as shown below. From the definition of the incidence matrix ZS1Z_{S_{1}} and (23), we have that A1(x)𝟏=ZS1θ(τ¯1(x))A^{\prime}_{1}(x){\bf 1}=Z_{S_{1}}\theta(\bar{\tau}^{\prime}_{1}(x)). By Lemma 5, ZS1θ(τ¯1(x))=τ¯1(x)Z_{S_{1}}\theta(\bar{\tau}^{\prime}_{1}(x))=\bar{\tau}^{\prime}_{1}(x). It then follows that

A1(x)𝟏=τ¯1(x),A^{\prime}_{1}(x){\bf 1}=\bar{\tau}^{\prime}_{1}(x), (24)

and, hence, 𝟏A1(x)𝟏=𝟏τ¯1(x)=1{\bf 1}^{\top}A^{\prime}_{1}(x){\bf 1}={\bf 1}^{\top}\bar{\tau}^{\prime}_{1}(x)=1. Furthermore, since A1(x)A^{\prime}_{1}(x) is symmetric, A1(x)𝟏=A1(x)𝟏{A^{\prime}_{1}(x)}{\bf 1}=A^{\prime}_{1}(x)^{\top}{\bf 1}. We thus have that A1(x)𝒜(S)A^{\prime}_{1}(x)\in\mathcal{A}(S). Since θf\theta_{f} is positive for every fF1f\in F_{1}, suppA1(x)\operatorname{supp}A^{\prime}_{1}(x) is constant over 𝒰{\cal U}. Moreover, by (15), the nonzero entries of A1(x)A^{\prime}_{1}(x) are uniformly bounded below by 12α\frac{1}{2}\alpha.

Next, we use A1A^{\prime}_{1} to construct A1A_{1}. There are two cases; one is straightforward and the other is more involved:

Case 1: n1A1(x)n_{1}^{\prime}A^{\prime}_{1}(x) is integer-valued: Set A1(x):=A1(x)A_{1}(x):=A^{\prime}_{1}(x).

Case 2: n1A1(x)n_{1}^{\prime}A^{\prime}_{1}(x) is not integer-valued: In this case, we appeal to the result [14, Theorem 2]: There, we have shown that there exists a matrix A1(x)=[a1,ij(x)]A_{1}(x)=[a_{1,ij}(x)] in 𝒜(S)\mathcal{A}(S), with

A1(x)𝟏=A1(x)𝟏=τ¯1(x)A_{1}(x){\bf 1}=A^{\prime}_{1}(x){\bf 1}=\bar{\tau}^{\prime}_{1}(x) (25)

such that A1(x)A_{1}(x) has the same support as A1(x)A^{\prime}_{1}(x) and

n1a1,ij(x)=n1a1,ij(x)orn1a1,ij(x)=n1a1,ij(x).n_{1}^{\prime}a_{1,ij}(x)=\lfloor n_{1}^{\prime}a^{\prime}_{1,ij}(x)\rfloor\quad\mbox{or}\quad n_{1}^{\prime}a_{1,ij}(x)=\lceil n_{1}^{\prime}a^{\prime}_{1,ij}(x)\rceil.

In particular, n1A1(x)n_{1}^{\prime}A_{1}(x) is integer-valued. Because a1,ij(x)=a1,ji(x)a^{\prime}_{1,ij}(x)=a^{\prime}_{1,ji}(x), it follows that

n1|a1,ij(x)a1,ji(x)|1,1i,jq.n_{1}^{\prime}|a_{1,ij}(x)-a_{1,ji}(x)|\leq 1,\quad\forall 1\leq i,j\leq q. (26)

Moreover, if a1,ij(x)>0a^{\prime}_{1,ij}(x)>0, then

a1,ij(x)>a1,ij(x)1n1α1n1α1nnn1α83nqα>(1112q)α.a_{1,ij}(x)>a^{\prime}_{1,ij}(x)-\frac{1}{n_{1}^{\prime}}\geq\alpha-\frac{1}{n_{1}^{\prime}}\geq\alpha-\frac{1}{n}\frac{n}{n_{1}^{\prime}}\geq\alpha-\frac{8}{3nq\alpha}>\left(1-\frac{1}{12q}\right)\alpha. (27)

where second to the last inequality follow from item 2 of Lemma 7 and the last inequality follows from the hypothesis on nn (specifically n>8αϵn>\frac{8}{\alpha\epsilon}) from the statement and the condition that ϵ<α/4\epsilon<\alpha/4 from Lemma 6.

Proof that AA satisfies the five items of the statement:

  1. 1.

    From (22), A0(x)𝟏=τ¯0(x)A_{0}(x){\bf 1}=\bar{\tau}^{\prime}_{0}(x). For A1A_{1}, it was shown that A1(x)𝟏=τ¯1(x)A_{1}(x){\bf 1}=\bar{\tau}^{\prime}_{1}(x) in (24) and (25) for Case 1 and Case 2, respectively. Since AA is the convex combination of A0A_{0} and A1A_{1} given in (21), it follows that

    A(x)𝟏=n0nτ¯0(x)+n1nτ¯1(x)=x.{A(x){\bf 1}}=\frac{n^{\prime}_{0}}{n}\bar{\tau}^{\prime}_{0}(x)+\frac{n_{1}^{\prime}}{n}\bar{\tau}^{\prime}_{1}(x)=x. (28)
  2. 2.

    By the construction of AA in (21) and the definitions of A0A_{0} and A1A_{1}, the diagonal of nA(x)nA(x) is

    n0A0(x)=n0Diagτ¯0(x)=nDiagτ0(x).n^{\prime}_{0}A_{0}(x)=n^{\prime}_{0}\operatorname{Diag}\bar{\tau}^{\prime}_{0}(x)=n\operatorname{Diag}\tau^{\prime}_{0}(x).

    By (19), all the entries of nτ0(x)n\tau^{\prime}_{0}(x) are even.

  3. 3.

    Using (19) again, we have that

    𝟏n(τ0(x)τ0(x))𝟏,-{\bf 1}\leq n(\tau_{0}(x)-\tau_{0}^{\prime}(x))\leq{\bf 1},

    from which it follows that

    ndiagA(x)τ0(x)=nτ0(x)τ0(x)1.n\|\operatorname{diag}A(x)-\tau_{0}(x)\|_{\infty}=n\|\tau^{\prime}_{0}(x)-\tau_{0}(x)\|\leq 1.
  4. 4.

    The off-diagonal entries aij(x)a_{ij}(x) of A(x)A(x) are those of n1nA1(x)\frac{n_{1}^{\prime}}{n}A_{1}(x), which we denoted by n1na1,ij(x)\frac{n_{1}^{\prime}}{n}a_{1,ij}(x). Thus,

    n|aij(x)aji(x)|=n1|a1,ij(x)a1,ji(x)|1,n|a_{ij}(x)-a_{ji}(x)|=n_{1}^{\prime}|a_{1,ij}(x)-a_{1,ji}(x)|\leq 1,

    where the last inequality is (26).

  5. 5.

    Case 1: SS does not have a self-loop: In this case, A(x)=A1(x)A(x)=A_{1}(x). By construction of A1A_{1}, suppA1(x)=S\operatorname{supp}A_{1}(x)=\vec{S}. If A1(x)A_{1}(x) is obtained via case 1 above, then, as argued after (24), its nonzero entries are bounded below by α/2\alpha/2. Otherwise, A1A_{1} is obtained via case 2 and its nonzero entries are lower bounded as shown in (27).

    Case 2: SS has at least one self-loop: In this case,

    suppA(x)=suppA0(x)suppA1(x).\operatorname{supp}A(x)=\operatorname{supp}A_{0}(x)\sqcup\operatorname{supp}A_{1}(x).

    By construction of A0A_{0} and item 1 of Lemma 7,

    suppA0(x)=suppDiagτ0(x)=suppDiagτ0(x)=S0,\operatorname{supp}A_{0}(x)=\operatorname{supp}\operatorname{Diag}\tau^{\prime}_{0}(x)=\operatorname{supp}\operatorname{Diag}\tau_{0}(x)=\vec{S}_{0}, (29)

    where the last equality follows from Lemma 4. Moreover, the nonzero entries of A0(x)A_{0}(x) are bounded below by 12(1ϵ/4)α\frac{1}{2}(1-\epsilon/4)\alpha. Also, by construction of A1A_{1},

    suppA1(x)=S1.\operatorname{supp}A_{1}(x)=\vec{S}_{1}. (30)

    Thus, by (29) and (30), suppA(x)=S\operatorname{supp}A(x)=\vec{S}. Finally, we verify that the nonzero entries of A(x)A(x) are uniformly bounded below by a positive number. By item 2 of Lemma 7, n0/nn^{\prime}_{0}/n and n1/nn_{1}^{\prime}/n are uniformly bounded below by positive numbers (note that |F0|1|F_{0}|\geq 1 in the current case). Thus, using (21), the nonzero entries of A(x)A(x) are also uniformly lower bounded by a positive number.

This completes the proof.  

2.4 Constructing a Hamiltonian decomposition from A(x)A(x)

In this subsection, we construct the map ρ~:A(x)H\tilde{\rho}:A(x)\mapsto H announced in Section 1, where A(x)A(x) will be taken from the statement of Proposition 2 and HH is a Hamiltonian decomposition in GnWG_{n}\sim W, with xx its empirical concentration vector. Throughout this subsection, we assume that WW is a binary step-graphon, i.e., WW is valued in {0,1}\{0,1\}.

Graphs sampled from a binary step-graphon have rather rigid structures as we will describe below. We refer to them as SS-multipartite graphs, see also Figure 4:

Definition 8 (SS-multipartite graph).

Let S=(U,F)S=(U,F) be an undirected graph, possibly with self-loops. An undirected graph GG is an SS-multipartite graph if there exists a graph homomorphism π:GS\pi:G\to S, so that

(vi,vj)E(π(vi),π(vj))F.(v_{i},v_{j})\in E\Rightarrow(\pi(v_{i}),\pi(v_{j}))\in F.

Further, GG is a complete SS-multipartite graph if

(vi,vj)E(π(vi),π(vj))F.(v_{i},v_{j})\in E\Leftrightarrow(\pi(v_{i}),\pi(v_{j}))\in F.
Refer to caption
(a)
Refer to caption
(b)
Figure 4: Left: An undirected graph SS on three nodes. Right: A complete SS-multipartite graph M(ω,S)M(\omega,S) with ω=(1,2,2)\omega=(1,2,2).

Let GG be an arbitrary complete SS-multipartite graph with S=(U,F)S=(U,F) and set ni:=|π1(ui)|n_{i}:=|\pi^{-1}(u_{i})| for i=1,,qi=1,\ldots,q. It should be clear that GG is completely determined by SS and the vector w:=(n1,,nq)w:=(n_{1},\ldots,n_{q}) . We will consequently use the notation M(w,S)M(w,S) to refer to a complete SS-multipartite graph. Now, returning to the case GnWG_{n}\sim W, where WW is a binary step-graphon with skeleton graph SS, the empirical concentration vector x(Gn)x(G_{n}) together with SS then completely determine GnG_{n} as announced above.

If GG is a (complete) SS-multipartite graph, then G\vec{G} is (complete) S\vec{S}-multipartite, and we use the same notation π\pi to denote the homomorphism. We next introduce a special class of cycles in G\vec{G}:

Definition 9 (Simple cycle).

Let GG be an SS-multipartite graph, and π:GS\pi:\vec{G}\to\vec{S} be the associated homomorphism. A directed cycle DD in G\vec{G} is called simple if π(D)\pi(D) is a cycle (rather than a closed walk) in S\vec{S}.

With the notions above, we state the main result of this subsection:

Proposition 3.

Let S=(U,F)S=(U,F) be an undirected graph, possibly with self-loops. Let G=M(nx,S)G=M(nx,S) be a complete SS-multipartite graph on nn nodes, where x𝒰x\in{\cal U}^{*} and nn is paired with xx (see Definition 7). Let A(x)A(x) be as in Proposition 2 and

mij(x):=nmin{aij(x),aji(x)}, for all 1i,jq.m_{ij}(x):=n\min\{a_{ij}(x),a_{ji}(x)\},\mbox{ for all }1\leq i,j\leq q. (31)

Then, there exists a Hamiltonian decomposition HH of G\vec{G}, with ρ(H)=A(x)\rho(H)=A(x), such that the following hold:

  1. 1.

    There exist exactly 12mii(x)\frac{1}{2}m_{ii}(x) disjoint 2-cycles in HH pairing mii(x)m_{ii}(x) nodes in π1(ui)\pi^{-1}(u_{i}) for every i=1,,qi=1,\ldots,q;

  2. 2.

    There are at least mij(x)m_{ij}(x) disjoint 2-cycles in HH pairing nodes in π1(ui)\pi^{-1}(u_{i}) to nodes in π1(uj)\pi^{-1}(u_{j}) for each (ui,uj)F1(u_{i},u_{j})\in F_{1}.

  3. 3.

    There are at most 23|F|\left\lceil\frac{2}{3}|F|\right\rceil cycles of length three or more in HH;

  4. 4.

    The length of every cycle of HH does not exceed 2|F|2|F|;

  5. 5.

    All cycles of length at least 33 of HH are simple.

We illustrate the Proposition on an example.

Example 1.

Consider a complete SS-multipartite graph GG for SS shown in Figure 4a. Set ni:=|π1(ui)|n_{i}:=|\pi^{-1}(u_{i})|, for i=1,2,3i=1,2,3, n:=i=13nin:=\sum_{i=1}^{3}n_{i}, and x:=1n(n1,n2,n3)x:=\frac{1}{n}(n_{1},n_{2},n_{3}). In this case, xintΔ2x\in\operatorname{int}\Delta^{2} if and only if the nin_{i}’s satisfy triangle inequalities ni+nj>nkn_{i}+n_{j}>n_{k}, where ii, jj, and kk are pairwise distinct. If these inequalities are satisfied, then G\vec{G} admits a Hamiltonian decomposition HH, which is comprised primarily (if not entirely) of 2-cycles. We plot in Figure 5 the corresponding undirected edges of GG. Specifically, there are two cases: (1) If n1n2+n3n_{1}-n_{2}+n_{3} is even, then HH is comprised solely of 2-cycles as shown in Figure 5a. (2) If n1n2+n3n_{1}-n_{2}+n_{3} is odd, then HH is comprised of 2-cycles and a single triangle as shown in Figure 5b.

Refer to caption
(a)
Refer to caption
(b)
Figure 5: Illustration of Example 1: Two complete SS-multipartite graphs GG with SS shown in Figure 4a, where the undirected edges plotted in each subfigure give rise to a Hamiltonian decomposition HH of G\vec{G}. Green nodes belong to π1(u1)\pi^{-1}(u_{1}), blue nodes to π1(u2)\pi^{-1}(u_{2}), and red nodes to π1(u3)\pi^{-1}(u_{3}).

The proof of Proposition 3 relies on a reduction argument for both the graph GnG_{n} and the matrix A(x)A(x): roughly speaking, we will first remove out of Gn\vec{G}_{n} a number of 2-cycles, which leads to a graph G\vec{G}^{\prime} of smaller size. With regards to the matrix AA, this reduction leads to another matrix A𝒜(S)A^{\prime}\in\mathcal{A}(S) with the property that diag(A)=0\operatorname{diag}(A^{\prime})=0. Finding a Hamiltonian decomposition HH for G\vec{G} with ρ(H)=A\rho(H)=A is then reduced to finding a Hamiltonian decomposition HH^{\prime} for G\vec{G}^{\prime} with ρ(H)=A\rho(H^{\prime})=A^{\prime}. For the arguments outlined above, we need a supporting lemma stated below, whose proof is relegated to Appendix C:

Lemma 8.

Let nn^{\prime} be a nonnegative integer. Let A𝒜(S)A^{\prime}\in\mathcal{A}(S) be such that nAn^{\prime}A^{\prime} is integer-valued, diag(A)=0\operatorname{diag}(A^{\prime})=0, and x:=A𝟏x^{\prime}:={A^{\prime}}{\bf 1}. Then, there exists a Hamiltonian decomposition HH^{\prime} of G\vec{G}^{\prime}, where G:=M(nx,S)G^{\prime}:=M(n^{\prime}x^{\prime},S) is the complete SS-multipartite graph, such that ρ(H)=A\rho(H^{\prime})=A^{\prime} and every cycle in HH^{\prime} is simple.

With the lemma above, we now establish Proposition 3:

Proof of Proposition 3.

We construct a Hamiltonian decomposition HH with the desired properties in two steps. We will fix xx in the proof and, to simplify the notation, we omit writing the argument xx for aij(x)a_{ij}(x), mij(x)m_{ij}(x), and A(x)A(x).

Step 1: We claim that the following selection of 2-cycles out of Gn\vec{G}_{n} is feasible:

  • For every self-loop (ui,ui)F0(u_{i},u_{i})\in F_{0}, mii=naiim_{ii}=na_{ii} is an even integer, and we select miim_{ii} pairwise distinct nodes in π1(ui)\pi^{-1}(u_{i}) that form mii/2m_{ii}/2 disjoint 2-cycles.

  • For every (ui,uj)F1(u_{i},u_{j})\in F_{1}, we select mijm_{ij} distinct nodes in π1(ui)\pi^{-1}(u_{i}) and mijm_{ij} distinct nodes in π1(uj)\pi^{-1}(u_{j}), to form mijm_{ij} disjoint 2-cycles (so the total number of such 2-cycles is (ui,uj)F1mij\sum_{(u_{i},u_{j})\in F_{1}}m_{ij}).

The above selection is feasible because (1) GnG_{n} is a complete SS-multipartite graph and, thus, there is an edge between any pair of nodes in π1(ui)\pi^{-1}(u_{i}), π1(uj)\pi^{-1}(u_{j}) provided that (ui,uj)F1(u_{i},u_{j})\in F_{1} and, (2) A𝟏=xA{\bf 1}=x which implies that j=1qmijni\sum^{q}_{j=1}m_{ij}\leq n_{i} and, hence, we can always pick the required number of distinct nodes.

Let VV^{\prime} be the set of remaining nodes in GG, i.e., VV^{\prime} is obtained by removing out of VV the (ui,uj)Fmij\sum_{(u_{i},u_{j})\in F}m_{ij} nodes picked in step 1. If VV^{\prime} is the empty set, we let HH be the union of the disjoint 2-cycles just exhibited. It should be clear that HH is a Hamiltonian decomposition of G\vec{G}. We claim that HH satisfies the desired properties. To see this, let A:=ρ(H)A^{\prime}:=\rho(H) and x:=A𝟏x^{\prime}:={A^{\prime}}{\bf 1}. Then, A𝒜(S)A^{\prime}\in\mathcal{A}(S) and, by construction of HH, naij=mijna^{\prime}_{ij}=m_{ij}. On the one hand, since VV^{\prime} is empty, nx1=nx1=n\|nx^{\prime}\|_{1}=\|nx\|_{1}=n and, hence, the sum of the entries of AA^{\prime} is equal to the sum of the entries of AA. On the other hand, since mijnaijm_{ij}\leq na_{ij}, we have that A=1n[mij]ijAA^{\prime}=\frac{1}{n}[m_{ij}]_{ij}\leq A. It then follows that A=AA^{\prime}=A and x=xx^{\prime}=x. Furthermore, items 1 and 2 follow from Step 1, respectively, and items 3, 4, and 5 hold trivially.

Step 2: We now assume that VV^{\prime} is non-empty. Let GG^{\prime} be the subgraph of GG induced by VV^{\prime}. We exhibit below a Hamiltonian decomposition HH^{\prime} for G\vec{G}^{\prime} such that the cycles in HH^{\prime}, together with the 2-cycles constructed in Step 1, yield a desired HH. Additionally, we will show that all the cycles of length at least 3 in HH^{\prime} satisfy items 3, 4, and 5.

To construct the above mentioned HH^{\prime}, we will appeal to Lemma 8. To this end, let nin^{\prime}_{i} be the number of nodes in π1(ui)V\pi^{-1}(u_{i})\cap V^{\prime}, i.e.,

ni=niujN(ui)mij.n^{\prime}_{i}=n_{i}-\sum_{u_{j}\in N(u_{i})}m_{ij}.

Let n:=i=1qnin^{\prime}:=\sum^{q}_{i=1}n^{\prime}_{i} be the total number of nodes in GG^{\prime}, and

x:=1n[n1;;nq].x^{\prime}:=\frac{1}{n^{\prime}}[n^{\prime}_{1};\cdots;n^{\prime}_{q}].

It follows that G=M(nx,S)G^{\prime}=M(n^{\prime}x^{\prime},S).

Because HH has to satisfy ρ(H)=A\rho(H)=A with A𝟏=xA{\bf 1}=x, HH should contain naijna_{ij} edges from nodes in π1(ui)\pi^{-1}(u_{i}) to nodes in π1(uj)\pi^{-1}(u_{j}). Since mijm_{ij} such edges have already been accounted for by the 2-cycles created in Step 1, we need an additional nijn^{\prime}_{ij} edges, where

nij:={naijmijif (ui,uj)F,0otherwise.n^{\prime}_{ij}:=\left\{\begin{array}[]{ll}na_{ij}-m_{ij}&\mbox{if $(u_{i},u_{j})\in F$},\\ 0&\mbox{otherwise}.\end{array}\right. (32)

Note that by (31), nii=0n^{\prime}_{ii}=0 for all i=1,,qi=1,\ldots,q. Correspondingly, we define a q×qq\times q matrix as follows:

A:=1n[nij].A^{\prime}:=\frac{1}{n^{\prime}}[n^{\prime}_{ij}].

Because mij=mjim_{ij}=m_{ji} for all (ui,uj)F(u_{i},u_{j})\in F and because A𝟏=A𝟏A{\bf 1}=A^{\top}{\bf 1}, we obtain that

j=1qnij=j=1qnji=ni,i=1,,q.\sum^{q}_{j=1}n^{\prime}_{ij}=\sum^{q}_{j=1}n^{\prime}_{ji}=n^{\prime}_{i},\quad\forall i=1,\ldots,q.

Thus, A𝒜(S)A^{\prime}\in\mathcal{A}(S) and, by construction, diagA=0\operatorname{diag}A^{\prime}=0 and A𝟏=x{A^{\prime}}{\bf 1}=x^{\prime}, so AA^{\prime} satisfies the conditions in the statement of Lemma 8.

By Lemma 8, there exists a Hamiltonian decomposition HH^{\prime} of G\vec{G}^{\prime} such that ρ(H)=A\rho(H^{\prime})=A^{\prime}, A𝟏=x{A^{\prime}}{\bf 1}=x^{\prime}, and all cycles in HH^{\prime} are simple. Now, let HH be the union of HH^{\prime} and the 2-cycles obtained in Step 1. Then,

ρ(H)=1n[mij+nij]=1n[naij]=A,\rho(H)=\frac{1}{n}[m_{ij}+n^{\prime}_{ij}]=\frac{1}{n}[na_{ij}]=A,

where the second equality follows from (32). Moreover, since diagA=0\operatorname{diag}A^{\prime}=0, there is no 22-cycle in HH^{\prime} connecting pairs of nodes in π1(ui)\pi^{-1}(u_{i}) for any i=1,,qi=1,\ldots,q. Thus, for each i=1,,qi=1,\ldots,q, HH contains exactly 12mii\frac{1}{2}m_{ii} disjoint two-cycles pairing miim_{ii} nodes in π1(ui)\pi^{-1}(u_{i}).

It now remains to show that all the cycles of length at least three in HH^{\prime} satisfy items 3 and 4. To do so, we first provide an upper bound on nin^{\prime}_{i}: Using items 3 and 4 of Proposition 2, we have that naijmij1na_{ij}-m_{ij}\leq 1. Thus,

niniujN(ui)(naij1)=nini+deg(ui)=deg(ui),n^{\prime}_{i}\leq n_{i}-\sum_{u_{j}\in N(u_{i})}(na_{ij}-1)=n_{i}-n_{i}+\deg(u_{i})=\deg(u_{i}),

where deg(ui)\deg(u_{i}) is the degree of uiu_{i} in SS. Since

i=1qdeg(ui)2|F|,\sum^{q}_{i=1}\deg(u_{i})\leq 2|F|,

there are at most 2|F|2|F| nodes in G\vec{G}^{\prime}. Consequently, the length of any cycle in HH^{\prime} is bounded above by 2|F|2|F| and, moreover, there exist at most 23|F|\left\lceil\frac{2}{3}|F|\right\rceil cycles of length three or more in HH^{\prime}. This completes the proof.  

Remark 3.

The fact that item 2 of the proposition provides a lower bound for the number of 2-cycles instead of an exact number can be understood as follows: The Hamiltonian decomposition HH^{\prime} of G\vec{G}^{\prime}, introduced in Step 2 of the proof, may contain additional 2-cycles pairing nodes from π1(ui)\pi^{-1}(u_{i}) to π1(uj)\pi^{-1}(u_{j}), for (ui,uj)F1(u_{i},u_{j})\in F_{1}.  

2.5 Proof of the Main Theorem

In Subsection 2.4, we dealt with the construction of a Hamiltonian decomposition in a graph Gn\vec{G}_{n} sampled from a binary step graphon. We will now extend the result to a general step-graphon WW, for which the existence of an edge between a pair of nodes is not a sure event. This will then complete the proof of the Main Theorem.

Refer to caption
Figure 6: A bipartite graph with VL={α1,,α4}V_{L}=\{\alpha_{1},\ldots,\alpha_{4}\} and VR={β1,,β5}V_{R}=\{\beta_{1},\ldots,\beta_{5}\}. The blue edges form a left-perfect matching.

To do so, we first recall some known facts about bipartite graphs. An undirected graph B=(V,E)B=(V,E) is called bipartite if its node set can be written as the disjoint union V=VLVRV=V_{L}\sqcup V_{R} so that there does not exist an edge between two nodes in VLV_{L} or VRV_{R}. Equivalently, a bipartite graph can be viewed as an SS-multipartite graph where SS is a graph with two nodes connected by a single edge. We refer to elements of VLV_{L} and VRV_{R} as left- and right-nodes, respectively. A left-perfect matching PP in BB is a set of edges so that each left-node is incident to exactly one edge in PP, and each right-node is incident to at most one edge in PP. See Figure 6 for an illustration. Similarly, we define a right-perfect matching by swapping the roles of left- and right-nodes. One can easily see that a left-perfect (resp. right-perfect) matching exists only if |VL||VR||V_{L}|\leq|V_{R}| (resp. |VL||VR||V_{L}|\geq|V_{R}|).

Further, we denote by B(n1,n2,p)B(n_{1},n_{2},p) an Erdős-Rényi random bipartite graph, with n1n_{1} left-nodes, n2n_{2} right-nodes, and edge probability pp for all edges between left- and right-nodes.

We need the following fact:

Lemma 9.

Let n1n_{1} and n2n_{2} be positive integers such that 1κn2n1κ\frac{1}{\kappa}\leq\frac{n_{2}}{n_{1}}\leq\kappa, where κ1\kappa\geq 1 is a constant. Let n:=n1+n2n:=n_{1}+n_{2} and p(0,1)p\in(0,1). Then, it holds a.s. that the random bipartite graph B(n1,n2,p)B(n_{1},n_{2},p) is connected and contains a left-perfect (resp. right-perfect) matching if n2n1n_{2}\geq n_{1} (resp. n1n2n_{1}\geq n_{2}).

The above lemma is certainly well known. For completeness of presentation, we present a proof in Appendix D.

We now return to the proof of Main Theorem. For the given step-graphon WW, we fix a partition σ\sigma, and let xx^{*} be the associated concentration vector and SS be the skeleton graph. We now consider a sequence of graphs GnWG_{n}\sim W, with nn\to\infty. We show below that the Hamiltonian decomposition HH for Gn\vec{G}_{n} described in Proposition 3 exists a.s..

Denote by WsW^{s} the saturation of WW: it is the binary step-graphon defined as

Ws(s,t)=1W(s,t)0.W^{s}(s,t)=1\Longleftrightarrow W(s,t)\neq 0.

We similarly construct a saturated version of Gn=(Vn,En)WG_{n}=(V_{n},E_{n})\sim W, denoted by Gns=(Vn,Ens)G_{n}^{s}=(V_{n},E_{n}^{s}), as follows: There is an edge (v,vk)Ens(v_{\ell},v_{k})\in E_{n}^{s} if and only if (π(v),π(vk))F(\pi(v_{\ell}),\pi(v_{k}))\in F. Said otherwise, the node set of GnG_{n} and GnsG_{n}^{s} are the same, but the edges in GnsG_{n}^{s} are obtained using the binary step-graphon WsW^{s}. It should be clear that GnGns=M(nx(Gn),S)G_{n}\subseteq G_{n}^{s}=M(nx(G_{n}),S), where we recall that x(Gn)x(G_{n}) is the empirical concentration vector of GnG_{n} defined in (5).

Let 𝒰{\cal U} be the closed neighborhood of xx^{*} mentioned in Remark 2. Let 0\mathcal{E}_{0} be the event that the empirical concentration vector x(Gn)x(G_{n}) of GnG_{n} belongs to 𝒰{\cal U}. By Chebyshev’s inequality, we have that

(x(Gn)x>ϵ)cn2ϵ2,\mathbb{P}(\|x(G_{n})-x^{*}\|>\epsilon)\leq\frac{c}{n^{2}\epsilon^{2}},

which implies that 0\mathcal{E}_{0} is almost sure as nn\to\infty. Thus, we can assume in the sequel that 0\mathcal{E}_{0} is true, i.e., the analysis and computation carried out below are conditioned upon 0\mathcal{E}_{0}.

Note that nx(Gn)nx(G_{n}) is always integer-valued. Since x(Gn)𝒰x(G_{n})\in{\cal U} by assumption, we let nn be sufficiently large so that nn is paired with x(Gn)x(G_{n}) (see Definition 7). We can thus appeal to Proposition 2 to obtain a matrix A(x(Gn))A(x(G_{n})), and to Proposition 3 to obtain a corresponding Hamiltonian decomposition HH of Gns\vec{G}_{n}^{s}. We now demonstrate that the same HH exists a.s. in Gn\vec{G}_{n}, up to re-labeling of the nodes of Gn\vec{G}_{n}. The proof comprises two parts: In part 1, we show that the cycles in HH whose lengths are greater than 22 exist a.s. in Gn\vec{G}_{n} and, then, in part 2, we show that the 2-cycles of HH do as well.

Part 1: On cycles of length greater than 2. For clarity of presentation, we denote by πs:GnsS\pi^{s}:G_{n}^{s}\to S the graph homomorphisms associated with GnsG^{s}_{n}. For any path u1uku_{1}\cdots u_{k} in SS, since GnsG^{s}_{n} is complete SS-multipartite, there surely exists a path v1vkv_{1}\cdots v_{k} in GnsG^{s}_{n} so that πs(vi)=ui\pi^{s}(v_{i})=u_{i}. The following result shows that such a path exist in GnG_{n} a.s..

Lemma 10.

Let u1uku_{1}\cdots u_{k} be a path in SS. Then, it is a.s. that there exists a path v1vkv_{1}\cdots v_{k} in GnG_{n}, with π(vi)=ui\pi(v_{i})=u_{i}.

Proof.

Since the closed set 𝒰{\cal U} is in the interior of Δq1\Delta^{q-1}, there exists a κ1\kappa\geq 1 such that for all x𝒰x\in{\cal U},

1κxixjκ, for all 1i,jq.\frac{1}{\kappa}\leq\frac{x_{i}}{x_{j}}\leq\kappa,\mbox{ for all }1\leq i,j\leq q.

Thus, by conditioning on 0\mathcal{E}_{0}, we have that

1κxi(Gn)xj(Gn)=|π1(ui)||π1(uj)|κ, for all 1i,jq.\frac{1}{\kappa}\leq\frac{x_{i}(G_{n})}{x_{j}(G_{n})}=\frac{|\pi^{-1}(u_{i})|}{|\pi^{-1}(u_{j})|}\leq\kappa,\mbox{ for all }1\leq i,j\leq q.

It then follows that the subgraphs of GnG_{n} induced by π1(ui)π1(ui+1)\pi^{-1}(u_{i})\cup\pi^{-1}(u_{i+1}) are bipartite and satisfy the hypothesis of Lemma 9, for 1ik11\leq i\leq k-1. Hence, it is a.s. that all of these bipartite graphs are connected. We now pick an arbitrary node v1π1(u1)v_{1}\in\pi^{-1}(u_{1}); by the above arguments, we can find v2π1(u2)v_{2}\in\pi^{-1}(u_{2}) so that (v1,v2)Gn(v_{1},v_{2})\in G_{n} a.s.. Iterating this procedure, we obtain the path in GnG_{n} sought.  

Now, let D1,,DmD_{1},\ldots,D_{m} be the cycles in HH whose lengths are greater than 22, and C1,,CmC_{1},\ldots,C_{m} be the corresponding undirected cycles in GnsG_{n}^{s}. From items 3 and 4 of Proposition 3, the number mm of these cycles, as well as their lengths, are each uniformly bounded above by constants independent of nn.

Let 1\mathcal{E}_{1} be the event that the cycles C1,,CmC_{1},\ldots,C_{m} exist in GnG_{n}; more precisely, it is the event that there exist disjoint cycles CiC^{\prime}_{i} in GnG_{n} such that π(Ci)=πs(Ci)\pi(C^{\prime}_{i})=\pi^{s}(C_{i}) for all i=1,,mi=1,\ldots,m. We have the following lemma:

Lemma 11.

The event 1\mathcal{E}_{1} is true a.s..

Proof.

Let 11\mathcal{E}_{11} be the event that there exists a cycle C1GnC^{\prime}_{1}\in G_{n} with π(C1)=πs(C1)\pi(C^{\prime}_{1})=\pi^{s}(C_{1}). We show that 11\mathcal{E}_{11} holds a.s.. To start, we write explicitly πs(C1)=u1uku1\pi^{s}(C_{1})=u_{1}\ldots u_{k}u_{1}. Since C1C_{1} is simple, u1uku1u_{1}\ldots u_{k}u_{1} is a cycle in SS. By Lemma 10, there exist a.s. nodes viπ1(ui)v_{i}\in\pi^{-1}(u_{i}), for 1ik11\leq i\leq k-1, such that v1vk1v_{1}\cdots v_{k-1} is a path in GnG_{n}.

In order to obtain the cycle C1C^{\prime}_{1}, it remains to exhibit a node vkπ1(uk)v_{k}\in\pi^{-1}(u_{k}) that is connected to both v1v_{1} and vk1v_{k-1} in GnG_{n}. We claim that such a node exists with probability at least

1(1p1kpk1,k)|π1(uk)|1-(1-p_{1k}p_{k-1,k})^{|\pi^{-1}(u_{k})|} (33)

where pij>0p_{ij}>0 is the value of the step-graphon WW over the rectangle [σi1,σi)×[σj1,σj)[\sigma_{i-1},\sigma_{i})\times[\sigma_{j-1},\sigma_{j}). The claim holds because the probability that no node of π1(uk)\pi^{-1}(u_{k}) connects to both v1v_{1} and vk1v_{k-1} is given by (1p1kpk1,k)|π1(uk)|(1-p_{1k}p_{k-1,k})^{|\pi^{-1}(u_{k})|}. Thus, the probability of the complementary event is give by (33).

Next, recall that a¯>0\underline{a}>0 is the uniform lower bound for the nonzero entries of A(x)A(x), for all x𝒰x\in{\cal U}^{*}, introduced in item 5 of Proposition 2. Because x(Gn)=A(x(Gn))𝟏x(G_{n})=A(x(G_{n})){\bf 1}, every entry of x(Gn)x(G_{n}) is bounded below by a¯\underline{a} as well, so

|π1(ui)|a¯n, for all i=1,,q.|\pi^{-1}(u_{i})|\geq\underline{a}n,\mbox{ for all }i=1,\ldots,q. (34)

Thus, the expression (33) can be lower bounded by

1(1p1kpk1,k)|π1(uk)|1(1p2)a¯n,1-(1-p_{1k}p_{k-1,k})^{|\pi^{-1}(u_{k})|}\geq 1-(1-p^{2})^{\underline{a}n},

where p:=min{pij(ui,uj)F}>0p:=\min\{p_{ij}\mid(u_{i},u_{j})\in F\}>0. Note that the right-hand-side of the above equation converges to 11 as nn\to\infty, so 11\mathcal{E}_{11} is true a.s..

Let n:=n|C1|n^{\prime}:=n-|C_{1}|. Conditioning on the event 11\mathcal{E}_{11}, we let GnG^{\prime}_{n^{\prime}} be the subgraph of GnG_{n} induced by the nodes not in C1C^{\prime}_{1}. Similarly as above, we have that there is a cycle C2C^{\prime}_{2}, with π(C2)=π(C2)\pi(C^{\prime}_{2})=\pi(C_{2}), in GnG^{\prime}_{n^{\prime}} a.s. (note that nn\to\infty implies nn^{\prime}\to\infty). Iterating this argument for finitely many steps, we have that 1\mathcal{E}_{1} is true a.s..  

In the sequel, we condition on the event 1\mathcal{E}_{1} and let D1,,DmD^{\prime}_{1},\ldots,D^{\prime}_{m} be the directed cycles in Gn\vec{G}_{n} corresponding to D1,DmD_{1},\ldots D_{m} in HH of Gns\vec{G}^{s}_{n}.

Part 2: On 2-cycles. Let n:=ni=1m|Ci|n^{\prime}:=n-\sum^{m}_{i=1}|C^{\prime}_{i}|, and GnG^{\prime}_{n^{\prime}} be the subgraph of GnG_{n} induced by the nodes that do not belong to any of the above cycles CiC^{\prime}_{i}, and GnsG^{\prime s}_{n^{\prime}} be its saturation. By removing the cycles DiD_{i} out of HH, we obtain a Hamiltonian decomposition HH^{\prime} for Gns\vec{G}^{\prime s}_{n^{\prime}}, which is comprised only of 2-cycles. It now suffices to show that HH^{\prime} appears, up to relabeling, in Gn\vec{G}^{\prime}_{n^{\prime}} a.s..

Let Vijπ1(ui)V_{ij}\subset\pi^{-1}(u_{i}) be the set of nodes paired to nodes in π1(uj)\pi^{-1}(u_{j}) by HH^{\prime} in Gns\vec{G}^{\prime s}_{n^{\prime}}. Since HH^{\prime} is a Hamiltonian decomposition, π1(ui)\pi^{-1}(u_{i}) can be expressed as the disjoint union of the VijV_{ij}’s, for uju_{j} such that (ui,uj)F(u_{i},u_{j})\in F. By items 1 and 2 of Proposition 3, the cardinality of VijV_{ij}, which is the same as the cardinality of VjiV_{ji}, is at least mij:=nmin{aij,aji}m_{ij}:=n\min\{a_{ij},a_{ji}\}. Because the nonzero aija_{ij}’s are bounded below by a¯\underline{a} by item 5 of Proposition 2, we have that mija¯nm_{ij}\geq\underline{a}n.

Suppose that uiu_{i} has a self-loop; then, we let KiK_{i} be the subgraph of GnG^{\prime}_{n^{\prime}} induced by the nodes ViiV_{ii}. The graph KiK_{i} is an Erdős-Rényi graph with parameter pii>0p_{ii}>0 and, by item 1 of Proposition 3, |Vii|=mii|V_{ii}|=m_{ii} is an even integer. Since nn\to\infty implies that miim_{ii}\to\infty, KiK_{i} has a perfect matching a.s.. This holds because one can split the node set ViiV_{ii} into two disjoint subsets of equal cardinality and apply Lemma 9. In other words, it is a.s.. that there are mii/2m_{ii}/2, for i=1,,qi=1,\ldots,q, disjoint 2-cycles in Gn\vec{G}^{\prime}_{n^{\prime}} pairing nodes in π1(ui)\pi^{-1}(u_{i}).

Suppose that (ui,uj)(u_{i},u_{j}) is an edge between two distinct nodes; then, we let Bij:=B(|Vij|,|Vji|,pij)B_{ij}:=B(|V_{ij}|,|V_{ji}|,p_{ij}) be the bipartite graph in GnG^{\prime}_{n^{\prime}} induced by VijVjiV_{ij}\cup V_{ji} (recall from above that |Vij|=|Vji||V_{ij}|=|V_{ji}|). Let ij\mathcal{E}_{ij} be the event that BijB_{ij} has a perfect matching. Since mija¯nm_{ij}\geq\underline{a}n, by Lemma 9, the event ij\mathcal{E}_{ij} holds a.s. and, hence, it is a.s. that there are |Vij||V_{ij}| disjoint 2-cycles in Gn\vec{G}^{\prime}_{n^{\prime}} pairing nodes from VijV_{ij} to VjiV_{ji}.

Since there are finitely many edges in SS, by the above arguments, we conclude that HH^{\prime} appears in Gn\vec{G}^{\prime}_{n^{\prime}} a.s.. This completes the proof.  

3 Conclusions

Hamiltonian decompositions underlie a wide range of structural properties of control systems, such as stability and ensemble controllability. We say that a graphon WW satisfies the HH-property if graphs GnWG_{n}\sim W have a Hamiltonian decomposition almost surely. In a series of papers, of which this is the second, we exhibited necessary and sufficient conditions for the HH-property to hold for the class of step-graphons. These conditions are geometric and revealed the fact that HH-property depends only on concentration vector and skeleton graph of WW. When these two objects are given, one can reconstruct a step-graphon modulo the exact value of WW on its support, thus giving rise to an equivalence relation on the space of step-graphons. We showed that the HH-property is essentially a “zero-one” property of the equivalence classes. The case of general graphons will be addressed in future work.

References

  • [1] M.-A. Belabbas, X. Chen, and T. Başar, “On the HH-property for step-graphons and edge polytopes,” IEEE Control Systems Letters, vol. 6, pp. 1766–1771, 2022.
  • [2] X. Chen, “Sparse linear ensemble systems and structural controllability,” IEEE Transactions on Automatic Control, 2021. Appeared online.
  • [3] M.-A. Belabbas, “Algorithms for sparse stable systems,” in Proceedings of the 52th IEEE Conference on Decision and Control, 2013.
  • [4] M.-A. Belabbas, “Sparse stable systems,” Systems & Control Letters, vol. 62, no. 10, pp. 981–987, 2013.
  • [5] S. Gao and P. E. Caines, “Graphon control of large-scale networks of linear systems,” IEEE Transactions on Automatic Control, vol. 65, no. 10, pp. 4090–4105, 2019.
  • [6] S. Gao, R. F. Tchuendom, and P. E. Caines, “Linear quadratic graphon field games,” Communications in Information and Systems, vol. 21, pp. 341–369, 2021.
  • [7] F. Parise and A. Ozdaglar, “Analysis and interventions in large network games,” Annual Review of Control, Robotics, and Autonomous Systems, vol. 4, pp. 455–486, 2021.
  • [8] L. Lovász and B. Szegedy, “Limits of dense graph sequences,” Journal of Combinatorial Theory, Series B, vol. 96, no. 6, pp. 933–957, 2006.
  • [9] C. Borgs, J. T. Chayes, L. Lovász, V. T. Sós, and K. Vesztergombi, “Convergent sequences of dense graphs I: Subgraph frequencies, metric properties and testing,” Advances in Mathematics, vol. 219, no. 6, pp. 1801–1851, 2008.
  • [10] H. Ohsugi and T. Hibi, “Normal polytopes arising from finite graphs,” Journal of Algebra, vol. 207, no. 2, pp. 409–426, 1998.
  • [11] M.-A. Belabbas, X. Chen, and T. Başar, “The HH-property for line graphons,” in Proceedings of the 1313th Asian Control Conference, 2021.
  • [12] C. Van Nuffelen, “On the incidence matrix of a graph,” IEEE Transactions on Circuits and Systems, vol. 23, no. 9, pp. 572–572, 1976.
  • [13] J. R. Munkres, Analysis on Manifolds. CRC Press, 2018.
  • [14] M.-A. Belabbas and X. Chen, “On integer balancing of directed graphs,” Systems & Control Letters, vol. 154, p. 104980, 2021.
  • [15] P. Erdős and A. Rényi, “On random matrices,” Publ. Math. Inst. Hungar. Acad. Sci., vol. 8, pp. 455–461, 1964.
  • [16] A. Frieze and M. Karoński, Introduction to Random Graphs. Cambridge University Press, 2016.

Appendix A Analysis and Proof of Proposition 1

We first have some preliminaries about refinements of partitions: given a partition σ\sigma, a refinement σ\sigma^{\prime} of σ\sigma, denoted by σσ\sigma\prec\sigma^{\prime}, is any sequence that has σ\sigma as a proper subsequence. For example, σ=(0,1/2,3/4,1)\sigma^{\prime}=(0,1/2,3/4,1) is a refinement of σ=(0,1/2,1)\sigma=(0,1/2,1). Given a step-graphon WW, if σ\sigma is a partition for WW, then so is σ\sigma^{\prime}.

We say that σ\sigma^{\prime} is a one-step refinement of σ\sigma if it is a refinement with |σ|=|σ|+1|\sigma^{\prime}|=|\sigma|+1. Any refinement of σ\sigma can be obtained by iterating one-step refinements. To fix ideas, and without loss of generality, we consider the refinement of σ=(σ0,,σq,σ)\sigma=(\sigma_{0},\ldots,\sigma_{q},\sigma_{*}) to σ=(σ0,,σq,σq+1,σ)\sigma^{\prime}=(\sigma_{0},\ldots,\sigma_{q},\sigma_{q+1},\sigma_{*}) with σq<σq+1<σ\sigma_{q}<\sigma_{q+1}<\sigma_{*}. If S=(U,F)S=(U,F), then S=(U,F)S^{\prime}=(U^{\prime},F^{\prime}), the skeleton graph of WW for σ\sigma^{\prime}, is given by

{U=U{uq+1},F=F{(ui,uq+1)(ui,uq)F}{(uq+1,uq+1) if (uq,uq)F}.\left\{\begin{aligned} U^{\prime}&=U\cup\{u_{q+1}\},\\ F^{\prime}&=F\cup\{(u_{i},u_{q+1})\mid(u_{i},u_{q})\in F\}\cup\{(u_{q+1},u_{q+1})\mbox{ if }(u_{q},u_{q})\in F\}.\end{aligned}\right. (35)

In essence, the node uq+1u_{q+1} is a copy of the node uqu_{q}. If there is a loop (uq,uq)(u_{q},u_{q}) in FF, then uqu_{q} and uq+1u_{q+1} are also connected and each has a self-loop. See Figure 7 for illustration. We say that a one-step refinement splits a node (here, uqu_{q}).

Refer to caption
(a)
Refer to caption
(b)
Figure 7: The graph SS^{\prime} on the right is obtained from the left SS via a one-step refinement. The node u4u_{4} in SS is split into u4u_{4} and u5u_{5}. Because g=(u1,u4)g=(u_{1},u_{4}) is an edge in SS, there exist two edges g=(u1,u4)g^{\prime}=(u_{1},u_{4}) and h=(u1,u5)h^{\prime}=(u_{1},u_{5}) in SS^{\prime}. Because u4u_{4} has a self-loop 4\ell_{4} in SS, both u4u_{4} and u5u_{5} have self-loops in SS^{\prime}, denoted by 4\ell^{\prime}_{4} and 5\ell^{\prime}_{5}, respectively. In addition, we have the edge k=(u4,u5)k^{\prime}=(u_{4},u_{5}) in SS^{\prime}.

We now prove Proposition 1:

Proof of Proposition 1.

Let σ\sigma and σ\sigma^{\prime} be as given in the statement of the proposition. It should be clear that there exists another partition σ′′\sigma^{\prime\prime} which is a refinement of both σ\sigma^{\prime} and σ\sigma and that σ′′\sigma^{\prime\prime} can be obtained via a sequence of one-step refinements starting with either σ\sigma^{\prime} or σ\sigma. Thus, combining the arguments at the beginning of the section, we can assume, without loss of generality, that σ\sigma^{\prime} is a one-step refinement of σ\sigma obtained by splitting the node u1Uu_{1}\in U.

Let xx^{*} and xx^{\prime*} be the concentration vectors for σ\sigma and σ\sigma^{\prime}, SS and SS^{\prime} be the corresponding skeleton graphs, and ZZ and ZZ^{\prime} be the corresponding incidence matrices. Note that ZZ^{\prime} has one more row than ZZ does due to the addition of the new node uq+1u_{q+1}; here, we let the last row of ZZ^{\prime} correspond to that node. It should be clear that ZZ^{\prime} contains ZZ as a submatrix. For clarity of presentation, we use ff (resp. ff^{\prime}) to denote edges of SS (resp. SS^{\prime}). Since the graph SS can be realized as a subgraph of SS^{\prime} in a natural way, we will write on occasion fFf^{\prime}\in F if ff^{\prime} is an edge of SS.

We now prove the invariance of each item listed in the statement of Proposition 1 under one-step refinements. The proofs of the first two items are direct consequence of the definition of one-step refinement.

Proof for item 1): If SS is connected, then from (35) we obtain that there exists a path from any node uiFu_{i}\in F to the new node uq+1u_{q+1}, so SS^{\prime} is also connected. Reciprocally, assume that SS has at least two connected components. Then, the node uq+1u_{q+1} obtained by splitting uqu_{q} will only be connected to nodes in the same component as uqu_{q} by definition of FF^{\prime}.

Proof for item 2): If SS has an odd cycle, then so does SS^{\prime} by (35). Reciprocally, we assume that SS is lacking an odd cycle. We show that SS^{\prime} has no odd cycle. Suppose, to the contrary, that it does. The cycle must then contain the node uq+1u_{q+1}. Replacing uq+1u_{q+1} with uqu_{q} yields a closed walk of odd length in SS. Since a closed walk can be decomposed edge-wise into a union of cycles and since the length of the walk is the sum of the lengths of the constituent cycles, there must exist an odd cycle in SS, which is a contradiction.

Proof for item 3): We prove each direction of the statement separately:

Part 1: x𝒳(S)x𝒳(S)x\in\mathcal{X}(S)\Rightarrow x^{\prime}\in\mathcal{X}(S^{\prime}) (xint𝒳(S)xint𝒳(S)x\in\operatorname{int}\mathcal{X}(S)\Rightarrow x^{\prime}\in\operatorname{int}\mathcal{X}(S^{\prime})): For ease of presentation, we let zfz_{f} (resp. zfz^{\prime}_{f^{\prime}}) be the edge of ZZ (resp. ZZ^{\prime}) corresponding to the element fFf\in F (resp. fFf^{\prime}\in F^{\prime}), and zf,iz_{f,i} be the iith entry of zfz_{f}. Because 𝒳(S)\mathcal{X}(S) is the convex hull of the columns of ZZ, there exist coefficients cf0c_{f}\geq 0, for fFf\in F, such that x=fFcfzfx=\sum_{f\in F}c_{f}z_{f}. If, further, xint𝒳(S)x\in\operatorname{int}\mathcal{X}(S), then these coefficients can be chosen to be strictly positive. We will use cfc_{f} to construct cf0c^{\prime}_{f^{\prime}}\geq 0, for fFf^{\prime}\in F^{\prime}, such that

x=fFcfzfx^{\prime}=\sum_{f^{\prime}\in F^{\prime}}c^{\prime}_{f^{\prime}}z^{\prime}_{f^{\prime}} (36)

and show that xint𝒳(S)x^{\prime}\in\operatorname{int}\mathcal{X}(S^{\prime}) if xint𝒳(S)x\in\operatorname{int}\mathcal{X}(S).

To proceed, let FuqF_{u_{q}} be the set of edges incident to node uqu_{q} in SS. Similarly, let FuqF^{\prime}_{u_{q}} and Fuq+1F^{\prime}_{u_{q+1}} be the sets of edges incident to uqu_{q} and uq+1u_{q+1} in SS^{\prime}, respectively. The coefficients cfc^{\prime}_{f^{\prime}} are defined as follows:

  1. a)

    If fFuqFuq+1f^{\prime}\notin F^{\prime}_{u_{q}}\cup F^{\prime}_{u_{q+1}}, then fFf^{\prime}\in F. Let cf:=cfc^{\prime}_{f^{\prime}}:=c_{f^{\prime}}.

  2. b)

    If fFuqf^{\prime}\in F^{\prime}_{u_{q}} and f(uq,uq+1)f^{\prime}\neq(u_{q},u_{q+1}), then fFf^{\prime}\in F. Let cf:=σq+1σqσσqcfc^{\prime}_{f^{\prime}}:=\frac{\sigma_{q+1}-\sigma_{q}}{\sigma_{*}-\sigma_{q}}c_{f^{\prime}}.

  3. c)

    If f=(ui,uq+1)f^{\prime}=(u_{i},u_{q+1}) and uiuqu_{i}\neq u_{q}, then we pick the fFf\in F such that

    f={(ui,uq) if uiuq+1,(uq,uq) if ui=uq+1.f=\begin{cases}(u_{i},u_{q})&\mbox{ if }u_{i}\neq u_{q+1},\\ (u_{q},u_{q})&\mbox{ if }u_{i}=u_{q+1}.\end{cases}

    Let cf:=σσq+1σσqcfc^{\prime}_{f^{\prime}}:=\frac{\sigma_{*}-\sigma_{q+1}}{\sigma_{*}-\sigma_{q}}c_{f}.

  4. d)

    If f=(uq,uq+1)f^{\prime}=(u_{q},u_{q+1}), then let cf:=0c^{\prime}_{f^{\prime}}:=0.

With the coefficients as above, we prove entry-wise that (36) holds. First, note that because we obtain SS^{\prime} by splitting the last node uqu_{q} of SS, the iith entry of xx^{\prime}, for 1iq11\leq i\leq q-1, is equal to xix_{i}, so xi=xi=(σiσi1)x^{\prime}_{i}=x_{i}=(\sigma_{i}-\sigma_{i-1}). For the iith entry of the right hand side of (36), we consider two cases:

Case 1: uiu_{i} is not incident to uqu_{q} in SS. In this case, uiu_{i} is not incident to either uqu_{q} or uq+1u_{q+1} in SS^{\prime}. Consequently, Fui=FuiF^{\prime}_{u_{i}}=F_{u_{i}} and zf,i=zf,iz^{\prime}_{f,i}=z_{f,i} for all fFuif\in F_{u_{i}}. Furthermore, by item (a), cf=cfc^{\prime}_{f}=c_{f} for any fFuif\in F_{u_{i}}. Thus, the iith entry of the right hand side of (36) is given by

fFuicfzf,i=fFuicfzf,i=xi=σiσi1.\sum_{f^{\prime}\in F^{\prime}_{u_{i}}}c^{\prime}_{f^{\prime}}z^{\prime}_{f^{\prime},i}=\sum_{f\in F_{u_{i}}}c_{f}z_{f,i}=x_{i}=\sigma_{i}-\sigma_{i-1}.

Case 2: uiu_{i} is incident to uqu_{q} in SS. In this case, uiu_{i} is incident to both uqu_{q} and uq+1u_{q+1} in SS^{\prime}. Let g:=(ui,uq)g^{\prime}:=(u_{i},u_{q}) and h:=(ui,uq+1)h^{\prime}:=(u_{i},u_{q+1}) be the corresponding edges in SS^{\prime}, see Fig. 7 for an illustration. Then, the iith entry of the right hand side of (36) is given by

fFuicfzf,i=cgzg,i+chzh,i+fFui{g,h}cfzf,i.\sum_{f^{\prime}\in F^{\prime}_{u_{i}}}c^{\prime}_{f^{\prime}}z^{\prime}_{f^{\prime},i}=c^{\prime}_{g^{\prime}}z^{\prime}_{g^{\prime},i}+c^{\prime}_{h^{\prime}}z^{\prime}_{h^{\prime},i}+\sum_{f^{\prime}\in F^{\prime}_{u_{i}}-\{g^{\prime},h^{\prime}\}}c^{\prime}_{f^{\prime}}z^{\prime}_{f^{\prime},i}. (37)

By items (b) and (c),

cg=σq+1σqσσqcgandch=σσq+1σσqch.c^{\prime}_{g^{\prime}}=\frac{\sigma_{q+1}-\sigma_{q}}{\sigma_{*}-\sigma_{q}}c_{g^{\prime}}\quad\mbox{and}\quad c^{\prime}_{h^{\prime}}=\frac{\sigma_{*}-\sigma_{q+1}}{\sigma_{*}-\sigma_{q}}c_{h^{\prime}}.

Also, note that

zg,i=zh,i=zg,i=12.z^{\prime}_{g^{\prime},i}=z^{\prime}_{h^{\prime},i}=z_{g^{\prime},i}=\frac{1}{2}.

Thus, the sum of the first two terms on the right hand side of (37) is cgzg,ic_{g^{\prime}}z_{g^{\prime},i}. For the last term, note that Fui{g,h}=Fui{g}F^{\prime}_{u_{i}}-\{g^{\prime},h^{\prime}\}=F_{u_{i}}-\{g^{\prime}\}. Also, by item (a) and the fact that zf,i=zf,iz^{\prime}_{f,i}=z_{f,i} for any fFuif\in F_{u_{i}},

fFui{g,h}cfzf,i=fFui{g}cfzf,i.\sum_{f^{\prime}\in F^{\prime}_{u_{i}}-\{g^{\prime},h^{\prime}\}}c^{\prime}_{f^{\prime}}z^{\prime}_{f^{\prime},i}=\sum_{f\in F_{u_{i}}-\{g^{\prime}\}}c_{f}z_{f,i}.

Combining the above arguments, we have that the right hand side of (37) is given by

fFuicfzf,i=xi=σiσi1.\sum_{f\in F_{u_{i}}}c_{f}z_{f,i}=x_{i}=\sigma_{i}-\sigma_{i-1}.

Next, the qqth entry of xx^{\prime} is (σq+1σq)(\sigma_{q+1}-\sigma_{q}) and the qqth entry of the right hand side of (36) is

fFuqcfzf,q\displaystyle\sum_{f^{\prime}\in F^{\prime}_{u_{q}}}c^{\prime}_{f^{\prime}}z^{\prime}_{f^{\prime},q} =σq+1σqσσqfFuqcfzf,q\displaystyle=\frac{\sigma_{q+1}-\sigma_{q}}{\sigma_{*}-\sigma_{q}}\sum_{f\in F_{u_{q}}}c_{f}z_{f,q}
=σq+1σqσσqxq=σq+1σq,\displaystyle=\frac{\sigma_{q+1}-\sigma_{q}}{\sigma_{*}-\sigma_{q}}x_{q}=\sigma_{q+1}-\sigma_{q},

where the first equality follows from the fact that

Fuq=Fuq{(uq,uq+1) if (uq,uq)F},F^{\prime}_{u_{q}}=F_{u_{q}}\cup\{(u_{q},u_{q+1})\mbox{ if }(u_{q},u_{q})\in F\},

items (b) and (d), and the last equality follows from the fact that xq=σσqx_{q}=\sigma_{*}-\sigma_{q}.

The last (i.e., (q+1)(q+1)th) entry of xx^{\prime} is (σσq+1)(\sigma_{*}-\sigma_{q+1}). The last entry of the right hand side of (36) is given by

fFuq+1cfzf,q+1\displaystyle\sum_{f^{\prime}\in F^{\prime}_{u_{q+1}}}c^{\prime}_{f^{\prime}}z^{\prime}_{f^{\prime},q+1} =σσq+1σσqfFuqcfzf,q\displaystyle=\frac{\sigma_{*}-\sigma_{q+1}}{\sigma_{*}-\sigma_{q}}\sum_{f\in F_{u_{q}}}c_{f}z_{f,q}
=σσq+1σσqxq=σσq+1,\displaystyle=\frac{\sigma_{*}-\sigma_{q+1}}{\sigma_{*}-\sigma_{q}}x_{q}=\sigma_{*}-\sigma_{q+1},

where the first equality follows from item (c) above. We have thus shown that (36) holds. In particular, since cfc^{\prime}_{f^{\prime}} are nonnegative by construction, (36) implies that x𝒳(S)x^{\prime}\in\mathcal{X}(S^{\prime}).

It now remains to show that if xint𝒳(S)x\in\operatorname{int}\mathcal{X}(S), then xint𝒳(S)x^{\prime}\in\operatorname{int}\mathcal{X}(S^{\prime}). Assuming xint𝒳(S)x\in\operatorname{int}\mathcal{X}(S), if uqu_{q} does not have a self-loop in SS, then the edge (uq,uq+1)(u_{q},u_{q+1}) does not exist in SS^{\prime}, so by items (a), (b), and (c), all coefficients cfc^{\prime}_{f^{\prime}} are positive, which implies that xint𝒳(S)x^{\prime}\in\operatorname{int}\mathcal{X}(S^{\prime}).

We now assume that uqu_{q} has a self-loop in SS. Then, k:=(uq,uq+1)k^{\prime}:=(u_{q},u_{q+1}) is an edge in SS^{\prime} (see Fig. 7 for an illustration), and thus ck=0c^{\prime}_{k^{\prime}}=0 per item (d) above. In this case, both uqu_{q} and uq+1u_{q+1} have self-loops in SS^{\prime}. Denote these two self-loops by q:=(uq,uq)\ell^{\prime}_{q}:=(u_{q},u_{q}) and q+1:=(uq+1,uq+1)\ell^{\prime}_{q+1}:=(u_{q+1},u_{q+1}). By (3), we have that

zk=12(zq+zq+1).z^{\prime}_{k^{\prime}}=\frac{1}{2}(z^{\prime}_{\ell^{\prime}_{q}}+z^{\prime}_{\ell^{\prime}_{q+1}}).

Since cqc^{\prime}_{\ell^{\prime}_{q}} and cq+1c^{\prime}_{\ell^{\prime}_{q+1}} are positive, there exists an ϵ>0\epsilon>0 such that ϵ<cq\epsilon<c^{\prime}_{\ell^{\prime}_{q}} and ϵ<cq+1\epsilon<c^{\prime}_{\ell^{\prime}_{q+1}}. It then follows that

cqzq+cq+1zq+1=2ϵzk+(cqϵ)zq+(cq+1ϵ)zq+1.c^{\prime}_{\ell^{\prime}_{q}}z^{\prime}_{\ell^{\prime}_{q}}+c^{\prime}_{\ell^{\prime}_{q+1}}z^{\prime}_{\ell^{\prime}_{q+1}}=2\epsilon z^{\prime}_{k^{\prime}}+(c^{\prime}_{\ell^{\prime}_{q}}-\epsilon)z^{\prime}_{\ell^{\prime}_{q}}+(c^{\prime}_{\ell^{\prime}_{q+1}}-\epsilon)z^{\prime}_{\ell^{\prime}_{q+1}}. (38)

Plugging in (36) the relation (38) shows that xx^{\prime} can be written as a convex combination of the zfz^{\prime}_{f^{\prime}}, for fFf^{\prime}\in F^{\prime}, with all positive coefficients, and thus xint𝒳(S)x^{\prime}\in\operatorname{int}\mathcal{X}(S^{\prime}).

Part 2: x𝒳(S)x𝒳(S)x^{\prime}\in\mathcal{X}(S^{\prime})\Rightarrow x\in\mathcal{X}(S) (xint𝒳(S)xint𝒳(S)x^{\prime}\in\operatorname{int}\mathcal{X}(S^{\prime})\Rightarrow x\in\operatorname{int}\mathcal{X}(S)). Because x𝒳(S)x^{\prime}\in\mathcal{X}(S^{\prime}) (resp. xint𝒳(S)x^{\prime}\in\operatorname{int}\mathcal{X}(S^{\prime})), we can write x=fFcfzfx^{\prime}=\sum_{f^{\prime}\in F^{\prime}}c^{\prime}_{f^{\prime}}z^{\prime}_{f^{\prime}}, with cf0c^{\prime}_{f^{\prime}}\geq 0 (resp. cf>0c^{\prime}_{f^{\prime}}>0), for all fFf^{\prime}\in F^{\prime}. We will use cfc^{\prime}_{f^{\prime}} to construct cfc_{f}, for fFf\in F, so that

x=fFcfzf.x=\sum_{f\in F}c_{f}z_{f}. (39)

To this end, we define cfc_{f} as follows:

  1. e)

    If ff is not incident to uqu_{q} in SS, then let cf:=cfc_{f}:=c^{\prime}_{f}.

  2. f)

    If f=(ui,uq)f=(u_{i},u_{q}) and uiuqu_{i}\neq u_{q}, then g:=(ui,uq)g^{\prime}:=(u_{i},u_{q}) and h:=(ui,uq+1)h^{\prime}:=(u_{i},u_{q+1}) are edges in SS^{\prime}, and let cf:=cg+chc_{f}:=c^{\prime}_{g^{\prime}}+c^{\prime}_{h^{\prime}}.

  3. g)

    If f=(uq,uq)f=(u_{q},u_{q}), then k:=(uq,uq+1)k^{\prime}:=(u_{q},u_{q+1}), q:=(uq,uq)\ell^{\prime}_{q}:=(u_{q},u_{q}), and q+1:=(uq+1,uq+1)\ell^{\prime}_{q+1}:=(u_{q+1},u_{q+1}) are edges in SS^{\prime}, and let cf:=ck+cq+cq+1c_{f}:=c^{\prime}_{k^{\prime}}+c^{\prime}_{\ell^{\prime}_{q}}+c^{\prime}_{\ell^{\prime}_{q+1}}.

Note that all the coefficients cfc_{f}, for fFf\in F, defined above are nonnegative. Further, if all the cfc^{\prime}_{f^{\prime}} are positive, i.e., xint𝒳(S)x^{\prime}\in\operatorname{int}\mathcal{X}(S^{\prime}), then the cfc_{f} are positive as well, which implies xint𝒳(S)x\in\operatorname{int}\mathcal{X}(S) provided that (39) holds.

We now show that the coefficients given above are such that (39) indeed holds. We do so by checking that (39) holds for each entry.

For the iith entry, with 1i<q1\leq i<q, the left hand side of (39) is xi=(σiσi1)x_{i}=(\sigma_{i}-\sigma_{i-1}). For the right-hand side, if (ui,uq)(u_{i},u_{q}) is an edge in SS, then gg^{\prime} and hh^{\prime}, as defined item (f), are two edges in SS^{\prime} and, consequently, Fui=Fui{h}F^{\prime}_{u_{i}}=F_{u_{i}}\cup\{h^{\prime}\}. Note that zf,i=zf,iz_{f,i}=z^{\prime}_{f,i} for all fFuif\in F_{u_{i}} and

zg,i=zg,i=zh,i=12.z_{g^{\prime},i}=z^{\prime}_{g^{\prime},i}=z^{\prime}_{h^{\prime},i}=\frac{1}{2}.

Thus, by items (e) and (f), we have that

fFcfzf,i\displaystyle\sum_{f\in F}c_{f}z_{f,i} =cgzg,i+fFui{g}cfzf,i\displaystyle=c_{g^{\prime}}z_{g^{\prime},i}+\sum_{f\in F_{u_{i}}-\{g^{\prime}\}}c_{f}z_{f,i}
=cgzg,i+chzh,i+fFui{g,h}cfzf,i\displaystyle=c^{\prime}_{g^{\prime}}z^{\prime}_{g^{\prime},i}+c^{\prime}_{h^{\prime}}z^{\prime}_{h^{\prime},i}+\sum_{f^{\prime}\in F^{\prime}_{u_{i}}-\{g^{\prime},h^{\prime}\}}c^{\prime}_{f^{\prime}}z^{\prime}_{f^{\prime},i}
=fFuicfzf,i=xi=(σiσi1).\displaystyle=\sum_{f^{\prime}\in F^{\prime}_{u_{i}}}c^{\prime}_{f^{\prime}}z^{\prime}_{f^{\prime},i}=x^{\prime}_{i}=(\sigma_{i}-\sigma_{i-1}).

Finally, for the last entry, i.e., the qqth entry, the left hand side of (39) is xq=σσqx_{q}=\sigma_{*}-\sigma_{q}. For the right hand side of (39), we let q:=(uq,uq)\ell_{q}:=(u_{q},u_{q}) be the loop on uqu_{q} (if it exists in SS) and thus have that

fFuqcfzf,q=cqzq,q+fFuq{q}cfzf,q.\sum_{f\in F_{u_{q}}}c_{f}z_{f,q}=c_{\ell_{q}}z_{\ell_{q},q}+\sum_{f\in F_{u_{q}}-\{\ell_{q}\}}c_{f}z_{f,q}. (40)

Let kk^{\prime}, q\ell^{\prime}_{q}, and q+1\ell^{\prime}_{q+1} be the three edges in SS^{\prime} as defined in item (g). Note that

zq,q=zq,q=zq+1,q+1=2zk,q=2zk,q+1=1.z_{\ell_{q},q}=z^{\prime}_{\ell^{\prime}_{q},q}=z^{\prime}_{\ell^{\prime}_{q+1},q+1}=2z^{\prime}_{k^{\prime},q}=2z^{\prime}_{k^{\prime},q+1}=1.

For the first term of (40), using item (g) and the above relations, we obtain

cqzq,q=cqzq,q+cq+1zq+1,q+1+ckzk,q+ckzk,q+1.c_{\ell_{q}}z_{\ell_{q},q}=c^{\prime}_{\ell^{\prime}_{q}}z^{\prime}_{\ell^{\prime}_{q},q}+c^{\prime}_{\ell^{\prime}_{q+1}}z^{\prime}_{\ell^{\prime}_{q+1},q+1}+c^{\prime}_{k^{\prime}}z^{\prime}_{k^{\prime},q}+c^{\prime}_{k^{\prime}}z^{\prime}_{k^{\prime},q+1}. (41)

For each addend in the second term of (40), the edge g=(ui,uq)g=(u_{i},u_{q}) in SS, for some uiuqu_{i}\neq u_{q}, has two corresponding edges in SS^{\prime}, namely g=(ui,uq)g^{\prime}=(u_{i},u_{q}) and h=(ui,uq+1)h^{\prime}=(u_{i},u_{q+1}). Note that

zg,q=zg,q=zh,q+1=12.z_{g,q}=z^{\prime}_{g^{\prime},q}=z^{\prime}_{h^{\prime},q+1}=\frac{1}{2}.

Then, by item (f),

cgzg,q=cgzg,q+chzh,q+1.c_{g}z_{g,q}=c^{\prime}_{g^{\prime}}z^{\prime}_{g^{\prime},q}+c^{\prime}_{h^{\prime}}z^{\prime}_{h^{\prime},q+1}. (42)

Combining (41) and (42), we obtain that

fFuqcfzf,1\displaystyle\sum_{f\in F_{u_{q}}}c_{f}z_{f,1} =fFuqcfzf,q+fFuq+1cfzf,q+1\displaystyle=\sum_{f^{\prime}\in F^{\prime}_{u_{q}}}c^{\prime}_{f^{\prime}}z^{\prime}_{f^{\prime},q}+\sum_{f^{\prime}\in F^{\prime}_{u_{q+1}}}c^{\prime}_{f^{\prime}}z^{\prime}_{f^{\prime},q+1}
=xq+xq+1=(σq+1σq)+(σσq+1)\displaystyle=x^{\prime}_{q}+x^{\prime}_{q+1}=(\sigma_{q+1}-\sigma_{q})+(\sigma_{*}-\sigma_{q+1})
=σσq.\displaystyle=\sigma_{*}-\sigma_{q}.

This concludes the proof.  

Appendix B Proof of Lemma 7

Proof of item 1: From the definitions of τi(x)\tau^{\prime}_{i}(x), we have that

𝟏n(τ0(x)τ0(x))=n(τ1(x)τ1(x))𝟏.-{\bf 1}\leq n(\tau_{0}(x)-\tau^{\prime}_{0}(x))=n(\tau^{\prime}_{1}(x)-\tau_{1}(x))\leq{\bf 1}. (43)

If τ0,i(x)=0\tau_{0,i}(x)=0, then τ0,i(x)=0\tau^{\prime}_{0,i}(x)=0, where τ0,i(x)\tau_{0,i}(x) is the iith entry of the vector τ0(x)\tau_{0}(x). Otherwise, by the definition of the incidence matrix (3) and by (14) and (15), we have that τ0,i(x)α\tau_{0,i}(x)\geq\alpha. For the latter case, by (43) and the hypothesis on nn in the statement of Proposition 2, we have that

τ0,i(x)τ0,i(x)1n(1ϵ/8)α>0.\tau^{\prime}_{0,i}(x)\geq\tau_{0,i}(x)-\frac{1}{n}\geq(1-\epsilon/8)\alpha>0. (44)

It then follows that

suppτ0(x)=suppτ0(x).\operatorname{supp}\tau^{\prime}_{0}(x)=\operatorname{supp}\tau_{0}(x). (45)

Similarly, for τ1(x)\tau_{1}(x), using (3), (14), and (15), we have that τ1(x)12α𝟏\tau_{1}(x)\geq\frac{1}{2}\alpha{\bf 1}. Then, again, by (43) and the hypothesis on nn,

τ1(x)τ1(x)1n𝟏12(1ϵ/4)α𝟏>0,\tau^{\prime}_{1}(x)\geq\tau_{1}(x)-\frac{1}{n}{\bf 1}\geq\frac{1}{2}(1-\epsilon/4)\alpha{\bf 1}>0,

from which we conclude that

suppτ1(x)=suppτ1(x)={1,,q}.\operatorname{supp}\tau^{\prime}_{1}(x)=\operatorname{supp}\tau_{1}(x)=\{1,\ldots,q\}.

This concludes the proof of the first item.

Proof of item 2: By (44), (45), and Lemma 4, the ratio n0/nn^{\prime}_{0}/n is uniformly lower bounded by

n0n=τ0(x)1(1ϵ/8)α|suppτ0(x)|=(1ϵ/8)α|F0|.\frac{n^{\prime}_{0}}{n}=\|\tau^{\prime}_{0}(x)\|_{1}\geq(1-\epsilon/8)\alpha|\operatorname{supp}\tau^{\prime}_{0}(x)|=(1-\epsilon/8)\alpha|F_{0}|. (46)

To obtain a lower bound for n1/nn_{1}^{\prime}/n, we let

η:=τ1(x)τ1(x).\eta:=\tau_{1}(x)-\tau^{\prime}_{1}(x).

From (43) and the hypothesis on nn, we have that

η1qn18qαϵ18qα.\|\eta\|_{1}\leq\frac{q}{n}\leq\frac{1}{8}q\alpha\epsilon\leq\frac{1}{8}q\alpha.

It then follows that

n1nτ1(x)1η112qα18qα=38qα.\frac{n_{1}^{\prime}}{n}\geq\|\tau_{1}(x)\|_{1}-\|\eta\|_{1}\geq\frac{1}{2}q\alpha-\frac{1}{8}q\alpha=\frac{3}{8}q\alpha. (47)

This concludes the proof of the second item.

Proof of item 3: By (43) and the hypothesis on nn, η\eta as introduced above satisfies

η1/n<αϵ/8<ϵ.\|\eta\|_{\infty}\leq 1/n<\alpha\epsilon/8<\epsilon.

Because τ¯1(x)=τ1(x)+η¯\bar{\tau}_{1}^{\prime}(x)=\overline{\tau_{1}(x)+\eta} and η<ϵ\|\eta\|_{\infty}<\epsilon, we have that τ¯1(x)𝒱\bar{\tau}^{\prime}_{1}(x)\in{\cal V} by Lemma 6.  

Appendix C Proof of Lemma 8

The proof is carried out by induction on nn^{\prime}. If n=0n^{\prime}=0, then GG^{\prime} is the empty graph and there is nothing to prove. For the inductive step, we set n>0n^{\prime}>0 and assume that the lemma holds for all n′′<nn^{\prime\prime}<n^{\prime}, and prove it for nn^{\prime}.

To proceed, we use AA^{\prime} to turn S\vec{S} into a weighted digraph on qq nodes: we assign to edge uiuju_{i}u_{j} the weight aija^{\prime}_{ij}. Then, S\vec{S} is a balanced graph, i.e.,

ujN(ui)aij=ujN+(ui)aji,i=1,,q,\sum_{u_{j}\in N_{-}(u_{i})}a^{\prime}_{ij}=\sum_{u_{j}\in N_{+}(u_{i})}a^{\prime}_{ji},\quad\forall i=1,\ldots,q, (48)

where we recall N(ui)N_{-}(u_{i}) and N+(ui)N_{+}(u_{i}) are the sets of out-neighbors and in-neighbors of uiu_{i}, respectively, in S\vec{S}.

Let S\vec{S}^{\prime} be the subgraph of S\vec{S} induced by the nodes incident to the edges with nonzero weights. Then, S\vec{S}^{\prime} has at least one cycle. To see this, note that if S\vec{S}^{\prime} is acyclic, then by relabeling the nodes, the matrix AA^{\prime} is upper-triangular and, from the hypothesis, diagA=0\operatorname{diag}A^{\prime}=0. It follows that the only nonnegative solution {aij}\{a^{\prime}_{ij}\} to (48) is that all the aija^{\prime}_{ij} are zero, which contradicts the fact that AA^{\prime} is nonzero.

Since S\vec{S}^{\prime} is a subgraph of S\vec{S}, any cycle of S\vec{S}^{\prime} is also a cycle of S\vec{S}; denote such cycle by DS:=ui1uikui1D_{S}:=u_{i_{1}}\ldots u_{i_{k}}u_{i_{1}}. By construction, the weights of the edges in the cycle are positive. It thus follows from A𝟏=xA^{\prime}{\bf 1}=x^{\prime} that the entries xijx^{\prime}_{i_{j}}, for j=1,,kj=1,\ldots,k, are positive; together with the fact that G=M(nx,S)G^{\prime}=M(n^{\prime}x^{\prime},S), it implies that the sets π1(uij)G\pi^{-1}(u_{i_{j}})\in G^{\prime}, for j=1,,kj=1,\ldots,k, are non-empty. We next pick a node vjv_{j} from each π1(uij)\pi^{-1}(u_{i_{j}}). Since the nodes ui1,,uiku_{i_{1}},\ldots,u_{i_{k}} are pairwise distinct, so are the nodes v1,,vkv_{1},\ldots,v_{k}. Also, since GG^{\prime} is a complete SS-multipartite graph, DG:=v1vkv1D_{G}:=v_{1}\cdots v_{k}v_{1} is a cycle in G\vec{G}^{\prime}. Moreover, by construction, π(DG)=DS\pi(D_{G})=D_{S} and, hence, DGD_{G} is simple.

We let G′′G^{\prime\prime} be the graph obtained by removing from GG^{\prime} the kk nodes v1,,vkv_{1},\ldots,v_{k}, and the edges incident to them. Then, G′′G^{\prime\prime} is a complete SS-multipartite graph on n′′:=nkn^{\prime\prime}:=n^{\prime}-k nodes. Define

x′′:=1n′′(nxj=1keij),x^{\prime\prime}:=\frac{1}{n^{\prime\prime}}(n^{\prime}x^{\prime}-\sum^{k}_{j=1}e_{i_{j}}),

where {e1,,eq}\{e_{1},\ldots,e_{q}\} is the canonical basis of q\mathbb{R}^{q}. Note that x′′0x^{\prime\prime}\geq 0; indeed, nxn^{\prime}x^{\prime} is integer valued and xijx^{\prime}_{i_{j}}, for j=1,,kj=1,\ldots,k, are positive. We can then write G′′=M(n′′x′′,S)G^{\prime\prime}=M(n^{\prime\prime}x^{\prime\prime},S). Correspondingly, we define a q×qq\times q matrix A′′A^{\prime\prime} as follows:

A′′:=1n′′(nAj=1k1eijeij+1eikei1).A^{\prime\prime}:=\frac{1}{n^{\prime\prime}}(n^{\prime}A^{\prime}-\sum^{k-1}_{j=1}e_{i_{j}}e_{i_{j+1}}^{\top}-e_{i_{k}}e_{i_{1}}^{\top}).

In words, to obtain n′′A′′n^{\prime\prime}A^{\prime\prime}, we decrease the ijijth entry of nAn^{\prime}A^{\prime}, which is a positive integer, by one when the edge uiuju_{i}u_{j} is used in the cycle DSD_{S} and keep the other entries unchanged.

By construction, we have that A′′𝒜(S)A^{\prime\prime}\in\mathcal{A}(S), with A′′𝟏=x′′{A^{\prime\prime}}{\bf 1}=x^{\prime\prime}, and n′′A′′n^{\prime\prime}A^{\prime\prime} is integer-valued. Because n′′<nn^{\prime\prime}<n^{\prime}, we can appeal to the induction hypothesis and exhibit a Hamiltonian decomposition H′′H^{\prime\prime} of G′′\vec{G}^{\prime\prime} such that ρ(H′′)=A′′\rho(H^{\prime\prime})=A^{\prime\prime} and every cycle in H′′H^{\prime\prime} is simple. It is clear that adding the simple cycle DGD_{G} to H′′H^{\prime\prime} yields a Hamiltonian decomposition HH^{\prime} of G\vec{G}^{\prime} with desired properties. This completes the proof.  

Appendix D Proof of Lemma 9

1. Proof that B(n1,n2,p)B(n_{1},n_{2},p) has a left-perfect matching a.s.: The proof of this part relies on the following statement, which is a consequence of a stronger result of Erdős and Rényi [15]: For p(0,1)p\in(0,1) a constant, the random bipartite graph B(m,m,p)B(m,m,p) contains a perfect matching a.s.. Now, without loss of generality, we assume that n1n2n_{1}\leq n_{2} and let B(n1,n1,p)B(n_{1},n_{1},p) be a subgraph of B(n1,n2,p)B(n_{1},n_{2},p). Since n2/n1n_{2}/n_{1} is bounded above by a constant κ\kappa, nn\to\infty implies that n1n_{1}\to\infty. Since B(n1,n1,p)B(n_{1},n_{1},p) has a (left-)perfect matching a.s., so does B(n1,n2,p)B(n_{1},n_{2},p).

2. Proof that B(n1,n2,p)B(n_{1},n_{2},p) is connected a.s.: It is well known (see, e.g., [16, Exercise 4.3.7]) that B(m,m,p)B(m,m,p) is connected a.s.. We now extend the result to the general case where n1n_{1} is not necessarily equal to n2n_{2}. Again, we can assume without loss of generality that n1n2n_{1}\leq n_{2}. Let VL={α1,,αn1}V_{L}=\{\alpha_{1},\ldots,\alpha_{n_{1}}\} and VR={β1,,βn2}V_{R}=\{\beta_{1},\ldots,\beta_{n_{2}}\} be the left- and right-node sets of BB. Because n2/n1κn_{2}/n_{1}\leq\kappa, we can choose κ\kappa subsets VR,iVRV_{R,i}\subseteq V_{R}, so that |VR,i|=n1|V_{R,i}|=n_{1} and i=1κVR,i=VR\cup^{\kappa}_{i=1}V_{R,i}=V_{R}.

Denote by i\mathcal{E}_{i} the event that the subgraph BiB_{i} of B:=B(n1,n2,p)B:=B(n_{1},n_{2},p) induced by VLV_{L} and VR,iV_{R,i} is disconnected and by \mathcal{E} the event that BB is disconnected. Note that if every BiB_{i} is connected, then so is BB. Conversely, we have that i=1κi\mathcal{E}\subseteq\cup_{i=1}^{\kappa}\mathcal{E}_{i}.

Note that Bi=(n1,n1,p)B_{i}=(n_{1},n_{1},p) and, as argued above, n1n_{1}\to\infty as nn\to\infty. Since BiB_{i} is connected a.s., limn(i)=0\lim_{n\to\infty}\mathbb{P}(\mathcal{E}_{i})=0 and, hence,

limn()limni=1κ(i)=0.\lim_{n\to\infty}\mathbb{P}(\mathcal{E})\leq\lim_{n\to\infty}\sum^{\kappa}_{i=1}\mathbb{P}(\mathcal{E}_{i})=0.

This completes the proof.