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

On the principal eigenvector of a graph

Yueheng Zhang The University of Chicago
(Date: Jul 29, 2021)
Abstract.

The principal ratio of a connected graph GG, γ(G)\gamma(G), is the ratio between the largest and smallest coordinates of the principal eigenvector of the adjacency matrix of GG. Over all connected graphs on nn vertices, γ(G)\gamma(G) ranges from 11 to ncnn^{cn}. Moreover, γ(G)=1\displaystyle{\gamma(G)=1} if and only if GG is regular. This indicates that γ(G)\gamma(G) can be viewed as an irregularity measure of GG, as first suggested by Tait and Tobin (El. J. Lin. Alg. 2018). We are interested in how stable this measure is. In particular, we ask how γ\gamma changes when there is a small modification to a regular graph GG. We show that this ratio is polynomially bounded if we remove an edge belonging to a cycle of bounded length in GG, while the ratio can jump from 11 to exponential if we join a pair of vertices at distance 22. We study the connection between the spectral gap of a regular graph and the stability of its principal ratio. A naive bound shows that given a constant multiplicative spectral gap and bounded degree, the ratio remains polynomially bounded if we add or delete an edge. Using results from matrix perturbation theory, we show that given an additive spectral gap greater than (2+ϵ)n(2+\epsilon)\sqrt{n}, the ratio stays bounded after adding or deleting an edge.

Written for the University of Chicago 2019 Math REU; mentored by Professor László Babai.

1. Introduction

It is known that the adjacency matrix of every connected graph has a simple largest eigenvalue, and this eigenvalue has an eigenvector with all-positive coordinates, called the principal eigenvector of the graph. It is natural to study how this eigenvector reflects the structure of the graph. Our goal is to obtain asymptotic information as the number of vertices approaches infinity in a family of graphs.

Cioabă and Gregory [6] defined the principal ratio of the graph GG, γ(G)=qmax/qmin\gamma(G)=q_{\max}/q_{\min}, to be the ratio between the largest and smallest coordinates of the principal eigenvector 𝐪\mathbf{q}. This ratio is 11 for regular graphs, while it can grow at factorial rate (i.e., γ(G)>ncn\gamma(G)>n^{cn} for some positive constant cc[6]. Since γ(G)1\gamma(G)\geq 1 where equality holds if and only if GG is regular, it is natural to think of γ(G)\gamma(G) as a measure of the irregularity of GG. This view was suggested by Tait and Tobin [8].

A basic observation is that, given a connected graph GG with largest eigenvalue λ1\lambda_{1} and diameter DD, the principal ratio satisfies

(1.1) γ(G)λ1D.\gamma(G)\leq\lambda_{1}^{D}.

We are interested in the stability of γ\gamma, i.e., how a slight change of GG influences γ(G)\gamma(G). In particular, given a dd-regular graph GG, we ask how γ(G)\gamma(G) changes from the constant 11 if we add or remove one edge in GG. (We call the resulting graphs G+eG+e and GeG-e, respectively.) We always assume the edge we remove will not disconnect GG (i.e., the edge ee is not a bridge), so that the principal eigenvector of GeG-e is defined.

In Section 4, we study the cases where the edge we add to or remove from a regular graph is between vertices at bounded distance. We show the following.

  • γ(G+e)\gamma(G+e) can jump to exponential in nn when the degree is bounded [Theorem 4.3]. In our example, ee connects two vertices at distance 22 in GG. We make use of Chebyshev polynomials in proving this result. We recall some properties of Chebyshev polynomials and show their connection with the principal eigenvectors of certain graphs in Sec. 3.2.
    It is easy to infer from (1.1) that boundedness of the degree is necessary here (see Cor. 3.4 and Fact 3.5).

  • If we remove an edge belonging to a cycle of bounded length in GG, then γ(Ge)\gamma(G-e) is always polynomially bounded regardless of the degree [Theorem 4.11].

We also study the relevance of the spectral gap to the stability of γ(G)\gamma(G) for regular graphs. In Section 5, based on (1.1), we note the following.

  • γ(G±e)\gamma(G\pm e) is always polynomially bounded in nn when GG is a bounded-degree expander graph, i.e., when the degree is bounded and the spectral gap of GG is bounded away from zero [Observation 5.3].

In Section 6.2, we put this problem in the more general context of perturbations of matrices. By adapting theorems and proofs from Stewart and Sun’s book [3] to our special case, we show the following.

  • If the additive spectral gap is greater than (2+ϵ)n(2+\epsilon)\sqrt{n}, then γ(G±e)\gamma(G\pm e) is bounded [Theorem 6.1].

This result does not follow from (1.1). Indeed, in Section 6.1 we construct graphs with degree of order n(2+t)/3n^{(2+t)/3} and additive spectral gap of order ntn^{t}, having diameter of order n(1t)/3n^{(1-t)/3}, for any constant 0<t<10<t<1. Similar applications of matrix perturbation theory in link analysis for networks can be found in [5].

The main technical results of this paper are the first and the fourth bullet points.

2. General preliminaries

2.1. Graphs

By a graph we mean what is often called a simple graph (undirected graph with no self-loops and no parallel edges). GG will always denote a connected graph with nn vertices. We denote by V(G)V(G) and E(G)E(G) the set of vertices and edges of GG, respectively. We usually identify the set of vertices with the set [n]={1,2,,n}[n]=\{1,2,\dots,n\}. We write iGji\sim_{G}j if vertices i,ji,j are adjacent in GG. We denote by NG(j)N_{G}(j) the set of neighbors of jj in GG. We use degG(j)\deg_{G}(j) to denote the degree of vertex jj in GG. We write G¯\overline{G} for the complement of GG. We let distG(i,j)\operatorname{dist}_{G}(i,j) denote the distance between vertices ii and jj in GG, and D(G):=maxi,jdist(i,j)D(G):=\max_{i,j}\operatorname{dist}(i,j) the diameter of GG.

Let CnC_{n} denote the cycle with nn vertices. Let PrP_{r} denote the path with rr vertices; it has r1r-1 edges. Let KsK_{s} denote the clique with ss vertices; it has (s2)\binom{s}{2} edges. Following the notation used in previous papers on this subject, we use PrKsP_{r}\cdot K_{s} to denote the graph obtained by merging the vertex at one end of PrP_{r} with one vertex in KsK_{s}. So PrKsP_{r}\cdot K_{s} has n=r+s1n=r+s-1 vertices, r1+(s2)r-1+\binom{s}{2} edges, and diameter rr. This has been called a kite graph or a lollipop graph. We will call it a kite graph.

2.2. Matrices, spectra

Orthonormality in n\mathbb{R}^{n} refers to the standard dot product. Given a matrix AA, we write aija_{ij} for the entry on the ii-th row and in the jj-th column of AA, and we write A=(aij)A=(a_{ij}).

We use Mn()M_{n}(\mathbb{R}) to denote the set of n×nn\times n real matrices. We write 𝐣n\mathbf{j}_{n} for the all-ones vector in n\mathbb{R}^{n}, and JnJ_{n} for the n×nn\times n all-ones matrix. We omit the dimension when it is clear from context.

For an m×nm\times n real matrix MM, we write M\|M\| for the operator norm of MM induced by 2\ell^{2} vector norm (\|\cdot\|),

M=sup𝐱n,𝐱𝟎M𝐱𝐱.\|M\|=\sup_{\mathbf{x}\in\mathbb{R}^{n},\;\mathbf{x}\neq\mathbf{0}}\frac{\|M\mathbf{x}\|}{\|\mathbf{x}\|}.

In the rest of Sec. 2.2, A=(aij)A=(a_{ij}) will always denote a real symmetric n×nn\times n matrix with eigenvalues λ1λ2λn\lambda_{1}\geq\lambda_{2}\geq\cdots\geq\lambda_{n}.

Definition 2.1.

A multiset based on a set AA is a function from AA to positive integers, denoted by {{am(a)aA}}\displaystyle{\{\{a^{m(a)}\mid a\in A\}\}}. The value m(a)m(a) is called the multiplicity of the element aa. For example, {{a3,b,c2}}\{\{a^{3},b,c^{2}\}\}. We also expand this notation to {{a,a,a,b,c,c}}\{\{a,a,a,b,c,c\}\}. The order in which the elements are listed does not matter. So, for instance,

{{a3,b,c2}}={{c2,b,a3}}={{a,a,a,b,c,c}}={{a,b,c,a,a,c}}.\{\{a^{3},b,c^{2}\}\}=\{\{c^{2},b,a^{3}\}\}=\{\{a,a,a,b,c,c\}\}=\{\{a,b,c,a,a,c\}\}.
Definition 2.2.

The spectrum of an n×nn\times n matrix MM is the multiset of its eigenvalues. We denote it by spec(M)\operatorname{spec}(M).

Fact 2.3.

The spectrum of the cycle CrC_{r} is

.
{{2cos(2πjr)|j=0,1,,r1}}.\left\{\left\{2\cos\left(\frac{2\pi j}{r}\right)\bigg{|}\>j=0,1,\dots,r-1\right\}\right\}.\qed
Notation 2.4.

Let B=(bij)B=(b_{ij}) be an m×nm\times n matrix and let MM be a p×qp\times q matrix. The Kronecker product of BB and MM, BMB\otimes M, is the mp×nqmp\times nq matrix

(b11Mb12Mb1nMb21Mb22Mb2nMbm1Mbm2MbmnM).\begin{pmatrix}b_{11}M&b_{12}M&\cdots&b_{1n}M\\ b_{21}M&b_{22}M&\cdots&b_{2n}M\\ \vdots&\vdots&\vdots&\vdots\\ b_{m1}M&b_{m2}M&\cdots&b_{mn}M\end{pmatrix}.
Fact 2.5.

Let BMn()B\in M_{n}(\mathbb{R}) and MMm()M\in M_{m}(\mathbb{R}). Let the eigenvalues of BB be λ1λn\displaystyle{\lambda_{1}\geq\cdots\geq\lambda_{n}} and let the eigenvalues of MM be μ1μm\displaystyle{\mu_{1}\geq\cdots\geq\mu_{m}}. Then

.
spec(BM)={{λiμji[n],j[m]}}.\operatorname{spec}(B\otimes M)=\{\{\lambda_{i}\mu_{j}\mid i\in[n],\>j\in[m]\}\}.\qed

2.3. The adjacency operator, principal eigenvector and eigenvalue

We write AGA_{G} to denote the adjacency matrix of a graph GG. We note that AGA_{G} is a real symmetric matrix, so its eigenvalues are real. We refer to the eigenvalues and eigenvectors of AGA_{G} as the eigenvalues and eigenvectors of the graph GG, respectively. We write λ1(G)λ2(G)λn(G)\lambda_{1}(G)\geq\lambda_{2}(G)\geq\cdots\geq\lambda_{n}(G) to denote the eigenvalues of GG.

Fact 2.6.

For every connected graph GG, the largest eigenvalue λ1(G)\lambda_{1}(G) is simple, and it has a unique-up-to-scaling all-positive eigenvector (the principal eigenvector). ∎

We write 𝐪(G)\mathbf{q}(G) for the principal eigenvector of GG scaled to have 2\ell^{2} norm 11. Let qi(G)q_{i}(G) denote the coordinate corresponding to vertex ii in 𝐪(G)\mathbf{q}(G). We write qmax(G)q_{\max}(G) and qmin(G)q_{\min}(G) for the maximum and minimum coordinates of 𝐪(G)\mathbf{q}(G), and vmax(G)v_{\max}(G) and vmin(G)v_{\min}(G) for corresponding vertices. Recall that the principal ratio of GG is defined as

(2.1) γ(G):=qmax(G)qmin(G).\gamma(G):=\frac{q_{\max}(G)}{q_{\min}(G)}.

In the rest of Sec. 2.3, we fix the graph GG and omit it from notations.

Observation 2.7.

Given 𝐲=(y1,,yn)\mathbf{y}=(y_{1},\dots,y_{n}), the ii-th coordinate of A𝐲A\mathbf{y} is given by

.
(A𝐲)i=j:ijyj.(A\mathbf{y})_{i}=\sum_{j:i\sim j}y_{j}.\qed
Corollary 2.8.

𝐲=(y1,,yn)\mathbf{y}=(y_{1},\dots,y_{n}) is an eigenvector of AA to eigenvalue ρ\rho if and only if

.
ρyi=j:jiyj.\rho y_{i}=\sum_{j:j\sim i}y_{j}.\qed

Recall that 𝐣\mathbf{j} denotes the all-ones vector.

Corollary 2.9.

𝐣\mathbf{j} is an eigenvector of AA if and only if GG is regular. ∎

Fact 2.10.

If GG is connected and 𝐲\mathbf{y} is an eigenvector to eigenvalue λ\lambda with all-positive coordinates, then λ\lambda is the largest eigenvalue of AA.

Proof.

Since λ1\lambda_{1} is simple, any eigenvector 𝐲\mathbf{y} to an eigenvalue other than λ1\lambda_{1} must be orthogonal to the principal eigenvector. ∎

Fact 2.11.

For a connected dd-regular graph GG, λ1=d\lambda_{1}=d. ∎

Let Δ\Delta denote max1indeg(i)\max_{1\leq i\leq n}\deg(i), the maximum degree of GG.

Fact 2.12.

For every graph GG, we have λ1Δ\lambda_{1}\leq\Delta. For connected graphs, equality holds if and only if GG is regular. ∎

We denote the quadratic mean of the degrees of vertices by

(2.2) degqavg:=i=1ndeg(i)2n.\deg_{\operatorname{qavg}}:=\sqrt{\frac{\sum_{i=1}^{n}\deg(i)^{2}}{n}}.
Fact 2.13.

For every graph GG, we have λ1degqavg\lambda_{1}\geq\deg_{\operatorname{qavg}}.

Proof.

Immediate using Rayleigh’s principle applied to A2A^{2} and the 𝐣n\mathbf{j}_{n} vector. ∎

2.4. Spectral gaps

For a graph GG, we call the quantity λ1(G)λ2(G)\lambda_{1}(G)-\lambda_{2}(G) the additive spectral gap of GG, and the quantity (λ1(G)λ2(G))/λ1(G)(\lambda_{1}(G)-\lambda_{2}(G))/\lambda_{1}(G) the multiplicative spectral gap of GG.

We write LGL_{G} for the Laplacian of GG, defined as the n×nn\times n matrix

LG=diag(deg(1),deg(2),,deg(n))AG.L_{G}=\operatorname{diag}(\deg(1),\deg(2),\dots,\deg(n))-A_{G}.

LGL_{G} is positive semidefinite; 𝐣\mathbf{j} is an eigenvector of LGL_{G} to eigenvalue zero. We write δ(G)\delta(G) for the smallest eigenvalue of LGL_{G} restricted to the orthogonal complement of 𝐣\mathbf{j}. We note that δ(G)=0\delta(G)=0 if and only if GG is disconnected. δ(G)\delta(G) is the algebraic connectivity of the graph GG, as first defined by Fiedler [1].

For a graph GG, let fGf_{G} be the characteristic polynomial of the adjacency matrix, and gGg_{G} the characteristic polynomial of the Laplacian. When GG is a dd-regular graph, we have

(2.3) fG(t)=gG(dt).f_{G}(t)=g_{G}(d-t).

It follows that

(2.4) δ(G)=dλ2(G)\delta(G)=d-\lambda_{2}(G)

is the additive spectral gap of GG, and

(2.5) δ(G)d=1λ2(G)d\frac{\delta(G)}{d}=1-\frac{\lambda_{2}(G)}{d}

is the multiplicative spectral gap of GG. We shall discuss spectral gaps for regular graphs only.

In all notation, we omit the graph GG when it is clear from context.

2.5. Rates of growth

By a family of graphs, we mean an infinite set of non-isomorphic finite graphs.

Let 𝒢={G1,G2,}\mathcal{G}=\{G_{1},G_{2},\dots\} be a family of graphs, and let nin_{i} be the number of vertices of GiG_{i}. We say γ(𝒢)\gamma(\mathcal{G}) is polynomially bounded if γ(Gi)<nic\displaystyle{\gamma(G_{i})<n_{i}^{c}} for some constant cc and all sufficiently large ii. We say γ(𝒢)\gamma(\mathcal{G}) is exponential if γ(Gi)ani\displaystyle{\gamma(G_{i})\geq a^{n_{i}}} for some constant a>1a>1 and all sufficiently large ii. We say γ(𝒢)\gamma(\mathcal{G}) has factorial growth if γ(Gi)nicni\displaystyle{\gamma(G_{i})\geq n_{i}^{cn_{i}}} for some positive constant cc and all sufficiently large ii.

2.6. Subgraphs, graph products

Fact 2.14.

If HH is a proper subgraph of a connected graph GG, i.e., HH is obtained from GG by deleting at least one edge and possibly some vertices, then λ1(G)>λ1(H)\lambda_{1}(G)>\lambda_{1}(H).

Proof.

Immediate using Rayleigh’s principle. ∎

Notation 2.15.

Given a graph GG and UV(G)U\subseteq V(G), we denote the induced subgraph of GG on the set UU by G[U]G[U].

Notation 2.16.

For graphs H=(W,F)H=(W,F) and G=(V,E)G=(V,E), the Cartesian product of HH and GG, denoted by HGH\square G, is the graph with the set W×VW\times V as vertices, and (w1,v1)(w2,v2)(w_{1},v_{1})\sim(w_{2},v_{2}) if and only if w1=w2w_{1}=w_{2} and v1Gv2v_{1}\sim_{G}v_{2}, or v1=v2v_{1}=v_{2} and w1Hw2w_{1}\sim_{H}w_{2}. For each vVv\in V, we call (HG)[W×{v}](H\square G)[W\times\{v\}] the horizontal layer corresponding to vv. For each wWw\in W, we call (HG)[{w}×V](H\square G)[\{w\}\times V] the vertical layer corresponding to ww. The horizontal layers are copies of HH and the vertical layers are copies of GG.

Notation 2.17.

For graphs H=(W,F)H=(W,F) and G=(V,E)G=(V,E), the lexicographic product of HH and GG, denoted by HGH\circ G, is the graph with the set W×VW\times V as vertices, and (w1,v1)(w2,v2)(w_{1},v_{1})\sim(w_{2},v_{2}) if and only if either w1Hw2w_{1}\sim_{H}w_{2} or w1=w2w_{1}=w_{2} and v1Gv2v_{1}\sim_{G}v_{2}. For each wWw\in W we call (HG)[{w}×V](H\circ G)[\{w\}\times V] the vertical layer corresponding to ww.

Recall that JnJ_{n} denotes the n×nn\times n all-ones matrix.

Observation 2.18.

The adjacency matrix of HGH\circ G is AHJ|V(G)|+I|V(H)|AGA_{H}\otimes J_{|V(G)|}+I_{|V(H)|}\otimes A_{G}. ∎

3. Preliminary results about the principal eigenvector

As previously introduced, GG will always denote a connected graph. We write λ1(G)\lambda_{1}(G) for the largest eigenvalue of the adjacency matrix of GG. We use 𝐪(G)\mathbf{q}(G) to mean the principal eigenvector of the adjacency matrix, the all-positive eigenvector to λ1(G)\lambda_{1}(G). We assume 𝐪(G)\mathbf{q}(G) is scaled to have 2\ell^{2} norm 11 unless otherwise stated. We write γ(G)\gamma(G) for the principal ratio qmax(G)/qmin(G)q_{\max}(G)/q_{\min}(G) of GG. We omit GG from these notations when the graph is clear from context.

3.1. Observations and naive bounds on the ratio

First, we note that 𝐪\mathbf{q} reflects the symmetries of GG.

Fact 3.1.

The principal eigenvector 𝐪\mathbf{q} is constant on orbits of Aut(G)\operatorname{Aut}(G), the automorphism group of graph GG. ∎

Next we note some basic bounds on the ratio between the coordinates of 𝐪\mathbf{q}.

Observation 3.2.

For two vertices i,ji,j in GG, let dist(i,j)=k\operatorname{dist}(i,j)=k. Then

qjqiλ1k.\frac{q_{j}}{q_{i}}\leq\lambda_{1}^{k}.
Proof.

If dist(i,j)=0\operatorname{dist}(i,j)=0, then qj/qi=1=λ10q_{j}/q_{i}=1=\lambda_{1}^{0}. If dist(i,j)=1\operatorname{dist}(i,j)=1, then by Corollary 2.8 and since all qwq_{w} are positive, λ1qi=w:wiqwqj\lambda_{1}q_{i}=\sum_{w:w\sim i}q_{w}\geq q_{j}. Now, suppose k2k\geq 2 and qw/qiλ1k1\displaystyle{q_{w}/q_{i}\leq\lambda_{1}^{k-1}} for all vertices ww at distance k1k-1 from ii. We know jj is adjacent to at least one vertex ww at distance k1k-1 from ii. Then

qjqi=qjqwqwq1λ1λ1k1=λ1k.\frac{q_{j}}{q_{i}}=\frac{q_{j}}{q_{w}}\cdot\frac{q_{w}}{q_{1}}\leq\lambda_{1}\cdot\lambda_{1}^{k-1}=\lambda_{1}^{k}.\qed

Recall that DD denotes the diameter of the graph.

Corollary 3.3.

For a connected graph GG of diameter DD,

.
γλ1DΔD(n1)D.\gamma\leq\lambda_{1}^{D}\leq\Delta^{D}\leq(n-1)^{D}.\qed
Corollary 3.4.

If DD is bounded for some family 𝒢\mathcal{G} of graphs , then γ(𝒢)\gamma(\mathcal{G}) is polynomially bounded in nn. ∎

Since DD is relevant in bounding the ratio, we introduce a bound on DD for regular graphs.

Fact 3.5.

Let GG be a connected dd-regular graph. Then D3nd\displaystyle{D\leq\frac{3n}{d}}.

Proof.

Pick v0,vDv_{0},v_{D} in GG so that dist(v0,vD)=D\operatorname{dist}(v_{0},v_{D})=D. Let v0,v1,,vDv_{0},v_{1},\dots,v_{D} be a shortest path from v0v_{0} to vDv_{D}. Any vi,vjv_{i},v_{j} with |ij|3|i-j|\geq 3 cannot have any common neighbors, since otherwise the path will not be a shortest path. Thus

dD3n.d\left\lceil\frac{D}{3}\right\rceil\leq n.

Therefore D3n/dD\leq 3n/d. ∎

We know D(G+e)D(G)D(G+e)\leq D(G) for any eG¯e\in\overline{G}. The following result shows that D(G+e)D(G+e), and consequently, also D(Ge)D(G-e) (if still connected), cannot differ from DGD_{G} by more than a factor of 22.

Fact 3.6.

For a connected graph GG with D(G)=DD(G)=D and eE(G¯)e\in E(\overline{G}),

D(G+e)12D(G).D(G+e)\geq\frac{1}{2}D(G).
Proof.

(Notation: By distl(x,y)\operatorname{dist}_{l}(x,y), where ll is a path, we mean the distance between xx and yy along the path.)
We need to prove that there is a pair of vertices at distance at least D2\frac{D}{2} in G+eG+e. Let u,vV(G)u,v\in V(G) be such that distG(u,v)=D\operatorname{dist}_{G}(u,v)=D. If distG+e(u,v)=D\operatorname{dist}_{G+e}(u,v)=D, we are done. Otherwise, let pp be a shortest path between uu and vv in GG, and qq a shortest path between uu and vv in G+eG+e. Then ee must be on qq. Denote by xx the endpoint of ee which is closer to uu on qq, and by yy the other endpoint. Pick the middle vertex ww of pp with distp(u,w)=D2\operatorname{dist}_{p}(u,w)=\lceil\frac{D}{2}\rceil. If distG+e(u,w)=distG(u,w)=D2\operatorname{dist}_{G+e}(u,w)=\operatorname{dist}_{G}(u,w)=\lceil\frac{D}{2}\rceil, we are done. Otherwise, any shortest path rr in G+eG+e from uu to ww must pass through edge ee. Suppose we go along rr from uu to ww. By the optimality of qq, we may assume that rr and qq overlap from uu to yy. Now we look at the vertex ww^{\prime} adjacent to ww on pp which is closer to uu than to vv. We have

(3.1) distG(w,v)=distp(w,v)=D2+1.\operatorname{dist}_{G}(w^{\prime},v)=\operatorname{dist}_{p}(w^{\prime},v)=\left\lfloor\frac{D}{2}\right\rfloor+1.

We claim that there is a shortest path in G+eG+e from ww^{\prime} to vv that does not pass through ee. Let ss be a shortest path in G+eG+e from ww^{\prime} to vv that passes through ee. By the optimality of qq, we may assume that ss and qq overlap from xx to vv. Then by the optimality of rr,

dists(x,w)+distG+e(w,w)distr(x,w)=distG+e(x,y)+distr(y,w),\operatorname{dist}_{s}(x,w^{\prime})+\operatorname{dist}_{G+e}(w^{\prime},w)\geq\operatorname{dist}_{r}(x,w)=\operatorname{dist}_{G+e}(x,y)+\operatorname{dist}_{r}(y,w),

that is,

dists(x,w)distr(y,w).\operatorname{dist}_{s}(x,w^{\prime})\geq\operatorname{dist}_{r}(y,w).

Therefore

dists(w,y)=dists(w,x)+distG+e(x,y)distG+e(w,w)+distr(w,y).\operatorname{dist}_{s}(w^{\prime},y)=\operatorname{dist}_{s}(w^{\prime},x)+\operatorname{dist}_{G+e}(x,y)\geq\operatorname{dist}_{G+e}(w^{\prime},w)+\operatorname{dist}_{r}(w,y).

Thus ss is equivalent to a path that does not pass through ee in G+eG+e. As a result, a shortest path between ww^{\prime} and vv in G+eG+e is also available in GG. Then by (3.1),

distG+e(w,v)=distG(w,v)=D2+1.\operatorname{dist}_{G+e}(w^{\prime},v)=\operatorname{dist}_{G}(w^{\prime},v)=\left\lfloor\frac{D}{2}\right\rfloor+1.\qed

3.2. Chebyshev polynomials and principal eigenvectors

The Chebyshev polynomials of the first kind, TnT_{n}, can be characterized by the recurrence

(3.2) Tn+1(t)=2tTn(t)Tn1(t),T_{n+1}(t)=2t\cdot T_{n}(t)-T_{n-1}(t),

with initial values T0(t)=1T_{0}(t)=1 and T1(t)=tT_{1}(t)=t.
The Chebyshev polynomials of the second kind, UnU_{n}, can be characterized by the same recurrence

(3.3) Un+1(t)=2tUn(t)Un1(t),U_{n+1}(t)=2t\cdot U_{n}(t)-U_{n-1}(t),

with initial values U0(t)=1U_{0}(t)=1 and U1(t)=2tU_{1}(t)=2t.

Fact 3.7.

When |t|1|t|\geq 1, the explicit formula for TnT_{n} is

(3.4) Tn(t)=12((tt21)n+(t+t21)n).T_{n}(t)=\frac{1}{2}\Big{(}\Big{(}t-\sqrt{t^{2}-1}\Big{)}^{n}+\Big{(}t+\sqrt{t^{2}-1}\Big{)}^{n}\Big{)}.

and the explicit formula for UnU_{n} is

Proof.
(3.5) Un(t)=(t+t21)n+1(tt21)n+12t21.U_{n}(t)=\frac{\Big{(}t+\sqrt{t^{2}-1}\Big{)}^{n+1}-\Big{(}t-\sqrt{t^{2}-1}\Big{)}^{n+1}}{2\sqrt{t^{2}-1}}.\qed
Fact 3.8.

When x>1x>1, both Tn(x)T_{n}(x) and Un(x)U_{n}(x) are strictly increasing. ∎

Definition 3.9.

A pendant path of length kk in GG consists of kk vertices such that the induced subgraph on them is a path; moreover, one vertex has degree 11 in GG and k2k-2 vertices have degree 22 in GG. For example, in the graph PrKsP_{r}\cdot K_{s}, there is a pendant path of length rr.

Observation 3.10.

Let 1,2,,k1,2,\dots,k be a pendant path in GG where consecutive vertices are adjacent and deg(1)=1\deg(1)=1. Then for 1jk1\leq j\leq k,

qjq1=Uj1(λ12).\frac{q_{j}}{q_{1}}=U_{j-1}\left(\frac{\lambda_{1}}{2}\right).
Proof.

By Corollary 2.8, λ1q1=q2\lambda_{1}q_{1}=q_{2} and qj+1=λ1qjqj1q_{j+1}=\lambda_{1}q_{j}-q_{j-1} for 1jn11\leq j\leq n-1. Therefore qj/q1q_{j}/q_{1} satisfies the initial values and recurrence relation of Uj1(λ1/2)U_{j-1}(\lambda_{1}/2). ∎

In fact, this occurence was observed by Tait and Tobin [8] when they proved that the maximum principal ratio over all graphs of nn vertices is attained by a kite graph.

4. Stability of the ratio: adding or removing edges in bounded distance

In this section and the subsequent sections, we are interested in the stability of the ratio with respect to a small change in the graph.

Let GG be a dd-regular graph, so γ(G)=1\gamma(G)=1. As introduced before, we use G+eG+e to denote the graph obtained by adding an edge eE(G¯)e\in E(\overline{G}) to GG, and GeG-e to denote the graph obtained by deleting an edge eE(G)e\in E(G) from GG. We always assume GeG-e is still connected. We are interested in the possible asymptotic behaviors of γ(G+e)\gamma(G+e) and γ(Ge)\gamma(G-e).

We first make two simple observations. For the first one, we do not require GG to be regular.

Observation 4.1.

If the diameter D(G)D(G) is bounded, then γ(G±e)\gamma(G\pm e) is polynomially bounded in nn.

Proof.

Fact 3.6 shows that D(G±e)D(G\pm e) is also bounded, and the statement follows from Cororllary 3.4. ∎

Observation 4.2.

Let GG be a dd-regular graph. If dd is linear in nn, i.e., d/nd/n is bounded away from 0, then γ(G±e)\gamma(G\pm e) is polynomially bounded in nn.

Proof.

Fact 3.5 shows that D(G)D(G) is bounded, and the statement follows from Observation 4.1. ∎

4.1. Adding an edge

We show that if we add an edge ee between two vertices at distance 22 to a connected regular graph GG of bounded degree, then γ(G+e)\gamma(G+e) can be exponentially large.

Theorem 4.3.

For any d2d\geq 2 and n>18d3n>18d^{3} such that n/(d1)n/(d-1) is an odd integer, there exists a dd-regular graph Gn,dG_{n,d} with nn vertices and an edge en,dE(Gn,d¯)e_{n,d}\in E(\overline{G_{n,d}}) between two vertices at distance 2 in Gn,dG_{n,d}, such that

(4.1) γ(Gn,d+en,d)>en/(18d3).\gamma(G_{n,d}+e_{n,d})>\mathrm{e}^{n/(18d^{3})}.

In particular, when dd is bounded, γ(Gn,d+en,d)\gamma(G_{n,d}+e_{n,d}) is exponential in nn.

The proof of this theorem will be based on a series of constructions. The graphs produced by Construction 4.6 and Construction 4.7 are the pair of graphs that are used in the proof.

Construction 4.4 (Ringr,d,G2\operatorname{Ring}_{r,d,G_{2}}).

Let r0r\geq 0 be a parameter. We label the vertices in P2r+1P_{2r+1} from one end to the other end as prp_{-r}, pr+1p_{-r+1}, …, p1p_{-1}, p0p_{0}, p1p_{1}, …, pr1p_{r-1}, prp_{r}. Let G1=P2r+1Kd1\displaystyle{G_{1}=P_{2r+1}\square K_{d-1}}. We label the vertical layer corresponding to pip_{i} as LiL_{i}. Let G2G_{2}, which we will call a “gadget,” be a connected graph with two vertices v1v_{1}, v2v_{2} of degree 11 and all other vertices of degree dd, and an automorphism that switches v1v_{1} and v2v_{2}. We connect v1v_{1} with every vertex in LrL_{-r} and connect v2v_{2} with every vertex in LrL_{r}. We call the graph obtained Ringr,d,G2\operatorname{Ring}_{r,d,G_{2}}.

Observation 4.5.

Let XAut(G1)X\leq\operatorname{Aut}(G_{1}) be the subgroup of automorphisms of G1G_{1} that fixes the vertical layers and permutes the horizontal layers. The orbits of XX are the vertical layers. XX is isomorphic to the symmetric group Sd1S_{d-1}. This group extends to GG. The automorphism of G1G_{1} that switches LjL_{j} and LjL_{-j} for 0jr0\leq j\leq r also extends to GG. Therefore the coordinates of the principal eigenvector of Ringr,d,G2\operatorname{Ring}_{r,d,G_{2}} corresponding to all vertices in LjLjL_{j}\cup L_{-j} are the same. We denote this value by aja_{j}, for 0jr0\leq j\leq r.

Construction 4.6.

[Ringr,d\operatorname{Ring}_{r,d}] Now we specify the gadget G2G_{2}. We take two copies of Kd+1K_{d+1} and call them H1H_{1}, H2H_{2}. We label the vertices in H1H_{1}, H2H_{2} from 11 to d+1d+1. We remove the edge {1,2}\{1,2\} from H2H_{2}, the edge {3,4}\{3,4\} from H1H_{1}, and add the edges {1,3}\{1,3\} and {2,4}\{2,4\}. We find two vertices u1u_{1}, u2u_{2} in V(H1)V(H_{1}) such that u2𝒪(u1)u_{2}\in\mathcal{O}(u_{1}) in H1{3,4}H_{1}-\{3,4\}. We remove the edge {u1,u2}\{u_{1},u_{2}\}. Finally, we attach a dangling vertex w1w_{1} to u1u_{1}, and a dangling vertex w2w_{2} to u2u_{2}. w1w_{1} and w2w_{2} are the vertices that connect with LrL_{-r} and LrL_{r}, respectively. For this specific G2G_{2}, Ring(r,d,G2)\operatorname{Ring}(r,d,G_{2}) is a dd-regular graph on n=(2r+1)(d1)+2+2(d+1)=2rd2r+3d+3n=(2r+1)(d-1)+2+2(d+1)=2rd-2r+3d+3 vertices. We write Ring(r,d)\operatorname{Ring}(r,d) for this graph.

Construction 4.7.

[Ringr,d+e\operatorname{Ring}_{r,d}+e] We add the edge e={3,4}E(Ringr,d¯)e=\{3,4\}\in E(\overline{\operatorname{Ring}_{r,d}}) to Ringr,d\operatorname{Ring}_{r,d}.

[Uncaptioned image]
Proposition 4.8.

For 0jr0\leq j\leq r,

aja0=Tj(λ1(Ringr,d+e)d+22)\frac{a_{j}}{a_{0}}=T_{j}\left(\frac{\lambda_{1}(\operatorname{Ring}_{r,d}+e)-d+2}{2}\right)

where aja_{j} is as defined in Observation 4.5.

Proof.

By Corollary 2.8,

λ1a0=(d2)a0+2a1\lambda_{1}a_{0}=(d-2)a_{0}+2a_{1}

and

λ1aj=(d2)aj+aj1+aj+1\lambda_{1}a_{j}=(d-2)a_{j}+a_{j-1}+a_{j+1}

for 1jn11\leq j\leq n-1. Therefore aj/a0a_{j}/a_{0} satisfies the initial values and recurrence relation of Tj((λ1d+2)/2)T_{j}((\lambda_{1}-d+2)/2) according to (3.2). ∎

Lemma 4.9.

λ1(Ringr,d+e)>d+c(d)\lambda_{1}(\operatorname{Ring}_{r,d}+e)>d+c(d), where c(d):=2/(3d(d+1))c(d):=2/(3d(d+1)).

Proof.

Let HH be the induced subgraph of Ringr,d\operatorname{Ring}_{r,d} on V(H1)V(H2)V(H_{1})\cup V(H_{2}). Then

degH+e(1)=degH+e(2)=d+1,degH+e(u1)=degH+e(u2)=d1,\deg_{H+e}(1)=\deg_{H+e}(2)=d+1,\quad\deg_{H+e}(u_{1})=\deg_{H+e}(u_{2})=d-1,

while the rest of the vertices in H+eH+e are of degree dd. Then by Fact 2.14 and Fact 2.13,

λ1(Ringr,d+e)>λ1(H+e)degqavg(H+e)=(2d+2)d2+42d+2=d1+2d2(d+1).\lambda_{1}(\operatorname{Ring}_{r,d}+e)>\lambda_{1}(H+e)\geq\deg_{\operatorname{qavg}}(H+e)=\sqrt{\frac{(2d+2)d^{2}+4}{2d+2}}=d\sqrt{1+\frac{2}{d^{2}(d+1)}}.

Since 1+x>1+13x\displaystyle{\sqrt{1+x}>1+\frac{1}{3}x} when 0<x<30<x<3, we have λ1(G+e)>d+c(d)\displaystyle{\lambda_{1}(G+e)>d+c(d)}, as claimed. ∎

Proof of Theorem 4.3.

By Lemma 4.9,

λ1(Ringr,d+e)d+22>1+13d(d+1).\frac{\lambda_{1}(\operatorname{Ring}_{r,d}+e)-d+2}{2}>1+\frac{1}{3d(d+1)}.

By Proposition 4.8, Fact 3.8, and Fact 3.7,

γ(Ringr,d)ara0=Tr(λ1(Ringr,d+e)d+22)>Tj(1+13d(d+1))>12(1+13d(d+1))r.\gamma(\operatorname{Ring}_{r,d})\geq\frac{a_{r}}{a_{0}}=T_{r}\left(\frac{\lambda_{1}(\operatorname{Ring}_{r,d}+e)-d+2}{2}\right)>T_{j}\left(1+\frac{1}{3d(d+1)}\right)>\frac{1}{2}\left(1+\frac{1}{3d(d+1)}\right)^{r}.

Plugging in r=n3d32d2\displaystyle{r=\frac{n-3d-3}{2d-2}} and assuming n>18d3n>18d^{3}, we infer by a simple calculation that

γ(Ringr,d)>en/(18d3).\gamma(\operatorname{Ring}_{r,d})>\mathrm{e}^{n/(18d^{3})}.\qed

4.2. Removing an edge

Let e={1,2}e=\{1,2\}.

We show that in the case of removing the edge ee when distGe(1,2)\operatorname{dist}_{G-e}(1,2) is bounded, γ(Ge)\gamma(G-e) is polynomially bounded for all DD. We make use of the following theorem.

Theorem 4.10 (Cioabă, Gregory, Nikiforov[7]).

If GG is a connected nonregular graph with nn vertices, diameter DD, and maximum degree Δ\Delta, then

Δλ1(G)1n(D+1).\Delta-\lambda_{1}(G)\geq\frac{1}{n(D+1)}.
Theorem 4.11.

For a connected dd-regular graph GG and an edge e={1,2}E(G)e=\{1,2\}\in E(G), if distGe(1,2)<c\operatorname{dist}_{G-e}(1,2)<c where cc is some constant, then γ(Ge)\gamma(G-e) is polynomially bounded in nn.

Lemma 4.12.

qmin(Ge)q_{\min}(G-e) is either q1q_{1} or q2q_{2}.

Proof.

If qminq_{\min} corresponds to some vertex jj with degree dd, then the average of the coordinates corresponding to the neighbors of jj would be

λ1(Ge)qjd<λ1(Ge)qjλ1(Ge)=qj,\frac{\lambda_{1}(G-e)q_{j}}{d}<\frac{\lambda_{1}(G-e)q_{j}}{\lambda_{1}(G-e)}=q_{j},

which is a contradiction. ∎

Proof of Theorem 4.11.

Without loss of generality suppose q1=qminq_{1}=q_{\min}. Then summing nn equations of the form λ1(Ge)qi=j:jiqj\lambda_{1}(G-e)q_{i}=\sum_{j:j\sim i}q_{j}, we have

λ1(Ge)i=1nqi=(d1)q1+(d1)q2+dq3++dqn=d(i=1nqi)q1q2.\lambda_{1}(G-e)\sum_{i=1}^{n}q_{i}=(d-1)q_{1}+(d-1)q_{2}+dq_{3}+\cdots+dq_{n}=d\left(\sum_{i=1}^{n}q_{i}\right)-q_{1}-q_{2}.

Therefore by Theorem 4.10,

q1+q2i=1nqi=dλ1(Ge)1n(D(Ge)+1)1n2.\frac{q_{1}+q_{2}}{\sum_{i=1}^{n}q_{i}}=d-\lambda_{1}(G-e)\geq\frac{1}{n(D(G-e)+1)}\geq\frac{1}{n^{2}}.

By Observation 3.2, q2λ1(Ge)cq1q_{2}\leq\lambda_{1}(G-e)^{c}q_{1}. Therefore

γ(Ge)<i=1nqiq1n2(1+q2q1)n2(1+λ1(Ge)c)<n2(1+dc).\gamma(G-e)<\frac{\sum_{i=1}^{n}q_{i}}{q_{1}}\leq n^{2}\left(1+\frac{q_{2}}{q_{1}}\right)\leq n^{2}(1+\lambda_{1}(G-e)^{c})<n^{2}(1+d^{c}).\qed

5. Stability of the ratio: multiplicative spectral gap

We use a known bound on the diameter of a graph in terms of the spectral gap to show that γ(G±e)\gamma(G\pm e) is polynomially bounded for bounded-degree expanders.

Definition 5.1.

For 0<ϵ<10<\epsilon<1, a regular graph of degree dd is an ϵ\epsilon-expander if λ2(1ϵ)d\displaystyle{\lambda_{2}\leq(1-\epsilon)d}, i.e., δϵd\displaystyle{\delta\geq\epsilon d}.

We note that this definition implies that an expander graph is connected.

Theorem 5.2 (N. Alon, V. D. Milman[2]).

Let GG be a connected graph on nn vertices with maximum degree Δ\Delta and let δ\delta denote the smallest positive eigenvalue of the Laplacian matrix. Then

D(G)22Δδlog2n.D(G)\leq 2\left\lfloor\sqrt{\frac{2\Delta}{\delta}}\log_{2}n\right\rfloor.
Corollary 5.3.

For dd-regular ϵ\epsilon-expander graphs GG, we have

(5.1) γ(G+e)n42/ϵlog2(d+1).\gamma(G+e)\leq n^{4\sqrt{2/\epsilon}\log_{2}(d+1)}.

In particcular, for expander graphs GG with bounded degree, γ(G±e)\gamma(G\pm e) is polynomially bounded in nn.

Proof.

By definition,

Δ(G)δ(G)=dδ(G)1ϵ.\frac{\Delta(G)}{\delta(G)}=\frac{d}{\delta(G)}\leq\frac{1}{\epsilon}.

Then

D(G)22ϵlog2n.D(G)\leq 2\left\lfloor\sqrt{\frac{2}{\epsilon}}\log_{2}n\right\rfloor.

By Fact 3.6,

D(G±e)42ϵlog2n.D(G\pm e)\leq 4\left\lfloor\sqrt{\frac{2}{\epsilon}}\log_{2}n\right\rfloor.

Since γΔD\gamma\leq\Delta^{D} (Corollary 3.3),

γ(G±e)\displaystyle\gamma(G\pm e) (d+1)42/ϵlog2n\displaystyle\leq(d+1)^{4\left\lfloor\sqrt{2/\epsilon}\log_{2}n\right\rfloor}
(d+1)42/ϵlog2n\displaystyle\leq(d+1)^{4\sqrt{2/\epsilon}\log_{2}n}
=242/ϵlog2nlog2(d+1)\displaystyle=2^{4\sqrt{2/\epsilon}\log_{2}n\log_{2}(d+1)}
=n42/ϵlog2(d+1).\displaystyle=n^{4\sqrt{2/\epsilon}\log_{2}(d+1)}.\qed

6. Stability of the ratio: additive spectral gap

We show that a large ((2+ϵ)n(2+\epsilon)\sqrt{n}) additive spectral gap implies that γ(G±e)\gamma(G\pm e) is bounded. Specifically, we prove the following.

Theorem 6.1.

Let GG be a connected dd-regular graph. If the additive spectral gap δ=dλ2\delta=d-\lambda_{2} of GG satisfies δ>2cn+2\displaystyle{\delta>\frac{2}{c}\sqrt{n}}+2 for some value 0<c<10<c<1, then

γ(G±e)<1+c1c.\gamma(G\pm e)<\frac{1+c}{1-c}.

To motivate this result, we first point out that graphs with such an additive spectral gap are not necessarily expanders. Indeed, when dd is greater than Θ(n)\Theta(\sqrt{n}), the multiplicative spectral gap of graphs with a Θ(n)\Theta(\sqrt{n}) additive spectral graph will go to zero. Moreover, the diameter of graphs with such an additive spectral gap can still grow quite fast (polynomially in nn), approaching the upper bound derived from Theorem 5.2.

6.1. Existence of graphs with large additive spectral gap and large diameter

Proposition 6.2.

For a regular graph with additive spectral gap δ\delta, the diameter DD is O((n/δ)1/3(logn)2/3)O((n/\delta)^{1/3}(\log n)^{2/3}).

Proof.

Let the regular graph have degree dd. From Theorem 5.2, we know that

(6.1) D2cdδ(logn)2D^{2}\leq c\frac{d}{\delta}(\log n)^{2}

where cc is some constant. By Fact 3.5, we also have

(6.2) D3nd.D\leq\frac{3n}{d}.

Multiplying (6.1) and (6.2), we have

(6.3) D(3c)1/3(nδ)1/3(logn)2/3.D\leq(3c)^{1/3}\left(\frac{n}{\delta}\right)^{1/3}(\log n)^{2/3}.\qed
Corollary 6.3.

For a regular graph with an Ω(n)\Omega(\sqrt{n}) additive spectral gap, the diameter is O(n1/6(logn)2/3)O(n^{1/6}(\log n)^{2/3}).

We show that this bound is nearly tight.

Proposition 6.4.

There are connected regular graphs with diameter (1/2)n1/6(1/2)n^{1/6} and an additive spectral gap of cn1/2cn^{1/2} where c=2π2(1+O(n1/3))c=2\pi^{2}(1+O(n^{-1/3})).

We prove a more general statement.

Proposition 6.5.

For any constant 0<t<10<t<1, there are connected regular graphs with diameter (1/2)n(1t)/3(1/2)n^{(1-t)/3} and an additive spectral gap of cntcn^{t} where
c=2π2(1+O(n(2t2)/3))c=2\pi^{2}(1+O(n^{(2t-2)/3})).

Proof.

Recall the lexicographic product and its properties in Definition 2.17, Observation 2.18, and Fact 2.5. Let G:=CrK¯sG:=C_{r}\circ\overline{K}_{s} where rs=nr\cdot s=n. Then GG is a connected regular graph of degree 2s2s and diameter r/2\lfloor r/2\rfloor. The adjacency matrix of GG can be expressed as

(6.4) AG=ACrJs+Ir𝟎=ACrJs.A_{G}=A_{C_{r}}\otimes J_{s}+I_{r}\otimes\mathbf{0}=A_{C_{r}}\otimes J_{s}.

Since

(6.5) spec(Js)={{s,0s1}},\operatorname{spec}(J_{s})=\{\{s,0^{s-1}\}\},

we have

(6.6) spec(G)={{sλ1(Cr),sλ2(Cr),,sλr(Cr),0(s1)r.}}\operatorname{spec}(G)=\{\{s\lambda_{1}(C_{r}),s\lambda_{2}(C_{r}),\dots,s\lambda_{r}(C_{r}),0^{(s-1)r}.\}\}

Thus

δ(G)=s(λ1(Cr)λ2(Cr)).\delta(G)=s(\lambda_{1}(C_{r})-\lambda_{2}(C_{r})).

By Fact 2.3,

λ1(Cr)λ2(Cr)=2(1cos(2πr))=4π2r2+ϵ\lambda_{1}(C_{r})-\lambda_{2}(C_{r})=2\left(1-\cos\left(\frac{2\pi}{r}\right)\right)=\frac{4\pi^{2}}{r^{2}}+\epsilon

where |ϵ|<25π44!r4\displaystyle{|\epsilon|<\frac{2^{5}\pi^{4}}{4!r^{4}}}. We take r=n(1t)/3\displaystyle{r=n^{(1-t)/3}} and s=n(2+t)/3\displaystyle{s=n^{(2+t)/3}}. Then

δ(G)=4π2nt(1+O(1r2))=4π2nt(1+O(n(2t2)/3)).\delta(G)=4\pi^{2}n^{t}\left(1+O\left(\frac{1}{r^{2}}\right)\right)=4\pi^{2}n^{t}\left(1+O(n^{(2t-2)/3})\right).\qed

6.2. Large additive spectral gap implies bounded ratio

Now we prove Theorem 6.1 which shows that γ(G±e)\gamma(G\pm e) is bounded when GG has an additive spectral gap of (2+ϵ)n(2+\epsilon)\sqrt{n} for some ϵ>0\epsilon>0. The proof is an adaptation of the theory developed in Chapter V. 2 in Stewart and Sun’s book [3], dealing with perturbation of invariant subspaces. In addition to adapting the proofs to our special case, we fill in some details to make the proof accessible to readers not versed in perturbation theory.

Notation 6.6.

Let UMn()U\in M_{n}(\mathbb{R}) be the orthogonal matrix whose columns are eigenvectors of AA. We can write it as

(6.7) U=(𝐱Y)=(𝐱𝐲2𝐲n),U=(\mathbf{x}\;Y)=(\mathbf{x}\;\mathbf{y}_{2}\;\cdots\;\mathbf{y}_{n}),

where the columns 𝐱,𝐲2,𝐲3,,𝐲n\mathbf{x},\mathbf{y}_{2},\mathbf{y}_{3},\dots,\mathbf{y}_{n} are eigenvectors of AA corresponding to eigenvalues λ1λ2λ3λn\lambda_{1}\geq\lambda_{2}\geq\lambda_{3}\geq\dots\geq\lambda_{n}. Then λ1=d\lambda_{1}=d, and we can assume

(6.8) 𝐱=(1n,,1n)T.\mathbf{x}=\left(\frac{1}{\sqrt{n}},\dots,\frac{1}{\sqrt{n}}\right)^{T}.

In the context of adding an edge ee to GG, we label the two endpoints of ee to be 11 and 22, and let EMn()E\in M_{n}(\mathbb{R}) have E21=E12=1E_{21}=E_{12}=1 while all other coordinates are zero. In the context of deleting an edge ee from GG, we also label the two endpoints of ee to be 11 and 22, and let EMn()E\in M_{n}(\mathbb{R}) have E21=E12=1E_{21}=E_{12}=-1 while all other coordinates are zero. In both cases, A+EA+E is the adjacency matrix of the graph obtained.

6.2.1. Rotating the eigenbasis

We know 𝐱\mathbf{x} is not an eigenvector of A+EA+E. We want to know how close 𝐱\mathbf{x} is to the principal eigenvector of G±eG\pm e, in the sense that we want to find a vector 𝐯\mathbf{v} with small norm such that 𝐱~=𝐱+𝐯𝐱+𝐯\displaystyle{\widetilde{\mathbf{x}}=\frac{\mathbf{x}+\mathbf{v}}{\|\mathbf{x}+\mathbf{v}\|}} is the unit principal eigenvector of G±eG\pm e.

Proposition 6.7.

Let 𝐩\mathbf{p} be a vector in n1\mathbb{R}^{n-1}. Let U=(𝐱Y)Mn()U=(\mathbf{x}\;Y)\in M_{n}(\mathbb{R}) be orthogonal, where 𝐱\mathbf{x} is a vector in n1\mathbb{R}^{n-1} and YY is an n×(n1)n\times(n-1) matrix. Define U~Mn()\widetilde{U}\in M_{n}(\mathbb{R}) as U~=(𝐱~Y~)\displaystyle{\widetilde{U}=(\widetilde{\mathbf{x}}\;\widetilde{Y})}, where

(6.9) 𝐱~=𝐱+Y𝐩1+𝐩2andY~=(Y𝐱𝐩T)(In1+𝐩𝐩T)1/2.\widetilde{\mathbf{x}}=\frac{\mathbf{x}+Y\mathbf{p}}{\sqrt{1+\|\mathbf{p}\|^{2}}}\quad\text{and}\quad\widetilde{Y}=(Y-\mathbf{x}\mathbf{p}^{T})(I_{n-1}+\mathbf{p}\mathbf{p}^{T})^{-1/2}.

Then U~\widetilde{U} is an orthogonal matrix.

Proof.
𝐱~T𝐱~=𝐱T𝐱+𝐩TYTY𝐩1+𝐩2=1+𝐩21+𝐩2=1.\widetilde{\mathbf{x}}^{T}\widetilde{\mathbf{x}}=\frac{\mathbf{x}^{T}\mathbf{x}+\mathbf{p}^{T}Y^{T}Y\mathbf{p}}{1+\|\mathbf{p}\|^{2}}=\frac{1+\|\mathbf{p}\|^{2}}{1+\|\mathbf{p}\|^{2}}=1.
Y~TY~\displaystyle\widetilde{Y}^{T}\widetilde{Y} =(In1+𝐩𝐩T)1/2(YT𝐩𝐱T)(Y𝐱𝐩T)(In1+𝐩𝐩T)1/2\displaystyle=(I_{n-1}+\mathbf{p}\mathbf{p}^{T})^{-1/2}(Y^{T}-\mathbf{p}\mathbf{x}^{T})(Y-\mathbf{x}\mathbf{p}^{T})(I_{n-1}+\mathbf{p}\mathbf{p}^{T})^{-1/2}
=(In1+𝐩𝐩T)1/2(In1+𝐩𝐩T)(In1+𝐩𝐩T)1/2\displaystyle=(I_{n-1}+\mathbf{p}\mathbf{p}^{T})^{-1/2}(I_{n-1}+\mathbf{p}\mathbf{p}^{T})(I_{n-1}+\mathbf{p}\mathbf{p}^{T})^{-1/2}
=In1.\displaystyle=I_{n-1}.
𝐱~TY~\displaystyle\widetilde{\mathbf{x}}^{T}\widetilde{Y} =(𝐱T+𝐩TYT)(Y𝐱𝐩T)(In1+𝐩𝐩T)1/21+𝐩2\displaystyle=\frac{(\mathbf{x}^{T}+\mathbf{p}^{T}Y^{T})(Y-\mathbf{x}\mathbf{p}^{T})(I_{n-1}+\mathbf{p}\mathbf{p}^{T})^{-1/2}}{\sqrt{1+\|\mathbf{p}\|^{2}}}
=(𝐩T+𝐩T)(In1+𝐩𝐩T)1/21+𝐩2\displaystyle=\frac{(-\mathbf{p}^{T}+\mathbf{p}^{T})(I_{n-1}+\mathbf{p}\mathbf{p}^{T})^{-1/2}}{\sqrt{1+\|\mathbf{p}\|^{2}}}
=0.\displaystyle=0.

Therefore

U~TU~=(𝐱~T𝐱~𝐱~TY~Y~T𝐱~Y~TY~)=In.\widetilde{U}^{T}\widetilde{U}=\begin{pmatrix}\widetilde{\mathbf{x}}^{T}\widetilde{\mathbf{x}}&\widetilde{\mathbf{x}}^{T}\widetilde{Y}\\ \widetilde{Y}^{T}\widetilde{\mathbf{x}}&\widetilde{Y}^{T}\widetilde{Y}\end{pmatrix}=I_{n}.\qed

6.2.2. Rotating 𝐱\mathbf{x} to be an eigenvector

We want to find 𝐩\mathbf{p} such that 𝐱~\widetilde{\mathbf{x}} is the principal eigenvector of A+EA+E.

Notation 6.8.

We define

e11:=𝐱TE𝐱,e21:=YTE𝐱n1,E22:=YTEYMn1(),e_{11}:=\mathbf{x}^{T}E\mathbf{x}\in\mathbb{R},\quad e_{21}:=Y^{T}E\mathbf{x}\in\mathbb{R}^{n-1},\quad E_{22}:=Y^{T}EY\in M_{n-1}(\mathbb{R}),
andL:=YTAY=diag(λ2,,λn)Mn1().\text{and}\quad L:=Y^{T}AY=\operatorname{diag}(\lambda_{2},\dots,\lambda_{n})\in M_{n-1}(\mathbb{R}).
Observation 6.9.

Since E=1\|E\|=1 and (𝐱Y)(\mathbf{x}\>Y) is orthogonal, |e11|,e21,E221\displaystyle{|e_{11}|,\|e_{21}\|,\|E_{22}\|\leq 1}.

Proposition 6.10.

If

(6.10) ((d+e11)In1(L+E22))𝐩=e21𝐩e21T𝐩,((d+e_{11})I_{n-1}-(L+E_{22}))\mathbf{p}=e_{21}-\mathbf{p}e_{21}^{T}\mathbf{p},

then 𝐱~\widetilde{\mathbf{x}} is an eigenvector of A+EA+E.

Proof.

(6.10) is equivalent to

(YT𝐩𝐱T)(A+E)(𝐱+Y𝐩)=0,(Y^{T}-\mathbf{p}\mathbf{x}^{T})(A+E)(\mathbf{x}+Y\mathbf{p})=0,

which gives

Y~T(A+E)𝐱~=0.\widetilde{Y}^{T}(A+E)\widetilde{\mathbf{x}}=0.

Since (𝐱~Y~)(\widetilde{\mathbf{x}}\;\widetilde{Y}) is an orthogonal matrix, 𝐱~\widetilde{\mathbf{x}} is an eigenvector of A+EA+E. ∎

6.2.3. Finding small 𝐩\mathbf{p} when the additive spectral gap is large

Notation 6.11.

We define MMn1()M\in M_{n-1}(\mathbb{R}) as M=(d+e11)In1LE22M=(d+e_{11})I_{n-1}-L-E_{22}.

Proposition 6.12.

If δ>2cn+2\displaystyle{\delta>\frac{2}{c}\sqrt{n}+2} for some value 0<c<10<c<1, then MM is non-singular and

(6.11) M11δ2<c2n.\|M^{-1}\|\leq\frac{1}{\delta-2}<\frac{c}{2\sqrt{n}}.
Proof.

Since (d+e11)In1L(d+e_{11})I_{n-1}-L is a diagonal matrix and E22E_{22} is a symmetric matrix, MM is symmetrix. Recall that E221\|E_{22}\|\leq 1. By Rayleigh’s principle and Observation 6.9,

minspec(M)\displaystyle\min\operatorname{spec}(M) =min𝐱=1𝐱T((d+e11)In1LE22)𝐱\displaystyle=\min_{\|\mathbf{x}\|=1}\mathbf{x}^{T}((d+e_{11})I_{n-1}-L-E_{22})\mathbf{x}
min𝐱=1𝐱T((d+e11)In1L)𝐱1\displaystyle\geq\min_{\|\mathbf{x}\|=1}\mathbf{x}^{T}((d+e_{11})I_{n-1}-L)\mathbf{x}-1
=min𝐱=1spec((d+e11)IL)1\displaystyle=\min_{\|\mathbf{x}\|=1}\operatorname{spec}((d+e_{11})I-L)-1
=d+e11λ21\displaystyle=d+e_{11}-\lambda_{2}-1
δ2.\displaystyle\geq\delta-2.

Therefore all eigenvalues of MM are positive, and

M1=maxspec(M1)=1minspec(M)1δ2<c2n.\|M^{-1}\|=\max\operatorname{spec}(M^{-1})=\frac{1}{\min\;\operatorname{spec}(M)}\leq\frac{1}{\delta-2}<\frac{c}{2\sqrt{n}}.\qed
Proposition 6.13.

We write (6.10) in terms of MM:

(6.12) M𝐩=e21𝐩e21T𝐩.M\mathbf{p}=e_{21}-\mathbf{p}e_{21}^{T}\mathbf{p}.

Let θ=M11\theta=\|M^{-1}\|^{-1} and η=e21\eta=\|e_{21}\|. We claim that if δ>2cn+2\delta>\frac{2}{c}\sqrt{n}+2 for some value 0<c<10<c<1, then there exists 𝐩\mathbf{p} with

(6.13) 𝐩<2ηθ+θ24η2\|\mathbf{p}\|<\frac{2\eta}{\theta+\sqrt{\theta^{2}-4\eta^{2}}}

such that (6.12) holds, and thus 𝐱~\widetilde{\mathbf{x}} defined by (6.9) is an eigenvector of A+EA+E.

Proof.

We want to find a solution with small norm to the non-linear equation (6.12). We do this by an iterative construction.

We define a sequence of vectors 𝐩0,𝐩1,\mathbf{p}_{0},\mathbf{p}_{1},\dots such that

𝐩0=0and𝐩i=M1(e21𝐩i1e21T𝐩i1),fori1.\mathbf{p}_{0}=0\quad\text{and}\quad\mathbf{p}_{i}=M^{-1}(e_{21}-\mathbf{p}_{i-1}e_{21}^{T}\mathbf{p}_{i-1}),\quad\text{for}\quad i\geq 1.

Then

𝐩iM1(|e21|+𝐩i12|e21|)η(1+𝐩i12)θ.\|\mathbf{p}_{i}\|\leq\|M^{-1}\|(|e_{21}|+\|\mathbf{p}_{i-1}\|^{2}|e_{21}|)\leq\frac{\eta(1+\|\mathbf{p}_{i-1}\|^{2})}{\theta}.

We claim that the sequence {𝐩i}\{\mathbf{p}_{i}\} converges. We define

ξ0=0,ξi=η(1+ξi12)θ,fori1.\xi_{0}=0,\quad\xi_{i}=\frac{\eta(1+\xi_{i-1}^{2})}{\theta},\quad\text{for}\quad i\geq 1.

Then 𝐩iξi\|\mathbf{p}_{i}\|\leq\xi_{i}. Since ξ1=ηθ>ξ0\xi_{1}=\frac{\eta}{\theta}>\xi_{0}, we can prove by induction that ξ0,ξ1,ξ2,\xi_{0},\xi_{1},\xi_{2},\dots is monotone increasing. Let

ϕ(ξ)=η(1+ξ2)θ.\phi(\xi)=\frac{\eta(1+\xi^{2})}{\theta}.

This function is monotone increasing in ξ\xi, and has a fixed point at ξ=2ηθ+θ24η2\displaystyle{\xi=\frac{2\eta}{\theta+\sqrt{\theta^{2}-4\eta^{2}}}}. Then ξi<ξi+1=ϕ(ξi)<ϕ(ξ)=ξ\xi_{i}<\xi_{i+1}=\phi(\xi_{i})<\phi(\xi)=\xi. Therefore the sequence {ξi}\{\xi_{i}\} converges to ξ\xi. Thus

𝐩iξiξ=2ηθ+θ24η2<2ηθ.\|\mathbf{p}_{i}\|\leq\xi_{i}\leq\xi=\frac{2\eta}{\theta+\sqrt{\theta^{2}-4\eta^{2}}}<\frac{2\eta}{\theta}.

Next we prove the convergence of {𝐩0,𝐩1,}\{\mathbf{p}_{0},\mathbf{p}_{1},\dots\}. For any i2i\geq 2,

𝐩i𝐩i1\displaystyle\|\mathbf{p}_{i}-\mathbf{p}_{i-1}\| =|M1(𝐩i1e21T𝐩i1𝐩i2e21T𝐩i2)\displaystyle=|M^{-1}(\mathbf{p}_{i-1}e_{21}^{T}\mathbf{p}_{i-1}-\mathbf{p}_{i-2}e_{21}^{T}\mathbf{p}_{i-2})
=|M1(𝐩i1+𝐩i2)e21T(𝐩i1𝐩i2)|\displaystyle=|M^{-1}(\mathbf{p}_{i-1}+\mathbf{p}_{i-2})e_{21}^{T}(\mathbf{p}_{i-1}-\mathbf{p}_{i-2})|
2M1𝐩i1𝐩i1η𝐩i1𝐩i2\displaystyle\leq 2\|M^{-1}\|\|\mathbf{p}_{i-1}\|\|\mathbf{p}_{i-1}\|\cdot\eta\cdot\|\mathbf{p}_{i-1}-\mathbf{p}_{i-2}\|
4η2θ2𝐩i1𝐩i2.\displaystyle\leq\frac{4\eta^{2}}{\theta^{2}}\|\mathbf{p}_{i-1}-\mathbf{p}_{i-2}\|.

Then 𝐩i𝐩0ρi𝐩1𝐩0\|\mathbf{p}_{i}-\mathbf{p}_{0}\|\leq\rho^{i}\|\mathbf{p}_{1}-\mathbf{p}_{0}\|, where ρ=4η2θ2<c24n<1\rho=\displaystyle{\frac{4\eta^{2}}{\theta^{2}}}<\frac{c^{2}}{4n}<1. Therefore {𝐩i}\{\mathbf{p}_{i}\} is a Cauchy sequence in n1\mathbb{R}^{n-1} and has a limit 𝐩\mathbf{p}. Thus a solution 𝐩\mathbf{p} exists, with norm satisfying (6.13).

6.2.4. Using the bound on 𝐩\|\mathbf{p}\| to show that 𝐱~\tilde{\mathbf{x}} is the principal eigenvector

Proposition 6.14.

If δ>2cn+2\delta>\frac{2}{c}\sqrt{n}+2 for some value 0<c<10<c<1, then the principal eigenvector 𝐱~\widetilde{\mathbf{x}} of A+EA+E can be writen in the form

𝐱~=𝐱+Y𝐩1+𝐩2where𝐩<cn.\widetilde{\mathbf{x}}=\frac{\mathbf{x}+Y\mathbf{p}}{\sqrt{1+\|\mathbf{p}\|^{2}}}\quad\text{where}\quad\|\mathbf{p}\|<\frac{c}{\sqrt{n}}.
Proof.

We showed that there exists 𝐩\mathbf{p} with

𝐩<2ηθ+θ24η2<2ηθ<cn\|\mathbf{p}\|<\frac{2\eta}{\theta+\sqrt{\theta^{2}-4\eta^{2}}}<\frac{2\eta}{\theta}<\frac{c}{\sqrt{n}}

such that 𝐱+Y𝐩1+𝐩2\displaystyle{\frac{\mathbf{x}+Y\mathbf{p}}{\sqrt{1+\|\mathbf{p}\|^{2}}}} is an eigenvector of A+EA+E. It remains to show that this is the principal eigenvector. Since Y=1\|Y\|=1 and 𝐩<cn\|\mathbf{p}\|<\displaystyle{\frac{c}{\sqrt{n}}} where 0<c<10<c<1, and since 𝐱=(1n,,1n)\mathbf{x}=\displaystyle{\left(\frac{1}{\sqrt{n}},\dots,\frac{1}{\sqrt{n}}\right)}, all coordinates of 𝐱~\widetilde{\mathbf{x}} are positive. Therefore by Fact 2.10, 𝐱~\widetilde{\mathbf{x}} is the principal eigenvector of A+EA+E. ∎

6.2.5. Completing the proof of Theorem 6.1

Following the argument above, we know the smallest possible coordinate of 𝐱~\widetilde{\mathbf{x}} is 1cn\displaystyle{\frac{1-c}{\sqrt{n}}}, and the largest possible coordinate of 𝐱~\widetilde{\mathbf{x}} is 1+cn\displaystyle{\frac{1+c}{\sqrt{n}}}. Therefore the ratio of G±eG\pm e is as claimed. This completes the proof. ∎

7. Open questions

In Section 4.2, we showed that if we remove a non-bridge edge ee from a connected regular graph such that the endpoints of ee are of bounded distance in GeG-e, then γ(Ge)\gamma(G-e) is polynomially bounded. Is there also a polynomial bound when the endpoints of ee are of unbounded distance in GeG-e?

Question 7.1.

If we remove a non-bridge edge ee from a connected regular graph GG, is γ(Ge)\gamma(G-e) always polynomially bounded?

In Section 4.1, we showed that if we add an edge ee between two vertices at distance 22 to a connected regular graph GG of bounded degree, then γ(G+e)\gamma(G+e) can be exponential. Can γ(G+e)\gamma(G+e) be exponential when ee is between two vertices of unbounded distance in GG?

Question 7.2.

If we add an edge ee to a connected regular graph GG with bounded degree dd, such that the disance between the endpoints of ee in GG is unbouneded, can γ(G+e)\gamma(G+e) be exponential in nn?

In Section 6.2, we showed that for a connected regular graph GG, an additive spectral gap greater than 2n2\sqrt{n} implies that γ(G±e)\gamma(G\pm e) is bounded (Theorem 6.1). However, this bound ceases to work at all for GG with an additive spectral gap δ2n\delta\leq 2\sqrt{n}. Is this a limitaton of the method, or is there really an abrupt change at δ=2n\delta=2\sqrt{n}?

Question 7.3.

Is there a family of connected regular graphs GG with an additive spectral gap slightly less than 2n2\sqrt{n} and with γ(G±e)\gamma(G\pm e) not bounded?

Acknowledgments

I am very grateful to Prof. Babai for his guidance and patience, and to Prof. Peter May for organizing the UChicago Math REU.

References

  • [1] Miroslav Fiedler. Algebraic connectivity of graphs. Czechoslovak Math. J. (1973), 23(2):298-305. DOI
  • [2] Noga Alon, Vitali D. Milman. λ1\lambda_{1}, isoperimetric inequalities for graphs, and superconcentrators. J. Combinatorial Theory, Series B (1985), 38:73-88. DOI
  • [3] Gilbert W. Stewart, Ji-guang Sun. Matrix Perturbation Theory, Academic press, 1990.
  • [4] Edwin R. van Dam, Willem H. Haemers. Eigenvalues and the Diameter of Graphs. Linear and Multilinear Algebra (1995), 39:33-44. DOI
  • [5] Andrew Y. Ng, Alice X. Zheng, Michael I. Jordan. Link analysis, eigenvectors and stability. In: 17th Internat. Joint Conf. on A. I. (IJCAI’01), vol. 2, pp. 903-910, Morgan Kaufmann, 2001. Author’s website
  • [6] Sebastian M. Cioabă, David A. Gregory. Principal eigenvectors of irregular graphs. Electr. J. Linear Algebra (2007), 16:366-379. DOI
  • [7] Sebastian M. Cioabă, David A. Gregory, Vladimir Nikiforov. Extreme eigenvalues of nonregular graphs. J. Combinatorial Theory, Series B (2007), 97(3):483-486. DOI
  • [8] Michael Tait, Josh Tobin. Characterizing graphs of maximum principal ratio. Electr. J. Linear Algebra (2018), 34:61-70. DOI