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

Support of Closed Walks and Second Eigenvalue Multiplicity of the Normalized Adjacency Matrix

Theo McKenzie111[email protected]. Supported by NSF Grant DGE-1752814.
UC Berkeley
   Peter M. R. Rasmussen222[email protected]. Supported in part by grant 16582, Basic Algorithms Research Copenhagen (BARC), from the VILLUM Foundation.
University of Copenhagen
   Nikhil Srivastava333[email protected]. Supported by NSF Grants CCF-1553751 and CCF-2009011.
UC Berkeley
Abstract

We show that the multiplicity of the second normalized adjacency matrix eigenvalue of any connected graph of maximum degree Δ\Delta is bounded by O(nΔ7/5/log1/5o(1)n)O(n\Delta^{7/5}/\log^{1/5-o(1)}n) for any Δ\Delta, and by O(nlog1/2d/log1/4o(1)n)O(n\log^{1/2}d/\log^{1/4-o(1)}n) for simple dd-regular graphs when dlog1/4nd\geq\log^{1/4}n. In fact, the same bounds hold for the number of eigenvalues in any interval of width λ2/logΔ1o(1)n\lambda_{2}/\log_{\Delta}^{1-o(1)}n containing the second eigenvalue λ2\lambda_{2}. The main ingredient in the proof is a polynomial (in kk) lower bound on the typical support of a closed random walk of length 2k2k in any connected graph, which in turn relies on new lower bounds for the entries of the Perron eigenvector of submatrices of the normalized adjacency matrix.

1 Introduction

The eigenvalues of matrices associated with graphs play an important role in many areas of mathematics and computer science, so general phenomena concerning them are of broad interest. In their recent beautiful work on the equiangular lines problem, Jiang, Tidor, Yao, Zhang, and Zhao [JTY+19] proved the following novel result constraining the distribution of the adjacency eigenvalues of all connected graphs of sufficiently low degree.

Theorem 1.1.

If GG is a connected graph of maximum degree Δ\Delta on nn vertices, then the multiplicity of the second largest eigenvalue of its adjacency matrix AGA_{G} is bounded by O(nlogΔ/loglog(n)).O(n\log\Delta/\log\log(n)).

For their application to equiangular lines, [JTY+19] only needed to show that the multiplicity of the second eigenvalue is o(n)o(n), but they asked whether the O(n/loglog(n))O(n/\log\log(n)) dependence in Theorem 1.1 could be improved, noting a huge gap between this and the best known lower bound of Ω(n1/3)\Omega(n^{1/3}) achieved by certain Cayley graphs of PSL(2,p)\mathrm{PSL}(2,p) (see [JTY+19, Section 4]). Apart from Theorem 1.1, there are as far as we are aware no known sublinear upper bounds on the second eigenvalue multiplicity for any general class of graphs, even if the question is restricted to Cayley graphs (unless one imposes a restriction on the spectral gap; see Section 1.2 for a discussion).

Meanwhile, in the theoretical computer science community, the largest eigenvalues of the normalized adjacency matrix A~G:=DG1/2AGDG1/2\tilde{A}_{G}:=D^{-1/2}_{G}A_{G}D^{-1/2}_{G} (for DGD_{G} the diagonal matrix of degrees) have received much attention over the past decade due to their relation with graph partitioning problems and the unique games conjecture (see e.g. [Kol11, BRS11, LRTV12, OGT13, LOGT14, ABS15, BGH+15, LOG18]); in particular, many algorithmic tasks become easier on graphs with few large normalized adjacency eigenvalues. Thus, it is of interest to know how many of these eigenvalues there can be in the worst case.

In this work, we prove significantly stronger upper bounds than Theorem 1.1 on the second eigenvalue multiplicity for the normalized adjacency matrix. Graphs are undirected and allowed to have multiedges and self-loops, unless specified to be simple. Order the eigenvalues of A~G\tilde{A}_{G} as λ1(A~G)λ2(A~G)λn(A~G)\lambda_{1}(\tilde{A}_{G})\geq\lambda_{2}(\tilde{A}_{G})\geq\ldots\geq\lambda_{n}(\tilde{A}_{G}), and let mG(I)m_{G}(I) denote the number of eigenvalues of A~G\tilde{A}_{G} in an interval II.

Theorem 1.2.

If GG is a connected graph of maximum degree Δ\Delta on nn vertices with λ2(A~G)=λ2\lambda_{2}(\tilde{A}_{G})=\lambda_{2}, then444All asymptotics are as nn\rightarrow\infty and the notation O~()\widetilde{O}(\cdot) suppresses polyloglog(n)\mathrm{polyloglog}(n) terms.

mG([(1loglogΔnlogΔn)λ2,λ2])=O~(nΔ7/5log1/5n).m_{G}\left([(1-\frac{\log\log_{\Delta}n}{\log_{\Delta}n})\lambda_{2},\lambda_{2}]\right)=\widetilde{O}\left(n\cdot\frac{\Delta^{7/5}}{\log^{1/5}n}\right). (1)

Because of the relationship A~G=1dAG\tilde{A}_{G}=\frac{1}{d}A_{G} when GG is regular, (1) gives a substantial improvement on Theorem 1.1 in the regular case (in the non-regular case, the results are incomparable as they concern different matrices). In addition to the stronger O(n/polylog(n))O(n/\textnormal{polylog}(n)) bound, a notable difference between our result and Theorem 1.1 is that we control the number of eigenvalues in a small interval containing λ2\lambda_{2}. Though we do not know whether the exponents in (1) are sharp, we show in Section 5.1 that constant degree bipartite Ramanujan graphs have at least Ω(n/log3/2n)\Omega(n/\log^{3/2}n) eigenvalues in the interval appearing in (1), indicating that O(n/polylog(n))O(n/\textnormal{polylog}(n)) is the correct regime for the maximum number of eigenvalues in such an interval when Δ\Delta is constant.

Theorem 1.2 is nontrivial for all Δ=o~(log1/7n)\Delta=\widetilde{o}(\log^{1/7}n); as remarked in [JTY+19], Paley graphs have degree Ω(n)\Omega(n) and second eigenvalue multiplicity Ω(n)\Omega(n), so some bound on the degree is required to obtain sublinear multiplicity. In Section 1.1, we present a variant of Theorem 1.2 (advertised in the abstract) which yields nontrivial bounds in the special case of simple dd-regular graphs with degrees as large as d=exp(log1/2δn)d=\exp(\log^{1/2-\delta}n), which is considerably larger than the regime d=O(polylog(n))d=O(\textnormal{polylog}(n)) handled by [JTY+19].

The main new ingredient in the proof of Theorem 1.2 is a polynomial lower bound on the support of (i.e., number of distinct vertices traversed by) a simple random walk of fixed length conditioned to return to its starting point. The bound holds for any connected graph and any starting vertex and may be of independent interest.

Theorem 1.3.

Suppose GG is connected and of maximum degree Δ\Delta on nn vertices and xx is any vertex in GG. Let γx2k=(x=X0,X1,,X2k)\gamma_{x}^{2k}=(x=X_{0},X_{1},\ldots,X_{2k}) denote a random walk of length 2k<n2k<n sampled according to the simple random walk on GG starting at xx. Then

(support(γx2k)s|X2k=X0)exp(k65Δ7s4)fors14(kΔ7logΔ)1/5.\mathbb{P}(\textnormal{support}(\gamma^{2k}_{x})\leq s|X_{2k}=X_{0})\leq\exp\left(-\frac{k}{65\Delta^{7}s^{4}}\right)\qquad\textrm{for}\quad s\leq\frac{1}{4}\left(\frac{k}{\Delta^{7}\log\Delta}\right)^{1/5}. (2)

In particular, this means that for constant Δ\Delta, the typical support of a closed random walk of length 2k2k is least Ω(k1/5)\Omega(k^{1/5}). It may be tempting to compare Theorem 1.3 with the familiar fact that a random closed walk of length 2k2k on \mathbb{Z} (or in continuous time, a standard Brownian bridge run for time 2k)2k) attains a maximum distance of Ω(k)\Omega(\sqrt{k}) from its origin. However, as seen in Figure 1, there are regular graphs for which a closed walk of length 2k2k from a particular vertex xx travels a maximum distance of only polylog(k)\textnormal{polylog}(k) with high probability. Theorem 1.3 reveals that nonetheless the number of distinct vertices traversed is always typically poly(k)\mathrm{poly}(k). We do not know if the specific exponent of k1/5k^{1/5} supplied by Theorem 1.3 is sharp, but considering a cycle graph shows that it is not possible to do better than k1/2k^{1/2}.

Refer to caption
Figure 1: For a regular graph composed of a near-clique attached to an infinite tree, a closed walk of length 2k2k starting from within the near-clique does not typically go deeper than O(logk)O(\log k) down the tree. However, the support of such a closed walk is typically kΘ(1)k^{\Theta(1)}. See Section B for a more detailed discussion.

Given Theorem 1.3, our proof of Theorem 1.2 follows the strategy of [JTY+19]: since most closed walks in GG have large support, the number of such walks may be drastically reduced by deleting a small number of vertices from GG. By a moment calculation relating the spectrum to self return probabilities and a Cauchy interlacing argument, this implies an upper bound on the multiplicity of λ2(A~G)\lambda_{2}(\tilde{A}_{G}). The crucial difference is that we are able to delete only n/polylog(n)n/\textnormal{polylog}(n) vertices whereas they delete n/polyloglog(n)n/\mathrm{poly}\log\log(n).

The key ingredient in our proof of Theorem 1.3 is a result regarding the Perron eigenvector (i.e., the unique, strictly positive eigenvector with eigenvalue λ1\lambda_{1}) of a submatrix of A~\tilde{A}.

Theorem 1.4.

For any graph G=(V,E)G=(V,E) of maximum degree Δ\Delta, take any set of vertices SVS\subsetneq V such that the induced subgraph on SS is connected, and let ψS\psi_{S} be the 2\ell_{2}-normalized Perron vector of A~S\tilde{A}_{S}, the principal submatrix of A~\tilde{A} corresponding to vertices in SS. Then there is a vertex uSu\in S which is adjacent to VSV\setminus S such that

ψS(u)1/(Δ5/2λ1(A~S)|S|5/2).\psi_{S}(u)\geq 1/(\Delta^{5/2}\lambda_{1}(\tilde{A}_{S})|S|^{5/2}). (3)

When we restrict this result to GG being a dd-regular graph and pass to the adjacency matrix, we achieve a result about the unnormalized adjacency matrix of irregular graphs that may be of independent interest.

Corollary 1.5.

Let H=(V,E)H=(V,E) be an irregular connected graph of maximum degree Δ\Delta with at least two vertices, and let ϕH\phi_{H} be the 2\ell_{2}-normalized Perron vector of AHA_{H}. Then there is a vertex uVu\in V with degree strictly less than Δ\Delta satisfying

ϕH(u)1/(Δ2λ1(AH)|V|5/2).\phi_{H}(u)\geq 1/(\Delta^{2}\lambda_{1}(A_{H})|V|^{5/2}). (4)

Corollary 1.5 may be compared with existing results in spectral graph theory on the “principal ratio” between the largest and smallest entries of the Perron vector of a connected graph. The known worst case lower bounds on this ratio are necessarily exponential in the diameter of the graph [CG07, TT15], and it is known that the Perron entry of the highest degree vertex cannot be very small (see e.g. [Ste14, Ch. 2]). Corollary 1.5 articulates that there is always at least one vertex of non-maximal degree for which the ratio is only polynomial in the number of vertices.

The proof of Theorem 1.4 is based on an analysis of hitting times in the simple random walk on GG via electrical flows, and appears in Section 2. Combined with a perturbation-theoretic argument, it enables us to show that any small connected induced subgraph SS of GG can be extended to a slightly larger induced subgraph with significantly larger Perron value λ1(A~S)\lambda_{1}(\tilde{A}_{S}). With some further combinatorial arguments, this implies that closed walks cannot concentrate on small sets, yielding Theorem 1.3 in Section 3, which is finally used to deduce Theorem 1.2 in Section 4.

We show in Section 5.2 via an explicit example (Figure 2) that the exponent of 5/25/2 appearing in Corollary 1.5 is sharp up to polylogarithmic factors. We conclude with a discussion of open problems in Section 6.

Refer to caption
Figure 2: An example of a graph where all vertices uu that are not of maximum degree have ψ(u)=O~(n5/2)\psi(u)=\tilde{O}(n^{-5/2}). The circled sets X0X_{0}, X1X_{1} and X2X_{2} will be used in the analysis of the graph in Section 5.2.
Remark 1.6 (Higher Eigenvalues).

An update of the preprint of [JTY+19] generalizes Theorem 1.1 to the multiplicity of the jjth eigenvalue. Our results can also be generalized in this manner by some nominal changes to the arguments in Section 4, but for simplicity we focus on λ2\lambda_{2} in this paper.

1.1 Higher degree regular graphs

If G=(V,E)G=(V,E) is a simple, dd-regular graph, and SVS\subsetneq V such that |S|d|S|\leq d, then necessarily all vertices of SS are adjacent to vertices in VSV\setminus S. Therefore we can improve the bound from Theorem 1.4 by assuming the vertex on the boundary is the maximizer of the Perron vector, which has value ψS(u)1/|S|\psi_{S}(u)\geq 1/\sqrt{|S|}. This leads to the following variants of our main results for simple, regular graphs of sufficiently high degree.

Theorem 1.7.

GG is simple, dd-regular, and connected with λ2=λ2(AG)\lambda_{2}=\lambda_{2}(A_{G}), then

mG([(1loglogdnlogdn)λ2,λ2])={O~(nd) when d=o(log1/4n)O~(nlog1/2dlog1/4n) when d=Ω(log1/4n).m_{G}\left([(1-\frac{\log\log_{d}n}{\log_{d}n})\lambda_{2},\lambda_{2}]\right)=\begin{cases}\widetilde{O}\left(\frac{n}{d}\right)&\textrm{ when }d=o(\log^{1/4}n)\\ \widetilde{O}\left(\frac{n\log^{1/2}d}{\log^{1/4}n}\right)&\textrm{ when }d=\Omega(\log^{1/4}n).\end{cases} (5)

The above theorem is based on the following corresponding result for closed walks.

Theorem 1.8.

If GG is simple, dd-regular, and connected on nn vertices and γ\gamma is a random closed walk of length 2k<n2k<n started at any vertex in GG, then:

Pr(support(γ)s)exp(k100s3)forsmin{18(klogd)1/4,d2}.\Pr(\mathrm{support}(\gamma)\leq s)\leq\exp\left(-\frac{k}{100s^{3}}\right)\qquad\textrm{for}\quad s\leq\min\left\{\frac{1}{8}\left(\frac{k}{\log d}\right)^{1/4},\frac{d}{2}\right\}. (6)

The proofs of both theorems appear in Appendix A

1.2 Related work

Eigenvalue Multiplicity.

Despite the straightforward nature of the question, relatively little is known about eigenvalue multiplicity of general graphs. As discussed in [JTY+19], if one assumes that GG is a bounded degree expander graph, then the bound of Theorem 1.1 can be improved to O(n/logn)O(n/\log n). On the other hand, if GG is assumed to be a Cayley graph of bounded doubling constant KK (indicating non-expansion), then [LM08] show that the multiplicity of the second eigenvalue is at most exp(log2K)\exp(\log^{2}K). In the context of Cayley graphs, one interesting new implication of Theorem 1.7 is that all Cayley graphs of degree O(exp(log1/2δn))O(\exp(\log^{1/2-\delta}n)) have second eigenvalue multiplicity O(n/logδ/2n)O(n/\log^{\delta/2}n).

Distance regular graphs of diameter DD have exactly D+1D+1 distinct eigenvalues (see [God93] 11.4.1 for a proof). However, besides the top eigenvalue (which must have multiplicity 1), generic bounds on the multiplicity of the other eigenvalues are not known. As expanding graphs have diameter Θ(logdn)\Theta(\log_{d}n), the average multiplicity of eigenvalues besides λ1\lambda_{1} for expanding distance regular graphs is Θ(n/logdn)\Theta(n/\log_{d}n). It is tempting to see this as a hint that the multiplicity of the second eigenvalue could be Ω(n/logdn)\Omega(n/\log_{d}n).

Sublinear multiplicity does not necessarily hold for eigenvalues in the interior of the spectrum even assuming bounded degree. In particular, Rowlinson has constructed connected dd-regular graphs with an eigenvalue of multiplicity at least n(d2)/(d+2)n(d-2)/(d+2) [Row19] for constant dd.

Higher Order Cheeger Inequalities.

The results of [LRTV12, LOGT14] imply that if a dd-regular graph GG has a second eigenvalue multiplicity of mm, then its vertices can be partitioned into Ω(m)\Omega(m) disjoint sets each having edge expansion O(d(1λ2)logm)O(\sqrt{d(1-\lambda_{2})\log m}). Combining this with the observation that a set cannot have expansion less than the reciprocal of its size shows that m=O(n/polylog(n))m=O(n/\textnormal{polylog}(n)) whenever 1λ2(A~G)1/logcn1-\lambda_{2}(\tilde{A}_{G})\leq 1/\log^{c}n for any c>1c>1, i.e., the graph is sufficiently non-expanding. Our main theorem may be interpreted as saying that this phenomenon persists for all graphs.

Support of Walks.

There are as far as we are aware no known lower bounds for the support of a random closed walk of fixed length in a general graph (or even Cayley graph). It is relatively easy to derive such bounds for bounded degree graphs if the length of the walk is sufficiently larger than the mixing time of the simple random walk on the graph; the key feature of Theorem 1.3, which is needed for our application, is that the length of the walk can be taken to be much smaller.

The support of open walks (namely removing the condition that the walk ends at the starting point) is better understood. There are Chernoff-type bounds on the size of the support of a random walk based on the spectral gap [Gil98, Kah97]. Such bounds and their variants are an important tool in derandomization.

Entries of the Perron Vector.

There is a large literature concerning the magnitude of the entries of the Perron eigenvector of a graph — see [Ste14, Chapter 2] for a detailed discussion of results up to 2014. Rowlinson showed sufficient conditions on the Perron eigenvector for which changing the neighborhood of a vertex increases the spectral radius [Row90]. Cvetković, Rowlinson, and Simić give a condition which, if satisfied, means a given edge swap increases the spectral radius [CRS93]. Cioabă showed that for a graph of maximum degree Δ\Delta and diameter DD, Δλ1>1/nD\Delta-\lambda_{1}>1/nD [Cio07]. Cioabă, van Dam, Koolen, and Lee then showed that λ1(n1)1/D\lambda_{1}\geq(n-1)^{1/D} [CVDKL10]. The results of [VMSK+11] prove a lemma similar to Lemma 3.2, giving upper and lower bounds on the change in spectral radius from the deletion of edges. However, their result does not quite imply Lemma 3.2, and we prove a slightly different statement.

1.3 Notation

All logarithms are base ee unless noted otherwise.

Electrical Flows.

We use ReffH(,)\mathrm{Reff}_{H}(\cdot,\cdot) to denote the effective resistance between two vertices in HH, viewing each edge of the graph as a unit resistor. See e.g. [DS84] or [Bol13, Chapter IX] for an introduction to electrical flows and random walks on graphs.

Graphs.

For a matrix MM, we use MSM_{S} to denote the principal submatrix of MM corresponding to the indices in SS. Consider a graph G=(V,E)G=(V,E) and a subset HVH\subset V. Let P:=AD1P:=AD^{-1} be the transition matrix of the simple random walk matrix on GG, where AA is the adjacency matrix and DD is the diagonal matrix of degrees. We will also use the normalized adjacency matrix A~:=D1/2AD1/2\tilde{A}:=D^{-1/2}AD^{-1/2}. Note that PP and A~\tilde{A} are similar, and that A~\tilde{A} is symmetric. PSP_{S} and A~S\tilde{A}_{S} are submatrices of PP and A~\tilde{A}; they are not the transition matrices and normalized adjacency matrices of the induced subgraph on SS. Note PSP_{S} and A~S\tilde{A}_{S} are also similar.

Perron Eigenvector.

We use ψS\psi_{S} to denote the 2\ell_{2}-normalized eigenvector corresponding to λ1(A~S)\lambda_{1}(\tilde{A}_{S}), which is a simple eigenvalue if SS is connected. Note that for connected SS, ψS\psi_{S} is strictly positive by the Perron-Frobenius theorem.

A simple graph refers to a graph without multiedges or self-loops. We assume Δ2\Delta\geq 2 for all connected regular graphs, since otherwise the graph is just an edge, so logΔ>0\log\Delta>0.

2 Lower Bounds on the Perron Eigenvector

In this section we prove Theorem 1.4, which is a direct consequence of the following slightly more refined result. In a graph G=(V,E)G=(V,E), define the boundary of SS as the set of vertices in SS adjacent to V\SV\backslash S in GG.

Theorem 2.1 (Large Perron Entry).

Let G=(V,E)G=(V,E) be a connected graph of maximum degree Δ\Delta and SVS\subsetneq V such that the induced subgraph on SS is connected. Then there is a vertex uSu\in S on the boundary of SS such that

ψS(u)/ψS(t)1/(Δ5/2λ1(A~S)|S|2)\psi_{S}(u)/\psi_{S}(t)\geq 1/(\Delta^{5/2}\lambda_{1}(\tilde{A}_{S})|S|^{2}) (7)

where t=argmaxwSψS(w)t=\arg\max_{w\in S}\psi_{S}(w).

At a high level, the proof proceeds as follows. First, we show that there exists a vertex xSx\in S adjacent to the boundary of SS such that a random walk started at xx is somewhat likely to hit tt before it hits the boundary of SS. Second, we express the ratio of DS1/2ψS(x)D^{1/2}_{S}\psi_{S}(x) and DS1/2ψS(t)D^{1/2}_{S}\psi_{S}(t) as a limit as kk\to\infty of the ratio Yxk/Ytk\mathbb{P}{Y^{k}_{x}}/\mathbb{P}{Y_{t}^{k}}, where YvkY^{k}_{v} is the event that the simple random walk started at vv remains in SS for kk steps; we bound this ratio from below using the hitting time estimate from the first step. Third, by the eigenvector equation the ratio of the entries of an eigenvector at two neighboring vertices is bounded. Hence, xx is adjacent to some vertex uu on the boundary of SS satisfying the theorem.

Proof.

Write S=MBS=M\sqcup B, where BB is the boundary of SS and M=SBM=S\setminus B. If tBt\in B then we are done, so assume not. Let xG()\mathbb{P}^{G}_{x}(\cdot) denote the law of the simple random walk (SRW) (Xi)i=0(X_{i})_{i=0}^{\infty} on GG started at X0=xX_{0}=x, and for any subset TVT\subset V, let τT:={mini:XiT}\tau_{T}:=\{\min i:X_{i}\in T\} denote the hitting time of the SRW to that subset; if T={u}T=\{u\} is a singleton we will simply write τu\tau_{u}.

Step 1. We begin by showing that there is a vertex xMx\in M adjacent to BB for which the random walk started at xx is reasonably likely to hit tt before BB. To do so, we use the well-known connection between hitting probabilities in random walks and electrical flows. Define a new graph K=(V=VB{s},E)K=(V^{\prime}=V\setminus B\cup\{s\},E^{\prime}) by contracting all vertices in BB to a single vertex ss. Let f:V[0,1]f:V^{\prime}\rightarrow[0,1] be the vector of voltages in the electrical flow in KK with boundary conditions f(s)=0,f(t)=1f(s)=0,f(t)=1, regarding every edge as a unit resistor. By Ohm’s law, the current flow from ss to tt is equal to 1/ReffK(s,t)1/\mathrm{Reff}_{K}(s,t). We have the crude upper bound

ReffK(s,t)distanceK(s,t)|S|,\mathrm{Reff}_{K}(s,t)\leq\mathrm{distance}_{K}(s,t)\leq|S|,

since SS is connected, so the outflow of current from ss is at least 1/|S|1/|S|. By Kirchhoff’s current law, there must be a flow of at least 1/(|S|degK(s))1/(|S|\deg_{K}(s)) on at least one edge (s,x)E(s,x)\in E^{\prime}. By Ohm’s law again, for this particular xVx\in V^{\prime} we must have

f(x)1|S|degK(s)=1|S||GB|1Δ|S|2,f(x)\geq\frac{1}{|S|\deg_{K}(s)}=\frac{1}{|S||\partial_{G}B|}\geq\frac{1}{\Delta|S|^{2}}, (8)

where GB\partial_{G}B denotes the edge boundary of BB in GG. Appealing to e.g. [Bol13, Chapter IX, Theorem 8], this translates to the probabilistic bound

xG(τt<τB)=xK(τt<τs)=f(x)1Δ|S|2.\mathbb{P}_{x}^{G}(\tau_{t}<\tau_{B})=\mathbb{P}_{x}^{K}(\tau_{t}<\tau_{s})=f(x)\geq\frac{1}{\Delta|S|^{2}}. (9)

Finally, since f(s)=f(y)=0f(s)=f(y)=0 for every yVSy\in V\setminus S, we must in fact have xMx\in M.

Refer to caption
Figure 3: In Step 1 of the proof of Theorem 2.1, we lower bound the probability that a random walk started at a certain vertex xx adjacent to BB reaches tt before reaching BB. We do this by contracting BB to a vertex ss, then lower bounding the current from ss to tt, which establishes the existence of the desired xx. The left graph in the figure is GG and the right graph is the contracted graph KK, with the dotted lines indicating edges leaving the set of interest S=MBS=M\sqcup B.

Step 2. We now use (9) to show that ψS(x)\psi_{S}(x) is large. Because A~S=DS1/2PSDS1/2\tilde{A}_{S}=D^{-1/2}_{S}P_{S}D^{1/2}_{S}, the top eigenvector of PSP_{S} is DS1/2ψS/DS1/2ψSD^{1/2}_{S}\psi_{S}/\|D^{1/2}_{S}\psi_{S}\|. Let P:(P+I)/2P^{\prime}:(P+I)/2 denote the lazy random walk555This modification is only to ensure non-bipartiteness; if SS is not bipartite we may take the simple random walk on GG, and to ease notation let x():=xG()\mathbb{P}^{\prime}_{x}(\cdot):={\mathbb{P}^{\prime}_{x}}^{G}(\cdot) denote the law of the lazy random walk on GG started at xx. Note that the eigenvectors of PSP_{S}, as well as x(τt<τB)\mathbb{P}_{x}(\tau_{t}<\tau_{B}), do not change when passing to PSP^{\prime}_{S}.

For the lazy random walk, the Perron-Frobenius theorem implies that

(DS1/2ψS)(w)DS1/2ψS=limk𝟏STPSkew𝟏STPSk,\frac{(D^{1/2}_{S}\psi_{S})(w)}{\|D^{1/2}_{S}\psi_{S}\|}=\lim_{k\rightarrow\infty}\frac{\mathbf{1}_{S}^{T}{P^{\prime}_{S}}^{k}e_{w}}{\|\mathbf{1}_{S}^{T}{P^{\prime}_{S}}^{k}\|},

for every wSw\in S, where 𝟏SS\mathbf{1}_{S}\in\mathbb{R}^{S} is the all ones vector. We further have

𝟏STPSkew=w(τV\S>k),\mathbf{1}_{S}^{T}{P^{\prime}_{S}}^{k}e_{w}=\mathbb{P}^{\prime}_{w}(\tau_{V\backslash S}>k),

namely the probability a random walk of length kk starting at ww stays in SS.

We are interested in the ratio

(DS1/2ψS)(x)(DS1/2ψS)(t)=limkx(τV\S>k)t(τV\S>k).\frac{(D^{1/2}_{S}\psi_{S})(x)}{(D^{1/2}_{S}\psi_{S})(t)}=\lim_{k\rightarrow\infty}\frac{\mathbb{P}^{\prime}_{x}(\tau_{V\backslash S}>k)}{\mathbb{P}^{\prime}_{t}(\tau_{V\backslash S}>k)}. (10)

Fix an integer k>0k>0. The numerator of (10) is bounded as

x(τV\S>k)\displaystyle\mathbb{P}^{\prime}_{x}(\tau_{V\backslash S}>k) x(τV\S>k|τt<τB)x(τt<τB)\displaystyle\geq\mathbb{P}^{\prime}_{x}(\tau_{V\backslash S}>k|\tau_{t}<\tau_{B})\mathbb{P}^{\prime}_{x}(\tau_{t}<\tau_{B})
1Δ|S|2x(τV\S>k|τt<τB)by (9)\displaystyle\geq\frac{1}{\Delta|S|^{2}}\mathbb{P}^{\prime}_{x}(\tau_{V\backslash S}>k|\tau_{t}<\tau_{B})\qquad\textrm{by \eqref{eqn:hittingst}}
1Δ|S|2θ=0k1x(τV\S>k|τt=θ,τt<τB)x(τt=θ|τt<τB)\displaystyle\geq\frac{1}{\Delta|S|^{2}}\sum_{\theta=0}^{k-1}\mathbb{P}^{\prime}_{x}(\tau_{V\backslash S}>k|\tau_{t}=\theta,\tau_{t}<\tau_{B})\mathbb{P}^{\prime}_{x}(\tau_{t}=\theta|\tau_{t}<\tau_{B}) (11)
=1Δ|S|2θ=0k1t(τV\S>kθ)x(τt=θ|τt<τB)\displaystyle=\frac{1}{\Delta|S|^{2}}\sum_{\theta=0}^{k-1}\mathbb{P}^{\prime}_{t}(\tau_{V\backslash S}>{k}-\theta)\mathbb{P}^{\prime}_{x}(\tau_{t}=\theta|\tau_{t}<\tau_{B})
1Δ|S|2θ=0k1t(τV\S>k)x(τt=θ|τt<τB).\displaystyle\geq\frac{1}{\Delta|S|^{2}}\sum_{\theta=0}^{k-1}\mathbb{P}^{\prime}_{t}(\tau_{V\backslash S}>{k})\mathbb{P}^{\prime}_{x}(\tau_{t}=\theta|\tau_{t}<\tau_{B}). (12)

Observe that 𝔼xτB<\mathbb{E}^{\prime}_{x}\tau_{B}<\infty since GG is connected. Thus,

θ=0k1x(τt=θ|τt<τB)\displaystyle\sum_{\theta=0}^{k-1}\mathbb{P}^{\prime}_{x}(\tau_{t}=\theta|\tau_{t}<\tau_{B}) =1x(τtk|τt<τB)\displaystyle=1-\mathbb{P}^{\prime}_{x}(\tau_{t}\geq k|\tau_{t}<\tau_{B})
1x(τBk)x(τt<τB)\displaystyle\geq 1-\frac{\mathbb{P}^{\prime}_{x}(\tau_{B}\geq k)}{\mathbb{P}^{\prime}_{x}(\tau_{t}<\tau_{B})}
1𝔼xτBkΔ|S|2by Markov and (9).\displaystyle\geq 1-\frac{\mathbb{E}^{\prime}_{x}\tau_{B}}{k}\cdot\Delta|S|^{2}\quad\textrm{by Markov and \eqref{eqn:hittingst}}.

Combining this bound with (12), we have

x(τV\S>k)1Δ|S|2(1𝔼xτBkΔ|S|2)t(τV\S>k)\mathbb{P}^{\prime}_{x}(\tau_{V\backslash S}>k)\geq\frac{1}{\Delta|S|^{2}}\left(1-\frac{\mathbb{E}^{\prime}_{x}\tau_{B}}{k}\cdot\Delta|S|^{2}\right)\mathbb{P}^{\prime}_{t}(\tau_{V\backslash S}>k)

Taking the limit as kk\rightarrow\infty in (10) yields

(DS1/2ψS)(x)(DS1/2ψS)(t)1Δ|S|2.\frac{(D^{1/2}_{S}\psi_{S})(x)}{(D^{1/2}_{S}\psi_{S})(t)}\geq\frac{1}{\Delta|S|^{2}}.

Step 3. Since xx is adjacent to BB, we can choose a uBu\in B adjacent to xx. The eigenvector equation and nonnegativity of the Perron vector now imply Δλ1(AS)ψS(u)ψS(x)\Delta\lambda_{1}(A_{S})\psi_{S}(u)\geq\psi_{S}(x), whence

(DS1/2ψS)(u)1λ1(A~S)Δ2|S|2(DS1/2ψS)(t).(D^{1/2}_{S}\psi_{S})(u)\geq\frac{1}{\lambda_{1}(\tilde{A}_{S})\Delta^{2}|S|^{2}}(D^{1/2}_{S}\psi_{S})(t). (13)

Therefore, as DD is a diagonal matrix, and the entries of DD range from 11 to Δ\Delta, it must be the case that

ψS(u)1λ1(A~S)Δ5/2|S|2ψS(t).\psi_{S}(u)\geq\frac{1}{\lambda_{1}(\tilde{A}_{S})\Delta^{5/2}|S|^{2}}\psi_{S}(t).

Remark 2.2.

As the proof shows, the right-hand side of (7) may be replaced with 1/Δ3/2λ1(A~S)|GB|R1/\Delta^{3/2}\lambda_{1}(\tilde{A}_{S})|\partial_{G}B|R where BB is the boundary of SS in GG and RR is the maximum effective resistance between two vertices in SS.

Proof of Corollary 1.5.

Given an irregular graph HH, construct a Δ\Delta-regular graph GG containing HH as an induced subgraph (it is trivial to do this if we allow GG to be a multigraph). Repeating the above proof on GG with S=HS=H and observing that DS1/2D_{S}^{1/2} is a multiple of the identity since GG is regular, (13) yields the desired conclusion. ∎

3 Support of Closed Walks

In this section we prove Theorem 1.3, which is an immediate consequence of the following slightly stronger result. Let W2k,sW^{2k,s} denote the event a simple random walk of length 2k2k has support at most ss and ends at its starting point.

Theorem 3.1 (Implies Theorem 1.3).

If GG is connected and of maximum degree Δ\Delta on nn vertices, then for every vertex xGx\in G and k<n/2k<n/2,

x(W2k,s)exp(k65Δ7s4)x(W2k,2s)fors14(kΔ7logΔ)1/5.\mathbb{P}_{x}(W^{2k,s})\leq\exp\left(-\frac{k}{65\Delta^{7}s^{4}}\right)\mathbb{P}_{x}(W^{2k,2s})\qquad\textrm{for}\quad s\leq\frac{1}{4}\left(\frac{k}{\Delta^{7}\log\Delta}\right)^{1/5}. (14)

The proof requires a simple lemma lower bounding the increase in the Perron value of a subgraph upon adding a vertex in terms of the Perron vector.

Lemma 3.2 (Perturbation of λ1\lambda_{1}).

Take the normalized adjacency matrix A~:=D1/2AD1/2\tilde{A}:=D^{-1/2}AD^{-1/2} of a graph G=(V,E)G=(V,E) of maximum degree Δ\Delta. For any SVS\subsetneq V and vertex uSu\in S, the submatrix which includes the subset S=(S{v},E(S){(u,v)})S^{\prime}=(S\cup\{v\},E(S)\cup\{(u,v)\}), which adds a vertex vv and the edge (u,v)(u,v) to SS, satisfies

λ1(A~S)12(λ1(A~S)+λ1(A~S)2+Δ2ψS(u)2).\lambda_{1}(\tilde{A}_{S^{\prime}})\geq\frac{1}{2}\left(\lambda_{1}(\tilde{A}_{S})+\sqrt{\lambda_{1}(\tilde{A}_{S})^{2}+\Delta^{-2}\psi_{S}(u)^{2}}\right).
Proof.

The largest eigenvalue of A~\tilde{A} is at least the quadratic form associated with the unit vectors

gα(x)={1α2ψS(x)xVαx=vg_{\alpha}(x)=\left\{\begin{array}[]{cc}\sqrt{1-\alpha^{2}}\psi_{S}(x)&x\in V\\ \alpha&x=v\end{array}\right.

for 0α10\leq\alpha\leq 1. We have gαTA~gα=(1α2)λ1(A~S)+du1/2dv1/2α1α2ψS(u)g_{\alpha}^{T}\tilde{A}g_{\alpha}=(1-\alpha^{2})\lambda_{1}(\tilde{A}_{S})+d_{u}^{-1/2}d_{v}^{-1/2}\alpha\sqrt{1-\alpha^{2}}\psi_{S}(u), where dud_{u} is the degree of uu in GG. This quantity is maximized when

α=12λ1(A~S)2λ1(A~S)2+du1dv1ψS(u)2,\alpha=\sqrt{\frac{1}{2}-\frac{\lambda_{1}(\tilde{A}_{S})}{2\sqrt{\lambda_{1}(\tilde{A}_{S})^{2}+d_{u}^{-1}d_{v}^{-1}\psi_{S}(u)^{2}}}},

at which

gαTA~gα=12(λ1(A~S)+λ1(A~S)2+du1dv1ψS(u)2).g_{\alpha}^{T}\tilde{A}g_{\alpha}=\frac{1}{2}\left(\lambda_{1}(\tilde{A}_{S})+\sqrt{\lambda_{1}(\tilde{A}_{S})^{2}+d_{u}^{-1}d_{v}^{-1}\psi_{S}(u)^{2}}\right).

Combining Lemma 3.2 and Theorem 2.1 yields a bound on the increase of the top eigenvalue of the submatrix corresponding to an induced subgraph that may be achieved by adding vertices to it.

Lemma 3.3 (Support Extension).

For any connected graph G=(V,E)G=(V,E) of maximum degree Δ\Delta, consider its normalized adjacency matrix A~\tilde{A}. For any connected subset SVS\subsetneq V such that 2|S|=s<|V|/22\leq|S|=s<|V|/2, there is a connected subset TVT\subset V containing SS such that |T|=2s|T|=2s and

λ1(A~T)λ1(A~S)(1+5128Δ7s4).\lambda_{1}(\tilde{A}_{T})\geq\lambda_{1}(\tilde{A}_{S})\left(1+\frac{5}{128\Delta^{7}s^{4}}\right).
Proof.

Define λ1:=λ1(A~S)\lambda_{1}:=\lambda_{1}(\tilde{A}_{S}) and note that λ11/Δ\lambda_{1}\geq 1/\Delta since SS contains at least one edge. As ψS\psi_{S} is a normalized vector with ss entries, ψS(t)1/s\psi_{S}(t)\geq 1/\sqrt{s}. Therefore ψS(u)1/(Δ5/2λ1s5/2)\psi_{S}(u)\geq 1/(\Delta^{5/2}\lambda_{1}s^{5/2}). Take vv to be any vertex in VSV\setminus S that neighbors uu in GG. By Lemma 3.2,

λ1(A~S{v})\displaystyle\lambda_{1}(\tilde{A}_{S\cup\{v\}}) 12(λ1+λ12+Δ2ψS(u)2)\displaystyle\geq\frac{1}{2}\left(\lambda_{1}+\sqrt{\lambda_{1}^{2}+\Delta^{-2}\psi_{S}(u)^{2}}\right)
λ1+ψS(u)24λ1Δ2ψS(u)416λ13Δ4\displaystyle\geq\lambda_{1}+\frac{\psi_{S}(u)^{2}}{4\lambda_{1}\Delta^{2}}-\frac{\psi_{S}(u)^{4}}{16\lambda_{1}^{3}\Delta^{4}}
λ1+16λ13Δ7s5as ψS(u)2/λ12Δ2\displaystyle\geq\lambda_{1}+\frac{1}{6\lambda_{1}^{3}\Delta^{7}s^{5}}\qquad\textrm{as $\psi_{S}(u)^{2}/\lambda_{1}^{2}\leq\Delta^{2}$}
λ1+16Δ7s5since λ11.\displaystyle\geq\lambda_{1}+\frac{1}{6\Delta^{7}s^{5}}\qquad\textrm{since $\lambda_{1}\leq 1$.} (15)

Assuming that s<|V|/2s<|V|/2, we can iterate this process ss times, adding the vertices {v1,vs}\{v_{1},\ldots v_{s}\}. At each step we add the vertex viv_{i} and increase the Perron eigenvalue of A~S{v1,,vi1}\tilde{A}_{S\cup\{v_{1},\ldots,v_{i-1}\}} by at least 1/(6Δ7(s+i1)5)1/(6\Delta^{7}(s+i-1)^{5}). Therefore, defining T=S{v1,vs}T=S\cup\{v_{1},\ldots v_{s}\}, we have

λ1(A~T)λ1+16Δ7i=1s1(s+i1)5λ1+5128Δ7s4,\lambda_{1}(\tilde{A}_{T})\geq\lambda_{1}+\frac{1}{6\Delta^{7}}\sum_{i=1}^{s}\frac{1}{(s+i-1)^{5}}\geq\lambda_{1}+\frac{5}{128\Delta^{7}s^{4}},

where the last inequality follows from approximating the sum with the integral. As λ11\lambda_{1}\leq 1, this translates to the desired multiplicative bound. ∎

Proof of Theorem 3.1.

We begin by showing (14). Let Γxs\Gamma_{x}^{s} be the set of connected subgraphs of GG with ss vertices containing xx. Choose SS to be the maximizer of exTA~S2kexe_{x}^{T}\tilde{A}_{S}^{2k}e_{x} among SΓxsS\in\Gamma_{x}^{s}, and let TΓx2sT\in\Gamma_{x}^{2s} be the extension of SS guaranteed by Lemma 3.3 to satisfy

λ1(A~T)(1+5128Δ7s4)λ1(A~S).\lambda_{1}(\tilde{A}_{T})\geq\left(1+\frac{5}{128\Delta^{7}s^{4}}\right)\lambda_{1}(\tilde{A}_{S}).

PS2kP^{2k}_{S} has the same diagonal entries as A~S2k\tilde{A}^{2k}_{S}, so

x(W2k,s)SΓxsexTA~S2kex,\mathbb{P}_{x}(W^{2k,s})\leq\sum_{S^{\prime}\in\Gamma_{x}^{s}}e_{x}^{T}\tilde{A}_{S^{\prime}}^{2k}e_{x},

since each walk of length 2k2k satisfying W2k,sW^{2k,s} is contained in at least one SΓxsS^{\prime}\in\Gamma_{x}^{s}. Furthermore, |Γxs|Δ2s|\Gamma_{x}^{s}|\leq\Delta^{2s} since each subgraph of Γxs\Gamma_{x}^{s} may be encoded by one of its spanning trees, which may in turn be encoded by a closed walk rooted at xx traversing the edges of the tree. We then have

x(W2k,s)\displaystyle\mathbb{P}_{x}(W^{2k,s}) |Γxs|exTA~S2kex\displaystyle\leq|\Gamma_{x}^{s}|e_{x}^{T}\tilde{A}_{S}^{2k}e_{x}
Δ2sλ1(A~S)2k\displaystyle\leq\Delta^{2s}\lambda_{1}(\tilde{A}_{S})^{2k}
Δ2s(1+5128Δ7s4)2kλ1(A~T)2k.\displaystyle\leq\Delta^{2s}\left(1+\frac{5}{128\Delta^{7}s^{4}}\right)^{-2k}\lambda_{1}(\tilde{A}_{T})^{2k}. (16)

We will bound the right hand side in terms of x(W2k,2s)\mathbb{P}_{x}(W^{2k,2s}).

We claim that for every zTz\in T,

exTA~T2kexΔ4sezTA~T2k4sez.e_{x}^{T}\tilde{A}_{T}^{2k}e_{x}\geq\Delta^{-4s}e_{z}^{T}\tilde{A}_{T}^{2k-4s}e_{z}. (17)

To see this, let π\pi be a path in TT of length 2s\ell\leq 2s between xx and zz, which must exist since TT is connected and has size 2s2s. Then every closed walk of length 2k22k-2\ell in TT rooted at zz may be extended to a walk of length 2k2k in TT rooted at xx by attaching π\pi and its reverse. Performing the walk of π\pi twice occurs with probability at least Δ2\Delta^{-2\ell}. Since all of the walks produced this way are distinct, we have

exTA~T2kexΔ2ezTA~T2k2ez.e_{x}^{T}\tilde{A}_{T}^{2k}e_{x}\geq\Delta^{-2\ell}e_{z}^{T}\tilde{A}_{T}^{2k-2\ell}e_{z}.

By the same argument ezTA~T2k2ezΔ4s+2ezTA~T2k4seze_{z}^{T}\tilde{A}_{T}^{2k-2\ell}e_{z}\geq\Delta^{-4s+2\ell}e_{z}^{T}\tilde{A}_{T}^{2k-4s}e_{z}, and inequality (17) follows.

Choose zTz\in T to be the maximizer of ezTA~T2k4seze_{z}^{T}\tilde{A}_{T}^{2k-4s}e_{z}, for which we have:

ezTA~T2k4sez12sTr(PT2k4s)λ1(A~T)2k4s2s.e_{z}^{T}\tilde{A}_{T}^{2k-4s}e_{z}\geq\frac{1}{2s}\operatorname{Tr}(P_{T}^{2k-4s})\geq\frac{\lambda_{1}(\tilde{A}_{T})^{2k-4s}}{2s}.

Combining this with (17) and substituting in (16), we obtain

x(W2k,s)\displaystyle\mathbb{P}_{x}(W^{2k,s}) Δ6s2s(1+5128Δ7s4)2kλ1(A~T)4sexTA~T2kex\displaystyle\leq\Delta^{6s}\cdot 2s\left(1+\frac{5}{128\Delta^{7}s^{4}}\right)^{-2k}\lambda_{1}(\tilde{A}_{T})^{4s}e_{x}^{T}\tilde{A}_{T}^{2k}e_{x}
Δ6s2s(1+5128Δ7s4)2kλ1(A~T)4sx(W2k,2s).\displaystyle\leq\Delta^{6s}\cdot 2s\left(1+\frac{5}{128\Delta^{7}s^{4}}\right)^{-2k}\lambda_{1}(\tilde{A}_{T})^{4s}\mathbb{P}_{x}(W^{2k,2s}).

Applying the inequality ex/21+xe^{x/2}\leq 1+x for 0<x<10<x<1 and the bound λ1(A~T)<1\lambda_{1}(\tilde{A}_{T})<1, we obtain

x(W2k,s)exp(6slogΔ+log(2s)5k128Δ7s4)x(W2k,2s),\mathbb{P}_{x}(W^{2k,s})\leq\exp\left(6s\log\Delta+\log(2s)-\frac{5k}{128\Delta^{7}s^{4}}\right)\mathbb{P}_{x}(W^{2k,2s}), (18)

which implies

x(W2k,s)exp(k65Δ7s4)x(W2k,2s)\mathbb{P}_{x}(W^{2k,s})\leq\exp\left(-\frac{k}{65\Delta^{7}s^{4}}\right)\mathbb{P}_{x}(W^{2k,2s})

whenever

s14(kΔ7log(Δ))1/5,s\leq\frac{1}{4}\left(\frac{k}{\Delta^{7}\log(\Delta)}\right)^{1/5},

establishing (14).

4 Bound on Eigenvalue Multiplicity

In this section we prove Theorem 1.2, restated below in slightly more detail.

Theorem 4.1 (Detailed Theorem 1.2).

Let GG be a maximum degree Δ\Delta connected graph on nn vertices. If666If Δlog1/7n/loglogn\Delta\geq\log^{1/7}n/\log\log n then (1) is vacuously true. Δlog1/7n/loglogn\Delta\leq\log^{1/7}n/\log\log n then the spectrum of the normalized adjacency matrix A~\tilde{A} satisfies

mG([(1loglogΔnlogΔn)λ2,λ2])=O(nΔ7/5(log2/5Δ)loglognlog1/5n).m_{G}\left([(1-\frac{\log\log_{\Delta}n}{\log_{\Delta}n})\lambda_{2},\lambda_{2}]\right)=O\left(n\cdot\frac{\Delta^{7/5}(\log^{2/5}\Delta)\log\log n}{\log^{1/5}n}\right). (19)
Proof.

For now, assume that |λn(P)||λ2(P)|\left|\lambda_{n}(P)\right|\leq\left|\lambda_{2}(P)\right|. Let ()\mathbb{P}(\cdot) denote the law of an SRW γ\gamma of length 2k2k on GG, started at a vertex chosen uniformly at random (i.e., not from the stationary measure of the SRW). Let W2k:=W2k,nW^{2k}:=W^{2k,n} denote the event that γ\gamma returns to its starting vertex after 2k2k steps. In an abuse of notation, let W2k,s+1:=W2k\W2k,sW^{2k,\geq s+1}:=W^{2k}\backslash W^{2k,s} be the event that a walk of length 2k2k is closed and has support at least s+1s+1.

Set k:=13logΔnk:=\frac{1}{3}\log_{\Delta}n and c:=2logkc:=2\log k and let ss be a parameter satisfying

(W2k,s)ec(W2k)\mathbb{P}(W^{2k,s})\leq e^{-c}\mathbb{P}(W^{2k}) (20)

to be chosen later. Delete cn/scn/s vertices from GG uniformly at random and call the resulting graph HH.

If γ\gamma has support at least s+1s+1, then the probability that none of the vertices of γ\gamma are deleted is at most

(1sn)cnsec.\left(1-\frac{s}{n}\right)^{\frac{cn}{s}}\leq e^{-c}.

Thus,

𝔼H(γH|γW2k,s+1)ec,\mathbb{E}_{H}\mathbb{P}(\gamma\subset H|\gamma\in W^{2k,\geq s+1})\leq e^{-c},

where 𝔼H\mathbb{E}_{H} is the expectation over HH. It then follows by the probabilistic method that there exists a deletion such that the resulting subgraph HH of GG satisfies

(W2k,s+1{γH})ec(W2k,s+1).\mathbb{P}(W^{2k,\geq s+1}\cap\{\gamma\subset H\})\leq e^{-c}\mathbb{P}(W^{2k,\geq s+1}).

Write λ2:=λ2(A~G)\lambda_{2}:=\lambda_{2}(\tilde{A}_{G}) and let mm^{\prime} be the number of eigenvalues of HH in the interval [(1ϵ)λ2,λ2][(1-\epsilon)\lambda_{2},\lambda_{2}] for ϵ:=c/2logΔ(n)\epsilon:=c/2\log_{\Delta}(n). Since 2k2k is even,

m(1ϵ)2kλ22k\displaystyle m^{\prime}(1-\epsilon)^{2k}\lambda_{2}^{2k} Tr(A~H2k)\displaystyle\leq\operatorname{Tr}(\tilde{A}_{H}^{2k})
=n(W2k{γH})\displaystyle=n\mathbb{P}(W^{2k}\cap\{\gamma\subset H\})
=n((W2k,s{γH})+(W2k,s+1{γH}))\displaystyle=n(\mathbb{P}(W^{2k,s}\cap\{\gamma\subset H\})+\mathbb{P}(W^{2k,\geq s+1}\cap\{\gamma\subset H\}))
n((W2k,s)+ec(W2k,s+1))by our choice of H\displaystyle\leq n(\mathbb{P}(W^{2k,s})+e^{-c}\mathbb{P}(W^{2k,\geq s+1}))\quad\textrm{by our choice of }H
n(ec(W2k)+ec(W2k,s+1))by (20)\displaystyle\leq n(e^{-c}\mathbb{P}(W^{2k})+e^{-c}\mathbb{P}(W^{2k,\geq s+1}))\quad\textrm{by \eqref{eqn:sets}}
2ecTr(A~G2k)\displaystyle\leq 2e^{-c}\operatorname{Tr}(\tilde{A}_{G}^{2k})
2ec(nλ22k+1).\displaystyle\leq 2e^{-c}(n\lambda_{2}^{2k}+1).

We may assume that the diameter of GG is at least 44 as otherwise Δn1/4\Delta\geq n^{1/4}, making the theorem statement vacuous. Because of the diameter, we can take two edges (u,v)(u,v) and (a,b)(a,b), such that the distance between the edges is at least 22. Then consider the vector ϕ\phi such that

ϕ(x)={1x{u,v}1x{a,b}0 otherwise\phi(x)=\left\{\begin{array}[]{cc}1&x\in\{u,v\}\\ -1&x\in\{a,b\}\\ 0&\textnormal{ otherwise}\end{array}\right.

As the top eigenvector of the normalized adjacency matrix is the all ones vector, ϕ\phi is orthogonal to it. Therefore by Courant Fisher

λ2ϕTD1/2AD1/2ϕϕTϕ4Δ14=1Δ.\lambda_{2}\geq\frac{\phi^{T}D^{-1/2}AD^{-1/2}\phi}{\phi^{T}\phi}\geq\frac{4\Delta^{-1}}{4}=\frac{1}{\Delta}.

By our choice of kk, this means nλ22k1n\lambda_{2}^{2k}\geq 1. Moreover,

ϵ2loglogn2logΔnlogΔloglognlogn<1/2,\displaystyle\epsilon\leq\frac{2\log\log n}{2\log_{\Delta}n}\leq\frac{\log\Delta\log\log n}{\log n}<1/2,

based on our assumptions on Δ\Delta. Thus, 1ϵe1.5ϵ1-\epsilon\geq e^{-1.5\epsilon}. Combining these facts,

mλ22k4e3kϵcnλ22k,m^{\prime}\lambda_{2}^{2k}\leq 4e^{3k\epsilon-c}n\lambda_{2}^{2k},

yielding

m4ne3kϵc4nec/2=O(nlogΔn).m^{\prime}\leq 4ne^{3k\epsilon-c}\leq 4ne^{-c/2}=O\left(\frac{n}{\log_{\Delta}n}\right).

As we created HH by deleting cn/scn/s vertices, it follows by Cauchy interlacing that the number of eigenvalues of A~\tilde{A} in [(1ϵ)λ2,λ2][(1-\epsilon)\lambda_{2},\lambda_{2}] is at most

cns+O(nlogΔn).\frac{cn}{s}+O\left(\frac{n}{\log_{\Delta}n}\right).

We now show that taking

s:=14(kΔ7logΔ)1/5s:=\frac{1}{4}\left(\frac{k}{\Delta^{7}\log\Delta}\right)^{1/5}

satisfies (20). Applying Theorem 3.1 equation (14) to each xGx\in G and summing, we have

(W2k,s)(W2k)\displaystyle\frac{\mathbb{P}(W^{2k,s})}{\mathbb{P}(W^{2k})} exp(k65Δ7s4)\displaystyle\leq\exp\left(-\frac{k}{65\Delta^{7}s^{4}}\right)
exp(Ω(lognlog2/5ΔΔ7/5))\displaystyle\leq\exp\left(-\Omega\left(\frac{\log n\log^{2/5}\Delta}{\Delta^{7/5}}\right)\right)
exp(c)=exp(Θ(loglogΔn)),\displaystyle\ll\exp(-c)=\exp(-\Theta(\log\log_{\Delta}n)),

satisfying (20) for sufficiently large nn, and we conclude that

mG([(1loglogΔnlogΔn)λ2,λ2])=O(nΔ7/5log2/5Δloglognlog1/5n),m_{G}\left([(1-\frac{\log\log_{\Delta}n}{\log_{\Delta}n})\lambda_{2},\lambda_{2}]\right)=O\left(n\cdot\frac{\Delta^{7/5}\log^{2/5}\Delta\log\log n}{\log^{1/5}n}\right),

as desired.

If |λn|>|λ2|\left|\lambda_{n}\right|>\left|\lambda_{2}\right|, we can do a lazy walk with probability of moving p=12p=\frac{1}{2}, therefore making all eigenvalues nonnegative. This is equivalent to doubling the degree of every vertex by adding loops. This is the equivalent of taking the simple random walk on a graph with maximum degree 2Δ2\Delta, requiring s111(kΔ7logΔ)1/5s\leq\frac{1}{11}\left(\frac{k}{\Delta^{7}\log\Delta}\right)^{1/5}, yielding the same asymptotics. ∎

5 Examples

In this section, we consider examples demonstrating some of the points raised in the introduction regarding the tightness of our results. As most of our results in this section are combinatorial rather than probabilistic, we will consider multiplicity in the non-normalized adjacency matrix AA. For regular graphs, this is equivalent.

5.1 Bipartite Ramanujan Graphs

We show that bipartite Ramanujan graphs (see [LPS88]; known to exist for every d3d\geq 3 by [MSS15]) have high multiplicity near λ2\lambda_{2}. This means that the bound of n/logΘ(1)nn/\log^{\Theta(1)}n of Theorem 1.2 is tight.

Lemma 5.1 (McKay [McK81] Lemma 3).

The number of closed walks on the infinite dd-regular tree of length 2k2k starting at a fixed vertex is Θ(4k(d1)kk3/2)\Theta\left(\frac{4^{k}(d-1)^{k}}{k^{3/2}}\right).

Proposition 5.2.

There exists a constant α>0\alpha>0 such that for fixed dd, every bipartite dd-regular bipartite Ramanujan graph GG on nn vertices satisfies

mG([λ2(1αloglog(n)log(n)),λ2])=Ω(n/log3/2(n)).m_{G}\left([\lambda_{2}(1-\alpha\frac{\log\log(n)}{\log(n)}),\lambda_{2}]\right)=\Omega(n/\log^{3/2}(n)).
Proof.

Let k=βlognk=\beta\log n for some constant β\beta to be set later and suppose that there are mm eigenvalues of AGA_{G} in the interval [λ2(1αloglog(n)log(n)),λ2][\lambda_{2}\left(1-\alpha\frac{\log\log(n)}{\log(n)}\right),\lambda_{2}]. Recall that the spectrum of a bipartite graph is symmetric around 0. From Lemma 5.1 it follows that for some constant CC,

Cn(4k(d1)kk3/2)\displaystyle Cn\left(\frac{4^{k}(d-1)^{k}}{k^{3/2}}\right) i=1nλi(AG)2k\displaystyle\leq\sum_{i=1}^{n}\lambda_{i}(A_{G})^{2k}
2d2k+(n2m)(2d1(1αloglog(n)log(n)))2k+2m(2d1)2k.\displaystyle\leq 2d^{2k}+(n-2m)\left(2\sqrt{d-1}\left(1-\alpha\frac{\log\log(n)}{\log(n)}\right)\right)^{2k}+2m(2\sqrt{d-1})^{2k}.

If we let β\beta be sufficiently small and α>32β\alpha>\frac{3}{2\beta}, rearranging yields

mn\displaystyle\frac{m}{n} C4k(d1)kk3/22d2kn(2d1(1αloglog(n)log(n)))2k2(2d1)2k(1(1αloglog(n)log(n))2k)\displaystyle\geq\frac{C\frac{4^{k}(d-1)^{k}}{k^{3/2}}-\frac{2d^{2k}}{n}-\left(2\sqrt{d-1}\left(1-\alpha\frac{\log\log(n)}{\log(n)}\right)\right)^{2k}}{2(2\sqrt{d-1})^{2k}\cdot\left(1-\left(1-\alpha\frac{\log\log(n)}{\log(n)}\right)^{2k}\right)}
=Ω(12n2β1k3/2(1αloglog(n)log(n))2k1(1αloglog(n)log(n))2k)\displaystyle=\Omega\left(\frac{1-2n^{2\beta-1}}{k^{3/2}}-\frac{\left(1-\alpha\frac{\log\log(n)}{\log(n)}\right)^{2k}}{1-\left(1-\alpha\frac{\log\log(n)}{\log(n)}\right)^{2k}}\right)
=Ω(1k3/21e2αβloglog(n))\displaystyle=\Omega\left(\frac{1}{k^{3/2}}-\frac{1}{e^{2\alpha\beta\log\log(n)}}\right)
=Ω(1k3/2).\displaystyle=\Omega\left(\frac{1}{k^{3/2}}\right).

5.2 Mangrove Tree

This section shows that the dependence on |V||V| in Corollary 1.5 is tight up to polylogarithmic factors. Our example begins with a path of multiedges containing nn vertices, where each multiedge of the path is composed of d/2d/2 edges for some even dd. At both ends of the path, we attach a tree of depth logd1n\log_{d-1}n. The roots have degree d/2d/2 and all other vertices (besides the leaves) have degree dd. Therefore the only vertices in the graph that are not degree dd are the leaves of the two trees. Call this graph QQ. An example of this graph is shown in Figure 2.

Proposition 5.3.

For every vertex uu of degree less than dd, ψQ(u)=O~(n5/2)\psi_{Q}(u)=\tilde{O}(n^{-5/2}), where O~\tilde{O} suppresses dependence on logarithmic factors and dd.

Therefore, we cannot hope to do significantly better than our analysis in Lemma 3.3, in which we find a vertex uu of non-maximal degree with ψ(u)1/(dλ1n5/2)\psi(u)\geq 1/(d\lambda_{1}n^{5/2}).

Proof.

For simplicity, call λ1(AQ)=λ1\lambda_{1}(A_{Q})=\lambda_{1} and ψQ=ψ\psi_{Q}=\psi. By the symmetry of the graph, the value of ψ\psi at vertices in the tree is determined by the distance from the root. Call the entries of ψ\psi corresponding to the tree r0,r1,,rr_{0},r_{1},\ldots,r_{\ell}, where the index indicates the distance from the root.

By the discussion in the proof of Kahale [Kah95] Lemma 3.3, if we define

θ:=log(λ12d1+λ124(d1)1),\theta:=\log\left(\frac{\lambda_{1}}{2\sqrt{d-1}}+\sqrt{\frac{\lambda_{1}^{2}}{4(d-1)}-1}\right),

then for 0i0\leq i\leq\ell, entries of the eigenvector must satisfy

rir0=sinh((+1i)θ)(d1)i/2sinh((+1)θ)\frac{r_{i}}{r_{0}}=\frac{\sinh((\ell+1-i)\theta)(d-1)^{-i/2}}{\sinh((\ell+1)\theta)}

where \ell is the depth of the tree.

Therefore, r/r0=sinh(θ)(d1)/2sinh((+1)θ)r_{\ell}/r_{0}=\frac{\sinh(\theta)(d-1)^{-\ell/2}}{\sinh((\ell+1)\theta)}. Examining the various terms, sinh(θ)d\sinh(\theta)\leq d and (d1)/2=1n(d-1)^{-\ell/2}=\frac{1}{\sqrt{n}}. To bound the third term, we use the definition sinh(x)=(exex)/2\sinh(x)=(e^{x}-e^{-x})/2, which yields

sinh((+1)θ)1on(1)2(λ12d1+λ124(d1)1)logd1n+1.\sinh((\ell+1)\theta)\geq\frac{1-o_{n}(1)}{2}\left(\frac{\lambda_{1}}{2\sqrt{d-1}}+\sqrt{\frac{\lambda_{1}^{2}}{4(d-1)}-1}\right)^{\log_{d-1}n+1}.

λ1\lambda_{1} is at least the spectral radius of the path of length nn with d/2d/2 multiedges between vertices. This spectral radius is dcos(π/(n+1))d\cos(\pi/(n+1)). This gives

sinh((+1)θ)\displaystyle\sinh((\ell+1)\theta) \displaystyle\geq 1on(1)2(2d1)logd1n+1(d(1π22n2)+d2(1π22n2)24d+4)logd1n+1\displaystyle\frac{1-o_{n}(1)}{2(2\sqrt{d-1})^{\log_{d-1}n+1}}\left(d(1-\frac{\pi^{2}}{2n^{2}})+\sqrt{d^{2}(1-\frac{\pi^{2}}{2n^{2}})^{2}-4d+4}\right)^{\log_{d-1}n+1}
\displaystyle\geq 1on(1)2(2d1)logd1n+1(d+d2)logd1n+1(1O(dn2))logd1n+1\displaystyle\frac{1-o_{n}(1)}{2(2\sqrt{d-1})^{\log_{d-1}n+1}}\left(d+d-2\right)^{\log_{d-1}n+1}\left(1-O\left(\frac{d}{n^{2}}\right)\right)^{\log_{d-1}n+1}
\displaystyle\geq 1on(1)2eO(dlogd1n/n2)nn3\displaystyle\frac{1-o_{n}(1)}{2}e^{-O({d\log_{d-1}n}/{n^{2}})}\sqrt{n}\geq\frac{\sqrt{n}}{3}

for large enough nn. Therefore

rr0=sinh(θ)(d1)/2sinh((+1)θ)3dn.\frac{r_{\ell}}{r_{0}}=\frac{\sinh(\theta)(d-1)^{-\ell/2}}{\sinh((\ell+1)\theta)}\leq\frac{3d}{n}. (21)

At this point, we know the ratio between rr_{\ell} and r0r_{0}, but still need to bound the overall mass of the eigenvector on the tree. A “regular partition” is a partition of vertices V=i=0kXjV=\bigsqcup_{i=0}^{k}X_{j} where the number of neighbors a vertex uXiu\in X_{i} has in XjX_{j} does not depend on uu. We can create a quotient matrix, where entry i,ji,j corresponds to the number of neighbors a vertex uXiu\in X_{i} has in XjX_{j}. For an overview of quotient matrices and their utility, see Godsil, [God93, Chapter 5]. In our partition, every vertex in the path is placed in a set by itself. The vertices of each of the two trees are partitioned into sets according to their distance from the two roots. Call the matrix according to this partition BQB_{Q}. We denote by BQ(Xi,Xj)B_{Q}(X_{i},X_{j}) the entry in BQB_{Q} corresponding to edges from a vertex in XiX_{i} to XjX_{j}.

Define X0,XX_{0},\ldots X_{\ell} as the sets corresponding to vertices in the first tree of distance 0,,0,\ldots,\ell from the root. For 1j11\leq j\leq\ell-1, BQ(X0,X1)=d/2B_{Q}(X_{0},X_{1})=d/2. BQ(Xj,Xj+1)=d1B_{Q}(X_{j},X_{j+1})=d-1. Moreover, for 0j10\leq j\leq\ell-1, BQ(Xj+1,Xj)=1B_{Q}(X_{j+1},X_{j})=1. All values between vertices in the path are unchanged at d/2d/2.

Consider the diagonal matrix DD with Di,i:=|Xi|1/2D_{i,i}:=|X_{i}|^{-1/2}. D1BQDD^{-1}B_{Q}D is a symmetric matrix. Define C:=D1BQDC:=D^{-1}B_{Q}D We now have C(Xj+1,Xj)=C(Xj,Xj+1)=d1C(X_{j+1},X_{j})=C(X_{j},X_{j+1})=\sqrt{d-1} for 1j11\leq j\leq\ell-1, and C(X0,X1)=C(X1,X0)=d/2C(X_{0},X_{1})=C(X_{1},X_{0})=\sqrt{d/2}.

If a vector ϕ\phi is an eigenvector of CC, then DϕD\phi is an eigenvector of BQB_{Q} with the same eigenvalue. By the definition of DD this means

ψC(Xi)2=uXiψAQ(u)2.\psi_{C}(X_{i})^{2}=\sum_{u\in X_{i}}\psi_{A_{Q}}(u)^{2}. (22)

Define CX0:C_{X_{0:\ell}} as the submatrix of CC corresponding the the sets {X0,,X}\{X_{0},\ldots,X_{\ell}\}, then extended with zeros to have the same size as CC. Every entry of C+d/2d1d1CX0:C+\frac{d/2-\sqrt{d-1}}{\sqrt{d-1}}C_{X_{0:\ell}} is less than or equal to the corresponding entry of the adjacency matrix of a path of length n+2logd1nn+2\log_{d-1}n with d/2d/2 edges between pairs of vertices. Also, ψC\psi_{C} is a nonnegative vector. Therefore the quadratic form ψCT(C+d/2d1d1CX0:)ψC\psi_{C}^{T}(C+\frac{d/2-\sqrt{d-1}}{\sqrt{d-1}}C_{X_{0:\ell}})\psi_{C} is at most the spectral radius of this path. Namely

ψCTCψC+d/2d1d1ψCTCX0:ψCdcos(π/(n+2logd1n+1)).\psi_{C}^{T}C\psi_{C}+\frac{d/2-\sqrt{d-1}}{\sqrt{d-1}}\psi_{C}^{T}C_{X_{0:\ell}}\psi_{C}\leq d\cos(\pi/(n+2\log_{d-1}n+1)).

Because CC contains the path of length nn, ψCTCψCdcos(π/(n+1))\psi_{C}^{T}C\psi_{C}\geq d\cos(\pi/(n+1)). Putting these together yields

ψCTCX0:ψCd1d/2d1d(cos(π/(n+2logd1n+1))cos(π/(n+1)))dd1d/2d13π2logdnn3.\psi_{C}^{T}C_{X_{0:\ell}}\psi_{C}\leq\frac{\sqrt{d-1}}{d/2-\sqrt{d-1}}\cdot d(\cos(\pi/(n+2\log_{d-1}n+1))-\cos(\pi/(n+1)))\leq\frac{d\sqrt{d-1}}{d/2-\sqrt{d-1}}\frac{3\pi^{2}\log_{d}n}{n^{3}}. (23)

Define ψC(X1:)\psi_{C}(X_{1:\ell}) as the projection of ψC\psi_{C} on {X1,X}\{X_{1},\ldots X_{\ell}\}. CψC(X1:)=CX0:ψC(X1:)C\psi_{C}(X_{1:\ell})=C_{X_{0:\ell}}\psi_{C}(X_{1:\ell}), so

ψCTCX0:ψCλ1ψC(X1:)2d(cos(π/(n+1)))ψC(X1:)2\psi_{C}^{T}C_{X_{0:\ell}}\psi_{C}\geq\lambda_{1}\|\psi_{C}(X_{1:\ell})\|^{2}\geq d(\cos(\pi/(n+1)))\|\psi_{C}(X_{1:\ell})\|^{2} (24)

Combining (23) and (24) yields

ψC(X1:)2d1d/2d1(3π2logdnn3)/cos(π/n+1)(21π2logdnn3)\|\psi_{C}(X_{1:\ell})\|^{2}\leq\frac{\sqrt{d-1}}{d/2-\sqrt{d-1}}\left(\frac{3\pi^{2}\log_{d}n}{n^{3}}\right)/\cos(\pi/n+1)\leq\left(\frac{21\pi^{2}\log_{d}n}{n^{3}}\right)

assuming d4d\geq 4 and nn is sufficiently large.

Using (22)(\ref{eq:mass}) and the eigenvalue equation, we obtain

ψQ(r0)=ψC(X0)λ1(AC)ψC(X1:)d5πlogd1/2nn3/2.\psi_{Q}(r_{0})=\psi_{C}(X_{0})\leq\lambda_{1}(A_{C})\|\psi_{C}(X_{1:\ell})\|\leq d\cdot\frac{{5}\pi\log^{1/2}_{d}n}{n^{3/2}}.

Therefore, according to (21)

r15d2π(logd1/2n)n5/2.r_{\ell}\leq\frac{15d^{2}\pi(\log^{1/2}_{d}n)}{n^{5/2}}.

6 Open Problems

We conclude with some promising directions for further research.

Beyond the Trace Method: Polynomial Multiplicity Bounds

There is a large gap between our upper bound of O(n/log1/5n)O(n/\log^{1/5}n) on the multiplicity of the second eigenvalue and the lower bound of n1/3n^{1/3} mentioned after Theorem 1.1. It is very natural to ask, whether the bound of this paper may be improved. To improve the bound beyond O(n/polylog(n))O(n/\textnormal{polylog}(n)), however, it appears that a very different approach is needed.

Open Problem 1 (Similar to Question 6.3 of [JTY+19]).

Let d>1d>1 be fixed integer. Does there exist an ε>0\varepsilon>0 such that for every connected dd-regular graph GG on nn vertices, the multiplicity of the second largest eigenvalue of AGA_{G} is O(n1ε)O(n^{1-\varepsilon})?

In the present paper, we rely on the trace method to bound eigenvalue multiplicity through closed walks. There are three drawbacks to this approach that stops a bound on the second eigenvalue multiplicity below n/polylog(n)n/\textnormal{polylog}(n). First, considering walks of length ω(log(n))\omega(\log(n)) makes the top eigenvalue dominate the trace, leaving no information behind. Second, considering the trace TrAGk\operatorname{Tr}A_{G}^{k} for k=O(log(n))k=O(\log(n)) it is impossible to distinguish eigenvalues that differ by O(1/log(n))O(1/\log(n)). Third, as covered in Section 5.1, there exist graphs such that there are Ω(n/polylog(n))\Omega(n/\textnormal{polylog}(n)) eigenvalues in a range of that size around the second eigenvalue. Thus, the trace method reaches a natural barrier at n/polylog(n))n/\textnormal{polylog}(n)).

Eigenvalue Multiplicity for Unnormalized Non-Regular Graphs

Another natural question is whether Theorem 1.2 may be extended to hold for the (non-normalized) adjacency matrix of non-regular graphs.

Open Problem 2.

Let Δ>1\Delta>1 be a fixed integer. Does it hold for every connected graph GG on nn vertices of maximum degree Δ\Delta that the multiplicity of the second largest eigenvalue of AGA_{G} is o(n/loglog(n))o(n/\log\log(n))?

In order to handle unnormalized irregular graphs via the approach in this paper, the key ingredient needed would be an “unnormalized” analogue of Theorem 1.3, showing that a uniformly random closed walk (from the set of all closed walks) in an irregular graph must have large support. We exhibit in Appendix B an irregular “lollipop” graph for which the typical support of a closed walk from a specific vertex is only O(polylog(k))O(\textnormal{polylog}(k)). It remains plausible that when starting from a random vertex, a randomly selected closed walk has poly(k)(k) support in irregular graphs.

Sharper Bounds for Closed Random Walks

We have no reason to believe that the exponent of 1/51/5 appearing in Theorem 1.3 is sharp. In fact, we know of no example where where the answer is o(k1/2)o(k^{1/2}). An improvement over Theorem 1.3 would immediately yield an improvement of Theorem 1.2.

Open Problem 3.

Let d>1d>1 be a fixed integer. Does there exist an α>1/5\alpha>1/5 such that for every connected dd-regular graph GG on nn vertices and every vertex xx of GG, a random closed walk of length 2k<n2k<n rooted at xx has support Ω(kα)\Omega(k^{\alpha}) in expectation? Is α=1/2\alpha=1/2 true? Does such a bound hold for SRW in general?

Acknowledgments

We thank Yufei Zhao for telling us about [JTY+19] at the Simons Foundation conference on High Dimensional Expanders in October, 2019. We thank Shirshendu Ganguly for helpful discussions. We thank Cyril Letrouit for pointing out an error in the previous proof of Proposition 5.2.

References

  • [ABS15] Sanjeev Arora, Boaz Barak, and David Steurer. Subexponential algorithms for unique games and related problems. Journal of the ACM (JACM), 62(5):1–25, 2015.
  • [BGH+15] Boaz Barak, Parikshit Gopalan, Johan Håstad, Raghu Meka, Prasad Raghavendra, and David Steurer. Making the long code shorter. SIAM Journal on Computing, 44(5):1287–1324, 2015.
  • [Bol13] Béla Bollobás. Modern graph theory, volume 184. Springer Science & Business Media, 2013.
  • [BRS11] Boaz Barak, Prasad Raghavendra, and David Steurer. Rounding semidefinite programming hierarchies via global correlation. In 2011 ieee 52nd annual symposium on foundations of computer science, pages 472–481. IEEE, 2011.
  • [CG07] Sebastian Cioabă and David Gregory. Principal eigenvectors of irregular graphs. The Electronic Journal of Linear Algebra, 16, 2007.
  • [Cio07] Sebastian M Cioabă. The spectral radius and the maximum degree of irregular graphs. The Electronic Journal of Combinatorics, 14(1):R38, 2007.
  • [CRS93] Dragoš Cvetković, Peter Rowlinson, and Slobodan Simić. A study of eigenspaces of graphs. Linear algebra and its applications, 182:45–66, 1993.
  • [CVDKL10] Sebastian M Cioabă, Edwin R Van Dam, Jack H Koolen, and Jae-Ho Lee. A lower bound for the spectral radius of graphs with fixed diameter. European Journal of Combinatorics, 31(6):1560–1566, 2010.
  • [DS84] Peter G Doyle and J Laurie Snell. Random walks and electric networks, volume 22. American Mathematical Soc., 1984.
  • [Fri91] Joel Friedman. Some geometric aspects of graphs and their eigenfunctions. Princeton University, Department of Computer Science, 1991.
  • [Gil98] David Gillman. A chernoff bound for random walks on expander graphs. SIAM Journal on Computing, 27(4):1203–1220, 1998.
  • [God93] Chris Godsil. Algebraic combinatorics, volume 6. CRC Press, 1993.
  • [JTY+19] Zilin Jiang, Jonathan Tidor, Yuan Yao, Shengtong Zhang, and Yufei Zhao. Equiangular lines with a fixed angle. arXiv preprint arXiv:1907.12466, 2019.
  • [Kah95] Nabil Kahale. Eigenvalues and expansion of regular graphs. Journal of the ACM (JACM), 42(5):1091–1106, 1995.
  • [Kah97] Nabil Kahale. Large deviation bounds for markov chains. Combinatorics Probability and Computing, 6(4):465–474, 1997.
  • [Kol11] Alexandra Kolla. Spectral algorithms for unique games. computational complexity, 20(2):177–206, 2011.
  • [LM08] James R Lee and Yury Makarychev. Eigenvalue multiplicity and volume growth. arXiv preprint arXiv:0806.1745, 2008.
  • [LOG18] Russell Lyons and Shayan Oveis Gharan. Sharp bounds on random walk eigenvalues via spectral embedding. International Mathematics Research Notices, 2018(24):7555–7605, 2018.
  • [LOGT14] James R Lee, Shayan Oveis Gharan, and Luca Trevisan. Multiway spectral partitioning and higher-order cheeger inequalities. Journal of the ACM (JACM), 61(6):1–30, 2014.
  • [LPS88] Alexander Lubotzky, Ralph Phillips, and Peter Sarnak. Ramanujan graphs. Combinatorica, 8(3):261–277, 1988.
  • [LRTV12] Anand Louis, Prasad Raghavendra, Prasad Tetali, and Santosh Vempala. Many sparse cuts via higher eigenvalues. In Proceedings of the forty-fourth annual ACM symposium on Theory of computing, pages 1131–1140, 2012.
  • [McK81] Brendan D McKay. The expected eigenvalue distribution of a large regular graph. Linear Algebra and its Applications, 40:203–216, 1981.
  • [MSS15] Adam W Marcus, Daniel A Spielman, and Nikhil Srivastava. Interlacing families i: Bipartite ramanujan graphs of all degrees. Annals of Mathematics, 182:307–325, 2015.
  • [OGT13] Shayan Oveis Gharan and Luca Trevisan. A new regularity lemma and faster approximation algorithms for low threshold rank graphs. In Approximation, Randomization, and Combinatorial Optimization. Algorithms and Techniques, pages 303–316. Springer, 2013.
  • [Row90] Peter Rowlinson. More on graph perturbations. Bulletin of the London Mathematical Society, 22(3):209–216, 1990.
  • [Row19] Peter Rowlinson. Eigenvalue multiplicity in regular graphs. Discrete Applied Mathematics, 269:11–17, 2019.
  • [Ste14] Dragan Stevanovic. Spectral radius of graphs. Academic Press, 2014.
  • [TT15] Michael Tait and Josh Tobin. Characterizing graphs of maximum principal ratio. arXiv preprint arXiv:1511.06378, 2015.
  • [VMSK+11] Piet Van Mieghem, Dragan Stevanović, Fernando Kuipers, Cong Li, Ruud Van De Bovenkamp, Daijie Liu, and Huijuan Wang. Decreasing the spectral radius of a graph by link removals. Physical Review E, 84(1):016101, 2011.

Appendix A Proofs for high degree regular graphs

Theorem A.1 (Detailed Theorem 1.8).

If GG is dd-regular, has exactly hh self-loops at every vertex, and no multi-edges777This technical assumption is used to handle the case when |λn(AG)|>λ2(AG)|\lambda_{n}(A_{G})|>\lambda_{2}(A_{G}) in Theorem A.2. Here we take h=0h=0., then

x(W2k,s)exp(k100s3)x(W2k,2s)forsmin{18(klogd)1/4,dh2}.\mathbb{P}_{x}(W^{2k,s})\leq\exp\left(-\frac{k}{100s^{3}}\right)\mathbb{P}_{x}(W^{2k,2s})\qquad\textrm{for}\quad s\leq\min\left\{\frac{1}{8}\left(\frac{k}{\log d}\right)^{1/4},\frac{d-h}{2}\right\}. (25)
Proof.

We show this via a small modification of the proof of Theorem 3.1. Assume s(dh)/2s\leq(d-h)/2. The key observation is that each vertex has at least dhd-h edges in GG to other vertices, so in a subgraph of size at most 2s12s-1 every vertex has at least one edge in GG leaving the subgraph. In this case, we can simply choose uSu\in S as u:=argmaxwSψS(w)u:=\arg\max_{w\in S}\psi_{S}(w) in Lemma 3.3. Therefore, considering the adjacency matrix, (15) can be improved to

λ1(AS{v})12(λ1+λ12+ψS(u)2)λ1+ψS(u)26λ12λ1+16λ12s.\lambda_{1}(A_{S\cup\{v\}})\geq\frac{1}{2}\left(\lambda_{1}+\sqrt{\lambda_{1}^{2}+\psi_{S}(u)^{2}}\right)\geq\lambda_{1}+\frac{\psi_{S}(u)^{2}}{6\lambda_{1}^{2}}\geq\lambda_{1}+\frac{1}{6\lambda_{1}^{2}s}.

Therefore, after adding ss vertices to SS according to the process of Lemma 3.3, we find a set TΓx2sT\in\Gamma_{x}^{2s} satisfying

λ1(AT)λ1+16λ12i=1s1s+i1λ1+log26λ12λ1(1+110λ13).\lambda_{1}(A_{T})\geq\lambda_{1}+\frac{1}{6\lambda_{1}^{2}}\sum_{i=1}^{s}\frac{1}{s+i-1}\geq\lambda_{1}+\frac{\log 2}{6\lambda_{1}^{2}}\geq\lambda_{1}\left(1+\frac{1}{10\lambda_{1}^{3}}\right).

Using this improved bound, and keeping in mind that λ1(AT)2s\lambda_{1}(A_{T})\leq 2s, we can replicate the argument above to get to the following improvement over (18):

x(W2k,s)\displaystyle\mathbb{P}_{x}(W^{2k,s}) exp(2slogd+4slog(2s)+log(2s)k80s3)x(W2k,2s).\displaystyle\leq\exp\left(2s\log d+4s\log(2s)+\log(2s)-\frac{k}{80s^{3}}\right)\mathbb{P}_{x}(W^{2k,2s}).

This implies

x(W2k,s)exp(7slogdk80s3)x(W2k,2s)exp(k100s3)x(Wx2k,2s)\displaystyle\mathbb{P}_{x}(W^{2k,s})\leq\exp\left(7s\log d-\frac{k}{80s^{3}}\right)\mathbb{P}_{x}(W^{2k,2s})\leq\exp\left(-\frac{k}{100s^{3}}\right)\mathbb{P}_{x}(W_{x}^{2k,2s})

whenever

s18(klogd)1/4,s\leq\frac{1}{8}\left(\frac{k}{\log d}\right)^{1/4},

establishing (25). ∎

Theorem A.2 (Detailed Theorem 1.7).

If GG is simple and dd-regular, then

mG([(1loglogdnlogdn)λ2,λ2])={O(nlogdloglognd) when dlog1/2dαlog1/4nO(nlog1/2dloglognlog1/4n) when dlog1/2dαlog1/4nm_{G}\left([(1-\frac{\log\log_{d}n}{\log_{d}n})\lambda_{2},\lambda_{2}]\right)=\begin{cases}{O}\left(n\cdot\frac{\log d\log\log n}{d}\right)&\textrm{ when }d\log^{1/2}d\leq\alpha\log^{1/4}n\\ {O}\left(n\cdot\frac{\log^{1/2}d\log\log n}{\log^{1/4}n}\right)&\textrm{ when }d\log^{1/2}d\geq\alpha\log^{1/4}n\end{cases} (26)

for all888If dexp(logn)d\geq\exp(\sqrt{\log n}) then (26) is vacuously true. dexp(logn)d\leq\exp(\sqrt{\log n}), where α:=34/4\alpha:=\sqrt[4]{3}/4.

Proof.

The proof is the same as the proof of Theorem 1.2 in Section 4, except we choose different ss.

  1. 1.

    If dlog1/2d<αlog1/4nd\log^{1/2}d<\alpha\log^{1/4}n set

    s:=min{18(klogd)1/4,dh2}=d2s:=\min\left\{\frac{1}{8}\left(\frac{k}{\log d}\right)^{1/4},\frac{d-h}{2}\right\}=\frac{d}{2}

    with h=0h=0. Applying Theorem A.1 it is easily checked that (20) is satisfied for large enough nn, yielding a bound of

    mG([(1loglogdnlogdn)λ2,λ2])=O(nlogdloglognd).m_{G}\left([(1-\frac{\log\log_{d}n}{\log_{d}n})\lambda_{2},\lambda_{2}]\right)=O\left(n\cdot\frac{\log d\log\log n}{d}\right).
  2. 2.

    If GG is simple, dd-regular and dlog1/2dαlog1/4nd\log^{1/2}d\geq\alpha\log^{1/4}n, set

    s:=min{18(klogd)1/4,dh2}=18(lognlog2d)1/4s:=\min\left\{\frac{1}{8}\left(\frac{k}{\log d}\right)^{1/4},\frac{d-h}{2}\right\}=\frac{1}{8}\left(\frac{\log n}{\log^{2}d}\right)^{1/4}

    with h=0h=0. Then (20) is again satisfied by applying 3.1 equation (25), and we conclude that

    mG([(1loglogdnlogdn)λ2,λ2])=O(nlog1/2dloglognlog1/4n).m_{G}\left([(1-\frac{\log\log_{d}n}{\log_{d}n})\lambda_{2},\lambda_{2}]\right)=O\left(n\cdot\frac{\log^{1/2}d\log\log n}{\log^{1/4}n}\right).

Appendix B Lollipop

Here, we show that if we do not assume that our graph is regular, the average support of a uniformly chosen (from the set of all such walks) closed walk of length kk from a fixed vertex is no longer necessarily kΘ(1)k^{\Theta(1)} (as opposed to the average support of a random walk) . We take the lollipop graph, which consists of a clique of (d+1)(d+1) vertices for fixed d3d\geq 3 and a path of length nn {u1,,un}\{u_{1},\ldots,u_{n}\} attached to a vertex vv of the clique, where nkn\gg k. Here ψ:=ψ(A)\psi:=\psi(A) and λ1:=λ1(A)\lambda_{1}:=\lambda_{1}(A) are the Perron eigenvector and eigenvalue of the adjacency matrix of the graph.

Lemma B.1.

ψ(v)1/d+2\psi(v)\geq 1/\sqrt{d+2}.

Proof.

By symmetry, the value on all entries of the clique besides vv are the same. Call this value ψ(b)\psi(b). Then by the eigenvalue equation we have λ1ψ(b)=ψ(v)+(d1)ψ(b)\lambda_{1}\psi(b)=\psi(v)+(d-1)\psi(b), so as λ1d\lambda_{1}\geq d, it must be that ψ(v)ψ(b)\psi(v)\geq\psi(b).

Similarly, to satisfy the eigenvalue equation, vertices on the path must satisfy the recursive relation

λ1ψ(ui)=ψ(ui1)+ψ(ui+1)1in1λ1ψ(un)=ψ(un1)\begin{array}[]{cc}\lambda_{1}\psi(u_{i})=\psi(u_{i-1})+\psi(u_{i+1})&1\leq i\leq n-1\\ \lambda_{1}\psi(u_{n})=\psi(u_{n-1})\end{array}

where we define v=u0v=u_{0}. To satisfy this equation, we must have ψ(ui)(λ11)ψ(ui+1)\psi(u_{i})\geq(\lambda_{1}-1)\psi(u_{i+1}) for each ii, so as λ1d3\lambda_{1}\geq d\geq 3, ψ(v)i=1nψ(uk)\psi(v)\geq\sum_{i=1}^{n}\psi(u_{k}). As the Perron vector is nonnegative, ψ(v)2i=1nψ(uk)2\psi(v)^{2}\geq\sum_{i=1}^{n}\psi(u_{k})^{2}, and

(d+2)ψ(v)2ψ(v)2+dψ(b)2+i=1nψ(uk)2=1,(d+2)\psi(v)^{2}\geq\psi(v)^{2}+d\psi(b)^{2}+\sum_{i=1}^{n}\psi(u_{k})^{2}=1,

so ψ(v)1/d+2\psi(v)\geq 1/\sqrt{d+2}. ∎

Call γv2k\gamma_{v}^{2k} the number of closed walks of length 2k2k, and γv2k,+d+1\gamma_{v}^{2k,\geq\ell+d+1} as the subset of these walks with support at least +d+1\ell+d+1.

Proposition B.2.

For 2log(k)/log(λ1/2)\ell\geq 2\log(k)/\log(\lambda_{1}/2),

|γv2k,+d+1|=O(k2)|γv2k|.|\gamma_{v}^{2k,\geq\ell+d+1}|=O(k^{-2})|\gamma_{v}^{2k}|.
Proof.

For a closed walk to have support +d+1\ell+d+1, it must contain uu_{\ell}. For such walks, once the path is entered, at least 22\ell steps must be spent in the path, as the walk must reach uu_{\ell} and return. Therefore, closed walks starting at vv that reach uu_{\ell} can be categorized as follows. First, there is a closed walk from vv to vv. Then there is a closed walk from vv to vv going down the path containing uu_{\ell}. On this excursion, the walk can only go forward or backwards, and it spends at least 22\ell steps within the path. For each of these steps, there are 22 options. If we remain in the path after 22\ell steps, upper bound the number of choices until returning to vv by λ1\lambda_{1} at each step. After returning to vv, the remaining steps form another closed walk. The number of closed walks from vv of length ii is at most λ1i\lambda_{1}^{i}. Therefore the number of closed walks with an excursion to uu_{\ell} is at most

i=02kλ1i22λ12k2i=(2k+1)λ12k222.\sum_{i=0}^{2k}\lambda_{1}^{i}2^{2\ell}\lambda_{1}^{2k-2\ell-i}=(2k+1)\lambda_{1}^{2k-2\ell}2^{2\ell}.

The total number of closed walks starting at vv is at least ψ(v)2λ1n\psi(v)^{2}\lambda_{1}^{n}. Therefore the fraction of closed walks that have support at least \ell is at most

(2k+1)22λ12k2λ12k/(d+2)=(d+2)(2k+1)22λ12\frac{(2k+1)2^{2\ell}\lambda_{1}^{2k-2\ell}}{\lambda_{1}^{2k}/(d+2)}=\frac{(d+2)(2k+1)2^{2\ell}}{\lambda_{1}^{2\ell}}

so for 2(logk)/log(λ1/2)\ell\geq 2(\log k)/\log(\lambda_{1}/2), this is O(k2)O(k^{-2}).

Remark B.3.

Instead of adding a path, we can add a tree (as exhibited in Figure 1). According to the same analysis, the probability a walk reaches depth further than Θ(logk)\Theta(\log k) is small. Therefore, in Theorem 1.3 we can not get a sufficient bound on support from passing to depth, but must deal with support itself.