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

New results in vertex sedentariness

Hermie Monterde ​​1
Abstract

A vertex in a graph is said to be sedentary if a quantum state assigned on that vertex tends to stay on that vertex. Under mild conditions, we show that the direct product and join operations preserve vertex sedentariness. We also completely characterize sedentariness in blow-up graphs. These results allow us to construct new infinite families of graphs with sedentary vertices. We prove that a vertex with a twin is either sedentary or admits pretty good state transfer. Moreover, we give a complete characterization of twin vertices that are sedentary, and provide sharp bounds on their sedentariness. As an application, we determine the conditions in which perfect state transfer, pretty good state transfer and sedentariness occur in complete bipartite graphs and threshold graphs of any order.

Keywords: quantum walk, sedentary vertex, twin, direct product, adjacency matrix, Laplacian matrix

MSC2010 Classification: 05C50; 05C22; 15A16; 15A18; 81P45;

11footnotetext: Department of Mathematics, University of Manitoba, Winnipeg, MB, Canada R3T 2N2

1 Introduction

The concept of sedentariness was introduced by Godsil to study the ‘stay-at-home’ behaviour of vertices in families of graphs as their number of vertices tend to infinity [SedQW]. In particular, Godsil showed that complete graphs and large classes of strongly regular graphs are sedentary families. Recently, Monterde formalized the notion of a sedentary vertex and systematically studied both sedentary families and sedentary vertices [Monterde2023]. She proved that Cartesian products preserve vertex sedentariness, and provided examples of strongly cospectral vertices that are sedentary. She also showed that a vertex with at least two twins is sedentary and established lower bounds on their sedentariness. Her results allowed for the construction of new graphs with sedentary vertices using joins, blow-ups and attachment of pendent paths.

In this paper, we continue the investigation of vertex sedentariness. In Section 2, we introduce some graph theory, matrix theory and quantum walks background. Section 3 is devoted to a summary of known results about sedentariness that are useful throughout the paper. We highlight the fact that pretty good state transfer and sedentariness are mutually exclusive types of state transfer. We also provide infinite families of graphs containing vertices that are neither sedentary nor involved in pretty good state transfer (Example 3 and Remark 36). In Section 4, we prove an interesting result which states that a vertex with a twin only exhibits either sedentariness or pretty good state transfer (Corollary 12). This allows us to characterize twin vertices that are sedentary (Theorem 14) and provide an improved bound on their sedentariness (Theorem 15). We then apply our results to characterize pretty good state transfer and sedentariness in complete multipartite graphs in Section 5 and threshold graphs in Section 6. In particular, we show that a huge fraction of complete multipartite graphs are Laplacian sedentary at every vertex (Corollary 22). In fact, the only complete multipartite graphs that do not contain a Laplacian sedentary vertex are cocktail party graphs on twice an even number of vertices (Corollary 23). For threshold graphs, we characterize Laplacian sedentary vertices (Theorem 27), and provide tight lower bounds on sedentariness, which vary depending on the cell containing the vertex and the form of the threshold graph in question (Theorem 28). We also establish that almost all complete multipartite graphs and threshold graphs contain a sedentary vertex (Corollary 31). Section 7 explores sedentariness in direct products. We give sufficient conditions such that the direct product of a complete graph with another graph yields sedentary vertices (Theorem 32(1)). This yields new families of graphs with sedentary vertices, such as the direct product of complete graphs under mild conditions (Theorem 34). We also characterize sedentariness in bipartite doubles (Theorem 32(2)). In particular, we show that the bipartite double of a complete graph does not exhibit sedentariness (Corollary 35(1)). This prompts us to provide an infinite family of connected bipartite doubles containing sedentary vertices (Theorem 37). Section 8 is dedicated to characterizing adjacency sedentariness in blow-up graphs. Prior to this work, Monterde showed that the blow-ups of m3m\geq 3 copies of a graph are sedentary at every vertex. Here, we resolve the case m=2m=2 (Corollary 41) and provide sharper bounds for sedentariness in blow-ups (Theorem 40). In Section 9, we prove that the join operation preserves adjacency and Laplacian sedentariness under mild conditions (Theorem 45). This provides yet another method for constructing new graphs with sedentary vertices. Finally, we present open questions in Section 10.

2 Preliminaries

Throughout, we let XX be a weighted and undirected graph with possible loops, no multiple edges, all positive edge weights and vertex set V(X)V(X). The adjacency matrix A(X)A(X) of XX is defined entrywise as

A(X)u,v={ωu,v,if u and v are adjacent0,otherwise,A(X)_{u,v}=\begin{cases}\omega_{u,v},&\text{if $u$ and $v$ are adjacent}\\ 0,&\text{otherwise},\end{cases}

where ωu,v\omega_{u,v} is the weight of the edge (u,v)(u,v). The degree matrix D(X)D(X) of XX is the diagonal matrix of vertex degrees of XX, where deg(u)=2ωu,u+juωu,j\operatorname{deg}(u)=2\omega_{u,u}+\sum_{j\neq u}\omega_{u,j} for each uV(X)u\in V(X). For each qq\in\mathbb{R}, we call the matrix Mq(X)=qD(X)+A(X)M_{q}(X)=qD(X)+A(X) a generalized adjacency matrix for XX. Note that M0(X)=A(X)M_{0}(X)=A(X) and M1(X)=L(X)M_{-1}(X)=-L(X), where L(X)L(X) is the Laplacian matrix of XX. We say that XX is weighted kk-regular if deg(u)ωu,u\operatorname{deg}(u)-\omega_{u,u} is a constant kk for each uV(X)u\in V(X). Equivalently, each row sum of A(X)A(X) is a constant kk. If XX is simple (i.e., loopless), then XX is weighted kk-regular if and only if D(X)=kID(X)=kI. If the context is clear, then we write Mq(X)M_{q}(X), A(X)A(X), and L(X)L(X) resp. as MqM_{q}, AA, and LL. We also write MqM_{q} as MM when qq is not needed. We represent the transpose of MM by MTM^{T}, and the characteristic polynomial of MM in the variable tt by ϕ(M,t)\phi(M,t). We denote the simple unweighted empty, complete, cycle and path graphs on nn vertices resp. by OnO_{n}, KnK_{n}, CnC_{n} and PnP_{n}. The all-ones vector of order nn, the zero vector of order nn, the m×nm\times n all-ones matrix, and the n×nn\times n identity matrix are denoted by 1n\textbf{1}_{n}, 0n\textbf{0}_{n}, Jm,n\textbf{J}_{m,n} and InI_{n}, respectively. If m=nm=n, then we write Jm,n\textbf{J}_{m,n} as Jn\textbf{J}_{n}. If the context is clear, then we simply write these matrices resp. as 1, 0, J and II.

Let HH be a Hermitian matrix whose rows and columns are indexed by the vertices of XX satisfying the property Hu,v=0H_{u,v}=0 if and only if no edge exists between uu and vv in XX. A (continuous-time) quantum walk on XX with respect to HH is determined by the unitary matrix

UH(t)=eitH.U_{H}(t)=e^{itH}. (1)

In this work, we focus on the case H=MqH=M_{q}. If qq is not important, then we simply write UMq(t)U_{M_{q}}(t) as UM(t)U_{M}(t). Unless otherwise specified, the results in this paper apply to Mq=MM_{q}=M for all qq\in\mathbb{R} Since MM is real symmetric, it admits a spectral decomposition

M=jλjEj,M=\sum_{j}\lambda_{j}E_{j}, (2)

where the λj\lambda_{j}’s are the distinct real eigenvalues of MM and each EjE_{j} is the orthogonal projection matrix onto the eigenspace associated with λj\lambda_{j}. The eigenvalue support of uV(X)u\in V(X) is the set

σu(M)={λj:Ejeu0},\sigma_{u}(M)=\{\lambda_{j}:E_{j}\textbf{e}_{u}\neq\textbf{0}\},

where eu\textbf{e}_{u} denotes the characteristic vector of vertex uu. From (2), we may write (1) as

UM(t)=jeitλjEjU_{M}(t)=\sum_{j}e^{it\lambda_{j}}E_{j}.

We say that two vertices uu and vv are cospectral if (Ej)u,u=(Ej)v,v(E_{j})_{u,u}=(E_{j})_{v,v} for each jj, and strongly cospectral if Ejeu=±EjevE_{j}\textbf{e}_{u}=\pm E_{j}\textbf{e}_{v} for each jj. Note that if uu and vv are cospectral, then UM(t)u,u=UM(t)v,vU_{M}(t)_{u,u}=U_{M}(t)_{v,v} for all tt.

With respect to the matrix MM, we say that perfect state transfer (PST) occurs between two vertices uu and vv at time τ\tau if |U(τ)u,v|=1|U(\tau)_{u,v}|=1. The minimum τ>0\tau>0 such that |U(τ)u,v|=1|U(\tau)_{u,v}|=1 is called the minimum PST time between uu and vv. We say that uu is periodic at time τ\tau if |U(τ)u,u|=1|U(\tau)_{u,u}|=1. The minimum τ>0\tau>0 such that |U(τ)u,u|=1|U(\tau)_{u,u}|=1 is called the minimum period of uu, which we denote by ρ\rho. We say that pretty good state transfer (PGST) occurs between uu and vv if limτk|U(τk)u,v|=1\lim_{\tau_{k}\rightarrow\infty}|U(\tau_{k})_{u,v}|=1 for some {τk}\{\tau_{k}\}\subseteq\mathbb{R}. Further, if XX is a simple weighted kk-regular graph, then UM(t)=eit(qkI+A)=eit(qk)UA(t)U_{M}(t)=e^{it(qkI+A)}=e^{it(qk)}U_{A}(t), and so for all qq\in\mathbb{R},

|UM(t)u,v|2=|UA(t)u,v|2.\left|U_{M}(t)_{u,v}\right|^{2}=\left|U_{A}(t)_{u,v}\right|^{2}.

holds for all tt and for all u,vV(X)u,v\in V(X). That is, the quantum walks with respect to AA and MM are equivalent for all qq\in\mathbb{R}, i.e., they exhibit the same types of state transfer.

The join of weighted graphs XX and YY, denoted XYX\vee Y, is the resulting graph obtained by joining every vertex of XX with every vertex of YY with an edge of weight one. In particular, the graph O1YO_{1}\vee Y is called a cone on YY and the lone vertex of O1O_{1} is called the apex. If X{O2,K2}X\in\{O_{2},K_{2}\}, then XYX\vee Y is called a double cone on YY and the two vertices of XX are called the apexes. We denote the cocktail party graph on 2k2k vertices by CP(2k)CP(2k), and the complete graph on nn vertices minus an edge by Kn\eK_{n}\backslash e.

3 Sedentariness

In this section, we gather some known results on sedentariness that will be useful throughout the paper.

Definition 1.

We say that vertex uu of XX is CC-sedentary if for some constant 0<C10<C\leq 1,

inft>0|UM(t)u,u|C.\inf_{t>0}\lvert U_{M}(t)_{u,u}\rvert\geq C. (3)

If equality holds in (3), then we say that uu is sharply CC-sedentary, while if the infimum in (3) is attained at some t0>0t_{0}>0, then we say that uu is tightly CC-sedentary at time t=t0t=t_{0}. Further, we say that uu is not sedentary if inft>0|UM(t)u,u|=0\inf_{t>0}\lvert U_{M}(t)_{u,u}\rvert=0.

If CC is not important, then we respectively say sedentary, sharply sedentary, and tightly sedentary. Sedentariness of a vertex uu implies that |UM(t)u,u|\lvert U_{M}(t)_{u,u}\rvert is bounded away from 0, and so physically speaking, the quantum state initially at vertex uu tends to stay at uu.

We now state an important property of sedentary vertices [Monterde2023, Proposition 2c].

Proposition 2.

A sedentary vertex cannot be involved in pretty good state transfer. Conversely, a vertex involved in pretty good state transfer cannot be sedentary.

Proposition 2 implies that sedentarines and PGST are mutually exclusive types of quantum state transfer, although it is possible that a given vertex in a graph exhibits neither sedentarines nor PGST. We illustrate the latter statement by providing an infinite family of graphs that satisfy this property.

Example 3.

For each n3n\geq 3, consider K1,nK_{1,n} with central vertex uu. Then UA(π2n)u,u=0U_{A}(\frac{\pi}{2\sqrt{n}})_{u,u}=0, and so uu is not sedentary. Moreover, since uu is not strongly cospectral with any vertex in XX, it cannot exhibit PGST.

In Remark 36, we provide an infinite family of graphs containing periodic and strongly cospectral vertices that are neither sedentary nor involved in PST with respect to the adjacency and Laplacian matrix.

The following result is due to Monterde [Monterde2023, Theorem 11].

Theorem 4.

Let uu be a vertex of XX with σu(M)={λ1,,λr}\sigma_{u}(M)=\{\lambda_{1},\ldots,\lambda_{r}\}, where EjE_{j} is the orthogonal projection matrix corresponding to λj\lambda_{j}. If SS is a non-empty proper subset of σu(M)\sigma_{u}(M), say S={λ1,,λs}S=\{\lambda_{1},\ldots,\lambda_{s}\}, such that

j=1s(Ej)u,u=a\sum_{j=1}^{s}(E_{j})_{u,u}=a (4)

for some 12a<1\frac{1}{2}\leq a<1, then

|UM(t)u,u||j=1seitλj(Ej)u,u|(1a)for allt.\lvert U_{M}(t)_{u,u}\rvert\geq\bigg{|}\sum_{j=1}^{s}e^{it\lambda_{j}}(E_{j})_{u,u}\bigg{|}-(1-a)\quad\text{for all}\ t. (5)

Moreover, if there exists a time t1>0t_{1}>0 such that |j=1seit1λj(Ej)u,u|1a\bigg{|}\sum_{j=1}^{s}e^{it_{1}\lambda_{j}}(E_{j})_{u,u}\bigg{|}\geq 1-a, and for all j{1,,s}j\in\{1,\ldots,s\} and k{s+1,,r}k\in\{s+1,\ldots,r\},

eit1(λ1λj)=1andeit1(λ1λk)=1,e^{it_{1}(\lambda_{1}-\lambda_{j})}=1\quad\text{and}\quad e^{it_{1}(\lambda_{1}-\lambda_{k})}=-1, (6)

then equality holds in (5), in which case |UM(t1)u,u|=2a1\lvert U_{M}(t_{1})_{u,u}\rvert=2a-1 and uu is periodic at time 2t12t_{1}.

Remark 5.

From (5), it suffices to find a non-empty proper subset SS of σu(M)\sigma_{u}(M) with jS(Ej)u,u=a12\sum_{j\in S}(E_{j})_{u,u}=a\geq\frac{1}{2} such that |j=1seitλj(Ej)u,u|(1a)\left\lvert\sum_{j=1}^{s}e^{it\lambda_{j}}(E_{j})_{u,u}\right\rvert-(1-a) is bounded away from 0 for all tt.

We also have the following result, which characterizes the occurrence of (6). In what follows, we denote the largest power of two that divides an integer aa by ν2(a)\nu_{2}(a).

Lemma 6.

Let uV(X)u\in V(X) and suppose Sσu(M)S\subseteq\sigma_{u}(M) satisfies (4). If ϕ(M,t)[t]\phi(M,t)\in\mathbb{Z}[t], then (6) holds if and only if all of the following conditions hold

  1. 1.

    Either (a) all eigenvalues in σu(M)\sigma_{u}(M) are integers or (b) all eigenvalues in λjσu(M)\lambda_{j}\in\sigma_{u}(M) are quadratic integers of the form 12(a+bjΔ)\frac{1}{2}(a+b_{j}\sqrt{\Delta}), where a,bj,Δa,b_{j},\Delta are integers and Δ>1\Delta>1 is square-free.

  2. 2.

    For all λj,λmS\lambda_{j},\lambda_{m}\in S and λk,λσu(M)\S\lambda_{k},\lambda_{\ell}\in\sigma_{u}(M)\backslash S, ν2(λjλmΔ)>ν2(λkλmΔ)=ν2(λλmΔ)\nu_{2}(\frac{\lambda_{j}-\lambda_{m}}{\sqrt{\Delta}})>\nu_{2}(\frac{\lambda_{k}-\lambda_{m}}{\sqrt{\Delta}})=\nu_{2}(\frac{\lambda_{\ell}-\lambda_{m}}{\sqrt{\Delta}}), where Δ=1\Delta=1 whenever (1a) holds.

Moreover, the minimum time in which (6) holds is t1=πgt_{1}=\frac{\pi}{g}, where g=gcd(𝒯)g=\operatorname{gcd}(\mathcal{T}), 𝒯={λ1λΔ:λσu(M)}\mathcal{T}=\{\frac{\lambda_{1}-\lambda}{\sqrt{\Delta}}:\lambda\in\sigma_{u}(M)\} and λ1S\lambda_{1}\in S is fixed, and uu is periodic with ρ=2t1\rho=2t_{1}.

Remark 7.

If uu and vv are strongly cospectral in XX and Lemma 6(1-2) holds, then PST occurs between uu and vv at t1t_{1} and S=σuv+(M)S=\sigma_{uv}^{+}(M). In this case, a=12a=\frac{1}{2} so (4) yields UM(t1)u,u=0U_{M}(t_{1})_{u,u}=0. Thus, uu is not sedentary.

We also state Lemmas 9 and 12 in [Monterde2023], respectively.

Lemma 8.

Let uu be a periodic vertex in XX. Then uu is sedentary if and only if UM(t)u,u0U_{M}(t)_{u,u}\neq 0 for all t[0,ρ]t\in[0,\rho]. In particular, if uu is sedentary, then it is tightly CC-sedentary, where C=mint[0,ρ]|U(t)u,u|C=\min_{t\in[0,\rho]}|U(t)_{u,u}|.

Lemma 9.

Let uu be a vertex of XX with σu(M)={λ1,,λr}\sigma_{u}(M)=\{\lambda_{1},\ldots,\lambda_{r}\}, where EjE_{j} is the orthogonal projection matrix corresponding to λj\lambda_{j}. Suppose SS is a non-empty proper subset of σu(M)\sigma_{u}(M), say S={λ1,,λs}S=\{\lambda_{1},\ldots,\lambda_{s}\}, such that (4) holds for some 12a<1\frac{1}{2}\leq a<1. The following are equivalent.

  1. 1.

    If j\ell_{j} and mjm_{j} are integers such that j=1smjλj+j=s+1rjλj=0\sum_{j=1}^{s}m_{j}\lambda_{j}+\sum_{j=s+1}^{r}\ell_{j}\lambda_{j}=0 and j=1smj+j=s+1rj=0\sum_{j=1}^{s}m_{j}+\sum_{j=s+1}^{r}\ell_{j}=0, then j=1smj\sum_{j=1}^{s}m_{j} is even.

  2. 2.

    There exists a sequence {tk}\{t_{k}\}\subseteq\mathbb{R} such that limk|UM(tk)u,u|=2a1\lim_{k\rightarrow\infty}\lvert U_{M}(t_{k})_{u,u}\rvert=2a-1.

We close this section with a remark that Lemmas 6 and 9 are equivalent for periodic vertices.

4 Graphs with twins

Denote the neighbourhood of a vertex uu in XX by NX(u)N_{X}(u). Two vertices uu and vv of XX are twins if

  1. 1.

    NX(u)\{u,v}=NX(v)\{u,v}N_{X}(u)\backslash\{u,v\}=N_{X}(v)\backslash\{u,v\},

  2. 2.

    the edges (u,w)(u,w) and (v,w)(v,w) have the same weight for each wNX(u)\{u,v}w\in N_{X}(u)\backslash\{u,v\}, and

  3. 3.

    the loops on uu and vv have the same weight, and this weight is zero if those loops are absent.

We allow twins to be adjacent. A maximal subset T=T(ω,η)T=T(\omega,\eta) of V(X)V(X) with at least two elements is a twin set in XX if the vertices in TT are pairwise twins, each uTu\in T has a loop of weight ω\omega, and every pair of vertices in TT are connected by an edge with weight η\eta. In particular, if XX is a simple unweighted graph, then ω=0\omega=0 and η{0,1}\eta\in\{0,1\}. As an example, the apexes of a double cone form a twin set of size two. Note that the vertices in a twin set TT are pairwise cospectral [Monterde2022, Corollary 2.11]. Thus, if u,vTu,v\in T, then UM(t)u,u=UM(t)v,vU_{M}(t)_{u,u}=U_{M}(t)_{v,v} for all tt, and so a vertex in TT is sedentary if and only if each vertex in TT is sedentary.

The following fact is due to Monterde [Monterde2022, Lemma 2.9].

Lemma 10.

Let T=T(ω,η)T=T(\omega,\eta) be a twin set in XX. Then u,vTu,v\in T if and only if euev\textbf{e}_{u}-\textbf{e}_{v} is an eigenvector associated to the eigenvalue θ=qdeg(u)+ωη\theta=q\operatorname{deg}(u)+\omega-\eta of MM.

Our next result is due to Kirkland et al. [Kirkland2023, Corollary 1].

Theorem 11.

Let TT be a twin set in XX. Then for any two vertices uu and vv in TT,

|UM(t)u,u|+|UM(t)u,v|1\lvert U_{M}(t)_{u,u}\rvert+\lvert U_{M}(t)_{u,v}\rvert\geq 1\quad for all tt\in\mathbb{R}.

Corollary 12.

Let TT be a twin set in XX. Then a vertex uTu\in T is either sedentary or involved in pretty good state transfer with some vertex vuv\neq u in XX. If the latter holds, then T={u,v}T=\{u,v\}.

Proof.

Theorem 11 yields the first statement. In particular, if uTu\in T is involved in PGST with vv, then they are strongly cospectral, and so [Monterde2022, Theorem 3.9(2)] yields the second statement. ∎

Since strong cospectrality is a necessary condition for PGST, Corollary 12 yields the next result.

Corollary 13.

Let TT be a twin set in XX. If uTu\in T is not involved in strong cospectrality, then each uTu\in T is sedentary.

We now prove one of our main results, which characterizes sedentariness in twin vertices.

Theorem 14.

Let TT be a twin set in XX and consider the eigenvalue θ\theta in Lemma 10. Then each vertex in TT is sedentary if and only if one of the conditions below hold.

  1. 1.

    Either (i) |T|3|T|\geq 3 or (ii) T={u,v}T=\{u,v\} and there is an eigenvector wspan{euev}\textbf{w}\notin\operatorname{span}\{\textbf{e}_{u}-\textbf{e}_{v}\} of MM associated with θ\theta such that wTeu0\textbf{w}^{T}\textbf{e}_{u}\neq 0 or wTev0\textbf{w}^{T}\textbf{e}_{v}\neq 0. Here, any vertex in TT cannot exhibit strong cospectrality.

  2. 2.

    θ\theta is a simple eigenvalue of MM, so that T={u,v}T=\{u,v\} and uu and vv are strongly cospectral vertices, and there are integers mjm_{j} such that

    λjσuv+(M)mj(λjθ)=0andλjσuv+(M)mjis odd.\sum_{\lambda_{j}\in\sigma_{uv}^{+}(M)}m_{j}(\lambda_{j}-\theta)=0\quad\text{and}\quad\sum_{\lambda_{j}\in\sigma_{uv}^{+}(M)}m_{j}\ \text{is odd}.

    If we add that ϕ(M,t)[t]\phi(M,t)\in\mathbb{Z}[t] and uu is periodic, then the latter statement is equivalent to each eigenvalue λjσuv+(M)\lambda_{j}\in\sigma_{uv}^{+}(M) is of the form λj=θ+bjΔ\lambda_{j}=\theta+b_{j}\sqrt{\Delta}, where bjb_{j} is an integer and either Δ=1\Delta=1 or Δ>1\Delta>1 is a square-free integer and the ν2(bj)\nu_{2}(b_{j})’s are not all equal. In this case, uu is tightly sedentary.

Proof.

Combining condition (1) with [Monterde2022, Corollary 3.14] implies that each vertex in TT cannot be strongly cospectral with another vertex. Thus, each vertex in TT is sedentary by Corollary 13. This proves (1). To prove (2), note that the condition that θ\theta is a simple eigenvalue of MM implies that T={u,v}T=\{u,v\} and uu and vv are strongly cospectral [Monterde2022, Corollary 3.14]. Combining this with Corollary 12 and the characterizations of PST and PGST between twin vertices (see Theorems 11 and 14 in [Kirkland2023]) yields the desired results in (2). The last statement in (2) follows from Lemma 8. ∎

We now prove another main result in this paper, which is an improvement of [Monterde2023, Theorem 16].

Theorem 15.

Let TT be a twin set in XX and fix uTu\in T. Let 1\mathcal{B}_{1} be the resulting set after orthonormalizing {euev:vT\{u}}\{\textbf{e}_{u}-\textbf{e}_{v}:v\in T\backslash\{u\}\}, and let =12\mathcal{B}=\mathcal{B}_{1}\cup\mathcal{B}_{2} be an orthonormal basis for the eigenspace of MM associated with the eigenvalue θ\theta in Lemma 10. Define F=w2wwTF=\sum_{\textbf{w}\in\mathcal{B}_{2}}\textbf{w}\textbf{w}^{T}, where FF is absent whenever 2=\mathcal{B}_{2}=\varnothing. Then

|UM(t)u,u|2(Eθ)u,u1=12|T|+2Fu,ufor all t,|U_{M}(t)_{u,u}|\geq 2(E_{\theta})_{u,u}-1=1-\frac{2}{|T|}+2F_{u,u}\quad\text{for all $t$}, (7)

with equality if and only if (6) holds. The following also hold.

  1. 1.

    Suppose 12|T|+2Fu,u>01-\frac{2}{|T|}+2F_{u,u}>0.

    1. (a)

      If ϕ(M,t)[t]\phi(M,t)\in\mathbb{Z}[t] and each vertex in TT is periodic, then each vertex in TT is tightly CC-sedentary, where C=12|T|+2Fu,uC=1-\frac{2}{|T|}+2F_{u,u} whenever Lemma 6 holds and C>12|T|+2Fu,uC>1-\frac{2}{|T|}+2F_{u,u} otherwise.

    2. (b)

      Suppose each vertex in TT is not periodic. Then each vertex in TT is sharply (12|T|+2Fu,u)(1-\frac{2}{|T|}+2F_{u,u})-sedentary if and only if Lemma 9 holds.

  2. 2.

    If |T|3|T|\geq 3, then each vertex in TT is (12|T|+2Fu,u)\left(1-\frac{2}{|T|}+2F_{u,u}\right)-sedentary.

  3. 3.

    If T={u,v}T=\{u,v\} and there is an eigenvector wspan{euev}\textbf{w}\notin\operatorname{span}\{\textbf{e}_{u}-\textbf{e}_{v}\} of MM associated with θ\theta such that wTeu0\textbf{w}^{T}\textbf{e}_{u}\neq 0 or wTev0\textbf{w}^{T}\textbf{e}_{v}\neq 0, then each vertex TT is 2Fu,u2F_{u,u}-sedentary.

Proof.

Let 0m0_{m} denote the m×mm\times m zero matrix. From our assumption, we deduce that

Eθ=(I|T|1|T|J|T|0n|T|)+F,E_{\theta}=\left(I_{|T|}-\frac{1}{|T|}\textbf{J}_{|T|}\oplus 0_{n-|T|}\right)+F,

Taking S={θ}S=\{\theta\} and a=(Eθ)u,u=11|T|+Fu,u12a=(E_{\theta})_{u,u}=1-\frac{1}{|T|}+F_{u,u}\geq\frac{1}{2} in Theorem 4 yields

|UM(t)u,u|a(1a)=2a1=12|T|+2Fu,u.\begin{split}|U_{M}(t)_{u,u}|&\geq a-(1-a)=2a-1=1-\frac{2}{|T|}+2F_{u,u}.\end{split} (8)

Note that (1) is immediate from Lemma 6 and Lemma 9, while (2) follows directly from (8). Moreover, the assumption in (4) implies that Fu,u>0F_{u,u}>0, and so (8) yields |UM(t)u,u|2Fu,u>0|U_{M}(t)_{u,u}|\geq 2F_{u,u}>0. ∎

uuvvn2vertices\underbrace{}_{n-2\ \text{vertices}}
Figure 1: The graph PnP_{n}^{\prime} with twin vertices uu and vv marked white
Remark 16.

Let TT be a twin set in XX. If Theorem 14(1) holds, then Theorem 15(2-3) provides a bound for the sedentariness of vertices in TT. Now, if Theorem 14(2) holds, then =1=span{euev}\mathcal{B}=\mathcal{B}_{1}=\operatorname{span}\{\textbf{e}_{u}-\textbf{e}_{v}\} in Theorem 15, and so (7) yields the trivial inequality |UM(t)u,u|0|U_{M}(t)_{u,u}|\geq 0. In this case, T={u,v}T=\{u,v\} and uu and vv are strongly cospectral in XX, and so one may invoke Theorem 14(2) to determine whether uu is sedentary in XX or not. By explicitly calculating lower bounds for |UM(t)|u,u|U_{M}(t)|_{u,u}, Monterde provided sharp bounds for the sedentariness of the apexes of O2YO_{2}\vee Y for M{A,L}M\in\{A,L\} with the additional condition that YY is regular when M=AM=A (see Theorems 32 and 36 in [Monterde2023]). But in general, if uTu\in T is sedentary in XX, then we currently do not have a straightforward way of determining a bound for the sedentariness of vertices in TT.

Monterde first established Theorem 15(2) in [Monterde2023, Theorem 16], and she used this result provide infinite families of join graphs, blow-up graphs and graphs with tails that contain sedentary vertices. In the next examples, we provide infinite families of graphs that satisfy Theorem 15(3).

Example 17.

Let M=LM=L and YY be a simple weighted graph on nn vertices. Consider K2YK_{2}\vee Y and let u=1u=1 and v=2v=2 be the apexes of K2YK_{2}\vee Y. Then σu(L(K2Y))={0,θ}\sigma_{u}(L(K_{2}\vee Y))=\{0,\theta\}, where θ=n+2\theta=n+2. If \mathcal{B} is an orthonormal basis for the eigenspace associated with the eigenvalue nn of L(Y)L(Y), then

{12(e1e2),12n(n+2)[n1221n]}{[0v]:v}\left\{\frac{1}{\sqrt{2}}(\textbf{e}_{1}-\textbf{e}_{2}),\frac{1}{\sqrt{2n(n+2)}}\left[\begin{array}[]{ccccc}n\textbf{1}_{2}\\ -2\textbf{1}_{n}\end{array}\right]\right\}\cup\left\{\left[\begin{array}[]{ccccc}\textbf{0}\\ \textbf{v}\end{array}\right]:\textbf{v}\in\mathcal{B}\right\}

is as an orthonormal basis for the eigenspace associated with the eigenvalue θ\theta of L(K2Y)L(K_{2}\vee Y) and the set on the right is absent if YY is not a join graph. Hence, if we let w=12n(n+2)[n1221n]\textbf{w}=\frac{1}{\sqrt{2n(n+2)}}\left[\begin{array}[]{ccccc}n\textbf{1}_{2}\\ -2\textbf{1}_{n}\end{array}\right], then the conditions in Theorem 15(3) are met, and we conclude that

|UL(K2Y)(t)u,u|2F1,1=nn+2for all t|U_{L(K_{2}\vee Y)}(t)_{u,u}|\geq 2F_{1,1}=\frac{n}{n+2}\quad\text{for all $t$}

with equality if and only if t=jπn+2t=\frac{j\pi}{n+2} for all odd jj. Thus, vertices uu and vv in K2YK_{2}\vee Y are tightly (nn+2)(\frac{n}{n+2})-sedentary. This result is consistent with [Monterde2023, Theorem 29].

Example 18.

Let M=AM=A, n3n\geq 3 be odd and u=1u=1 be an end vertex of PnP_{n}. Let PnP_{n}^{\prime} be the resulting graph after adding vertex v:=n+1v:=n+1 to PnP_{n} such that uu and vv non-adjacent twins (see Figure 1). If x=12(e1en+1)\textbf{x}=\frac{1}{\sqrt{2}}(\textbf{e}_{1}-\textbf{e}_{n+1}) and w=2n+1[1,0,1,0,1,,(1)n12,0]T\textbf{w}=\frac{\sqrt{2}}{\sqrt{n+1}}[1,0,-1,0,1,\ldots,(-1)^{\frac{n-1}{2}},0]^{T}, then x and w are eigenvectors asscociated with the eigenvalue θ=0\theta=0 of AA. Orthonormalizing {x,w}\{\textbf{x},\textbf{w}\} yields an orthonormal basis {x,w}\{\textbf{x},\textbf{w}^{\prime}\} for the eigenspace associated to θ\theta, where w=12n[1,0,2,0,2,,0,2n+12,1]T\textbf{w}^{\prime}=\frac{1}{\sqrt{2n}}[1,0,2,0,-2,\ldots,0,2^{\frac{n+1}{2}},1]^{T}. Hence, Theorem 15(3) yields

|UA(Pn)(t)u,u|2F1,1=1nfor all t.|U_{A(P_{n}^{\prime})}(t)_{u,u}|\geq 2F_{1,1}=\frac{1}{n}\quad\text{for all $t$}.

If the nonzero eigenvalues in σu(A(P5))\sigma_{u}(A(P_{5}^{\prime})) are linearly independent over \mathbb{Q}, then a direct application of Lemma 9 implies that uu is sharply (1n)(\frac{1}{n})-sedentary in PnP_{n}^{\prime}. In particular, if n=3n=3, then uu is tightly (13)(\frac{1}{3})-sedentary in P3=K1,3P_{3}^{\prime}=K_{1,3}, while if n=5n=5, then σu(A(P5))={0,±(5±5)/2}\sigma_{u}(A(P_{5}^{\prime}))=\{0,\pm\sqrt{(5\pm\sqrt{5})/2}\}, and so uu is sharply (15)(\frac{1}{5})-sedentary in P5P_{5}^{\prime}. We conjecture that uu is sharply (1n)(\frac{1}{n})-sedentary in PnP_{n}^{\prime} for all odd n5n\geq 5.

5 Complete multipartite graphs

A complete bipartite graph Kn1,,nkK_{n_{1},\ldots,n_{k}} is the graph Kn1,,nk=j=1kOnjK_{n_{1},\ldots,n_{k}}=\bigvee_{j=1}^{k}O_{n_{j}}. If n=1n_{\ell}=1 (resp., n=2n_{\ell}=2) for some {1,,k}\ell\in\{1,\ldots,k\}, then we may write Kn1,,nk=O1jOnjK_{n_{1},\ldots,n_{k}}=O_{1}\vee\bigvee_{j\neq\ell}O_{n_{j}} (resp., Kn1,,nk=O2jOnjK_{n_{1},\ldots,n_{k}}=O_{2}\vee\bigvee_{j\neq\ell}O_{n_{j}}), and so we may view Kn1,,nkK_{n_{1},\ldots,n_{k}} as a cone (resp., double cone) on Y=jOnjY=\bigvee_{j\neq\ell}O_{n_{j}}. Our goal in this section to characterize PGST and sedentariness in complete multipartite graphs with respect to M{A,L}M\in\{A,L\}.

Laplacian case

We first analyze the quantum walk on a complete multipartite graphs using its Laplacian matrix.

Theorem 19.

Let X=Kn1,,nkX=K_{n_{1},\ldots,n_{k}} and n=j=1knjn=\sum_{j=1}^{k}n_{j}. If {1,,k}\ell\in\{1,\ldots,k\} and uu is a vertex of XX that belong to the partite set of size nn_{\ell}, then the following hold with respect to L=L(X)L=L(X).

  1. 1.

    If n=1n_{\ell}=1, then uu is tightly (12n)(1-\frac{2}{n})-sedentary at time t=jπnt=\frac{j\pi}{n} for any odd jj.

  2. 2.

    If n=2n_{\ell}=2 and n0n\equiv 0 (mod 4), then the two vertices of XX that belong to the partite set of size nn_{\ell} admit perfect state transfer with minimum PST time π2\frac{\pi}{2}.

  3. 3.

    If n=2n_{\ell}=2 and n2n\equiv 2 (mod 4), then uu is tightly (2n)(\frac{2}{n})-sedentary at time t=jπ2t=\frac{j\pi}{2} for any odd jj.

  4. 4.

    Let n=2n_{\ell}=2 and nn be odd. If n=3n=3, then uu is tightly (13)(\frac{1}{3})-sedentary at time t=jπt=j\pi for any odd jj. If n5n\geq 5, then uu is tightly (2n)(\frac{\sqrt{2}}{n})-sedentary at time t=jπ2t=\frac{j\pi}{2} for any odd jj.

  5. 5.

    If n3n_{\ell}\geq 3, then uu is tightly CC-sedentary, where C=12nC=1-\frac{2}{n_{\ell}} at time t=jπgt=\frac{j\pi}{g} whenever ν2(n)>ν2(n)\nu_{2}(n)>\nu_{2}(n_{\ell}), where g=gcd(n,n)g=\operatorname{gcd}(n,n_{\ell}) and jj is any odd integer and C>12nC>1-\frac{2}{n_{\ell}} otherwise.

Proof.

Since L(X)L(X) has integer eigenvalues, (1), (2), (3-4) and (5) resp. follow from [Monterde2023, Theorem 29], [Alvir2016, Corollary 4], [Monterde2023, Theorem 32] and Theorem 15(2)-[Monterde2023, Theorem 35] combined. ∎

Since CP(2k)=kO2CP(2k)=\bigvee_{k}O_{2}, Theorem 19(2-3) yields our next result.

Corollary 20.

Let X=CP(2k)X=CP(2k). If kk is even, then perfect state transfer occurs between pairs of non-adjacent vertices in XX with minimum PST time π2\frac{\pi}{2}. Otherwise, every vertex in XX is tightly (1k)(\frac{1}{k})-sedentary.

Next, since Kn\e=O2Kn2K_{n}\backslash e=O_{2}\vee K_{n-2}, Theorem 19 gives us our next result.

Corollary 21.

For all n3n\geq 3, let X=Kn\eX=K_{n}\backslash e with non-adjacent vertices uu and vv.

  1. 1.

    For all n3n\geq 3, each degree n1n-1 vertex of XX is (12n)(1-\frac{2}{n})-sedentary at time t=jπnt=\frac{j\pi}{n} for any odd jj.

  2. 2.

    If n0n\equiv 0 (mod 4), then perfect state transfer occurs between uu and vv with minimum time π2\frac{\pi}{2}. Otherwise, uu is tightly CC-sedentary, where C=2nC=\frac{2}{n} whenever n2n\equiv 2 (mod 4) at time t=jπ2t=\frac{j\pi}{2}, C=13C=\frac{1}{3} whenever n=3n=3 at time t=jπt=j\pi and C=2nC=\frac{\sqrt{2}}{n} whenever n5n\geq 5 is odd at time t=jπ2t=\frac{j\pi}{2}, where jj is odd.

In contrast to complete graphs where each vertex is sedentary, we now show that a huge fraction of the family of complete multipartite bipartite graphs are Laplacian sedentary at every vertex.

Corollary 22.

Let X=Kn1,,nkX=K_{n_{1},\ldots,n_{k}} and n=j=1knjn=\sum_{j=1}^{k}n_{j}. If either (i) n0n\not\equiv 0 (mod 4) or (ii) n0n\equiv 0 (mod 4) and nj2n_{j}\neq 2 for each j{1,,k}j\in\{1,\ldots,k\}, then every vertex in XX is sedentary.

If n0n\equiv 0 (mod 4), then nn only has one partition where all parts are equal to two. Thus:

Corollary 23.

CP(2k)CP(2k) for even kk are the only complete multipartite graphs with no sedentary vertex.

Adjacency case

The next result is an analogue of Theorem 19 for the adjacency case.

Theorem 24.

Let X=Kn1,,nkX=K_{n_{1},\ldots,n_{k}} and n=j=1knjn=\sum_{j=1}^{k}n_{j}. If {1,,k}\ell\in\{1,\ldots,k\} and uu is a vertex of XX that belong to the partite set of size nn_{\ell}, then the following hold with respect to A=A(X)A=A(X).

  1. 1.

    Let n=1n_{\ell}=1 for each \ell\in\mathcal{I}, where {1,,k}\varnothing\neq\mathcal{I}\subseteq\{1,\ldots,k\}.

    1. (a)

      If ||=1|\mathcal{I}|=1 and nr=mn_{r}=m for all rr\notin\mathcal{I}, then uu is tightly ((nm1)2(nm1)2+4(n1))(\frac{(n-m-1)^{2}}{(n-m-1)^{2}+4(n-1)})-sedentary at time t=jπ(nm1)2+4(n1)t=\frac{j\pi}{\sqrt{(n-m-1)^{2}+4(n-1)}} for any odd integer jj.

    2. (b)

      Let ||=2|\mathcal{I}|=2 and nr=mn_{r}=m for all rr\notin\mathcal{I}. Let Δ=(nm3)2+8(n2)\Delta=(n-m-3)^{2}+8(n-2).

      1. i.

        If Δ\Delta is not a perfect square, then pretty good state transfer occurs between the two vertices that belong to partite sets of size one.

      2. ii.

        If Δ\Delta is a perfect square and ν2(nm+1)ν2(Δ)\nu_{2}(n-m+1)\neq\nu_{2}(\sqrt{\Delta}), then perfect state transfer occurs between the two vertices that belong to partite sets of size one.

      3. iii.

        If Δ\Delta is a perfect square and ν2(nm+1)=ν2(Δ)\nu_{2}(n-m+1)=\nu_{2}(\sqrt{\Delta}), then XX is tightly sedentary at uu.

    3. (c)

      Let 3||n3\leq|\mathcal{I}|\leq n. Then |UA(t)u,u|12|||U_{A}(t)_{u,u}|\geq 1-\frac{2}{|\mathcal{I}|} for all tt. If we add that nr=mn_{r}=m for all rr\notin\mathcal{I} and let Δ=(nm2||+1)2+4||(n||)\Delta=(n-m-2|\mathcal{I}|+1)^{2}+4|\mathcal{I}|(n-|\mathcal{I}|), then the following hold.

      1. i.

        Let Δ\Delta be a perfect square. If ν2(nm+1)ν2(Δ)\nu_{2}(n-m+1)\neq\nu_{2}(\sqrt{\Delta}), then uu is tightly (12||)(1-\frac{2}{{|\mathcal{I}|}})-sedentary at time t=jπgt=\frac{j\pi}{g}, where g=gcd(nm+1+Δ,nm+1Δ)g=\operatorname{gcd}(n-m+1+\sqrt{\Delta},n-m+1-\sqrt{\Delta}) and jj is odd. Otherwise, uu is tightly CC-sedentary for some C>12||C>1-\frac{2}{{|\mathcal{I}|}}.

      2. ii.

        If Δ\Delta is not a perfect square, then uu is sharply (12||)(1-\frac{2}{{|\mathcal{I}|}})-sedentary.

    4. (d)

      Let ||n4|\mathcal{I}|\leq n-4 and nr=2n_{r}=2 for all rr\notin\mathcal{I} (so n||n-|\mathcal{I}| is even). Let Δ=(n1)2+4||\Delta=(n-1)^{2}+4|\mathcal{I}|.

      1. i.

        If Δ\Delta is not a perfect square, then pretty good state transfer occurs between every pair of vertices in a partite set of size two if and only if n3n\equiv 3 (mod 4).

      2. ii.

        If Δ\Delta is a perfect square, then then perfect state transfer occurs between every pair of vertices in a partite set of size two if and only if ν2(n3)ν2(Δ)\nu_{2}(n-3)\neq\nu_{2}(\sqrt{\Delta}).

      3. iii.

        If either (i) Δ\Delta is not a perfect square and n3n\not\equiv 3 (mod 4), or (ii) Δ\Delta is a perfect square and ν2(n3)=ν2(Δ)\nu_{2}(n-3)=\nu_{2}(\sqrt{\Delta}), then each vertex in a partite set of size two is sedentary.

  2. 2.

    Let n=2n_{\ell}=2 and nr=mn_{r}=m for all rr\neq\ell. Let Δ=(nm2)2+8(n2)\Delta=(n-m-2)^{2}+8(n-2) and vv be a twin of uu.

    1. (a)

      If either (i) Δ\Delta is not a perfect square, (ii) Δ\Delta is a perfect square and ν2(nm2)ν2(Δ)\nu_{2}(n-m-2)\neq\nu_{2}(\sqrt{\Delta}) or (iii) n=m+2n=m+2, then XX is not sedentary at uu. In particular, if (i) holds, then pretty good state transfer occurs between uu and vv. Otherwise, perfect state transfer occurs between them.

    2. (b)

      Let n>m+2n>m+2 and n2=12s(nm2+s)n-2=\frac{1}{2}s(n-m-2+s) for some integer ss such that ν2(nm2)ν2(s)\nu_{2}(n-m-2)\leq\nu_{2}(s). Let d1=(nm2)/gd_{1}=(n-m-2)/g and s1=s/gs_{1}=s/g, where g=gcd(nm2,s)g=\operatorname{gcd}(n-m-2,s).

      1. i.

        If s1=1s_{1}=1, then |UA(t)u,u|1d1+2|U_{A}(t)_{u,u}|\geq\frac{1}{d_{1}+2} for all tt with equality if and only if t=jπgt=\frac{j\pi}{g} for odd jj.

      2. ii.

        If s12s_{1}\geq 2, then |UA(t)u,u|2d1+2s1|U_{A}(t)_{u,u}|\geq\frac{\sqrt{2}}{d_{1}+2s_{1}} for all tt with equality if and only if t=jπgt=\frac{j\pi}{g} for odd jj.

  3. 3.

    If n3n_{\ell}\geq 3, then |UA(t)u,u|12n|U_{A}(t)_{u,u}|\geq 1-\frac{2}{n_{\ell}} for all tt. Moreover, uu is tightly (12n)(1-\frac{2}{n_{\ell}})-sedentary at time t=jπgt=\frac{j\pi}{g}, whenever nr=mn_{r}=m for all rr\neq\ell, Δ=(nnm)2+4n(nn)\Delta=(n-n_{\ell}-m)^{2}+4n_{\ell}(n-n_{\ell}) is a perfect square and ν2(nnm)ν2(Δ)\nu_{2}(n-n_{\ell}-m)\neq\nu_{2}(\sqrt{\Delta}), where g=gcd(nnm+Δ,nnmΔ)g=\operatorname{gcd}(n-n_{\ell}-m+\sqrt{\Delta},n-n_{\ell}-m-\sqrt{\Delta}) and jj is odd. Furthermore, if σu(A(X))\sigma_{u}(A(X)) is a linearly independent set over \mathbb{Q}, then uu is sharply (12n)(1-\frac{2}{n_{\ell}})-sedentary.

Proof.

In (1a), XX can be viewed as a cone over an (nm1)(n-m-1)-regular graph on n1n-1 vertices with apex uu. If XX is a cone on a dd-regular graph on nn vertices, then |UA(X)(t)u,u|d2d2+4n|U_{A(X)}(t)_{u,u}|\geq\frac{d^{2}}{d^{2}+4n} with equality if and only if t=πd2+4nt=\frac{\pi}{\sqrt{d^{2}+4n}} [SedQW]. Thus, (1a) is immediate. In (1b), we have n=(k2)m+2n=(k-2)m+2 and we may write X=K2YX=K_{2}\vee Y, where Y=k2OmY=\bigvee_{k-2}O_{m} is a (nm2)(n-m-2)-regular graph on n2n-2 vertices. Invoking Corollary 12 and Theorems 12(1) and 14(2b) in [Kirkland2023] yields (1b). Now, (1ci) is immediate from Theorem 15(1,2). To prove (1cii), note that X=K||Km,,mX=K_{|\mathcal{I}|}\vee K_{m,\ldots,m}, and so σu(A(X))={1,12(d±Δ)}\sigma_{u}(A(X))=\{-1,\frac{1}{2}(d\pm\sqrt{\Delta})\} by [kirkland2023quantum, Lemma 3(2)], where d=nm1d=n-m-1. Let S={1}S=\{-1\} and a=||1||a=\frac{|\mathcal{I}|-1}{|\mathcal{I}|} in Lemma 9. If a,b,ca,b,c are integers such that a[12(d+Δ)]+b[12(dΔ)]+(1)c=0a[\frac{1}{2}(d+\sqrt{\Delta})]+b[\frac{1}{2}(d-\sqrt{\Delta})]+(-1)c=0, then a=ba=b and c=a(nm1)c=a(n-m-1) because Δ\Delta is not a perfect square. Thus, a+b+c=a(nm+1)=0a+b+c=a(n-m+1)=0 for any integer aa if and only if n=m1n=m-1, a contradiction. Thus, Lemma 9 yields (1cii). Lastly in (1d), we have X=K||CP(n||)X=K_{|\mathcal{I}|}\vee CP(n-|\mathcal{I}|), where n||n-|\mathcal{I}| is even and σw(A(CP(n||)))={0,2,n2}\sigma_{w}(A(CP(n-|\mathcal{I}|)))=\{0,-2,n-2\} for each vertex ww of CP(n||)CP(n-|\mathcal{I}|) [Kirkland2023, Example 2]. Invoking [kirkland2023quantum, Lemma 3(2)], we get σw(A(X))={0,2,12(n3±Δ)}\sigma_{w}(A(X))=\{0,-2,\frac{1}{2}(n-3\pm\sqrt{\Delta})\}. Since ww is strongly cospectral with its twin in CP(n||)CP(n-|\mathcal{I}|), say vv, [kirkland2023quantum, Corollary 20(2)] implies that ww and vv are also strongly cospectral in XX with σuv+(A(X))={2,12(n3±Δ)}\sigma_{uv}^{+}(A(X))=\{-2,\frac{1}{2}(n-3\pm\sqrt{\Delta})\} and σuv(A(X))={0}\sigma_{uv}^{-}(A(X))=\{0\}. To prove (1di), let a,b,ca,b,c be integers such that a[12(n3+Δ)]+b[12(n3Δ)]+(2)c=0a[\frac{1}{2}(n-3+\sqrt{\Delta})]+b[\frac{1}{2}(n-3-\sqrt{\Delta})]+(-2)c=0. Since Δ\Delta is not a perfect square, we get a=ba=b and c=a(n3)2c=\frac{a(n-3)}{2}. In this case, a+b+c=a(2+n32)a+b+c=a\left(2+\frac{n-3}{2}\right) is even for any integer aa if and only if n3n\equiv 3 (mod 4). Applying [Kirkland2023, Theorem 13] yields the desired result for (1di), while (1dii) and (1diii) are resp. immediate from [kirkland2023quantum, Theorem 11(1)] and Corollary 12.

In (2), we have X=O2YX=O_{2}\vee Y, and so a direct application of [Monterde2023, Theorem 36] and [Kirkland2023, Theorem 11(1)] yields the desired result. In particular, the assumption in (2b) is equivalent to Δ\Delta is a perfect square and ν2(nm2)=ν2(Δ)\nu_{2}(n-m-2)=\nu_{2}(\sqrt{\Delta})). Finally, (3) is immediate from Theorem 15(1,2). ∎

Remark 25.

Let X=CP(2k)X=CP(2k). Taking n=2kn=2k, s=2s=2, and m=2m=2 in Theorem 24(2) yields nm2=2k4n-m-2=2k-4 and Δ=2k\sqrt{\Delta}=2k. If kk is even, then ν2(2k4)ν2(2k)\nu_{2}(2k-4)\neq\nu_{2}(2k), and so PST occurs in XX by Theorem 24(2a). If kk is odd, then ν2(2k4)=ν2(2k)\nu_{2}(2k-4)=\nu_{2}(2k) and g=2g=2, so that each uV(X)u\in V(X) is tightly (1k)(\frac{1}{k})-sedentary at t=jπgt=\frac{j\pi}{g} for any odd jj by Theorem 24(2bi). This coincides with Theorem 20 since XX is regular.

We now present an analogue of Corollary 21 for the adjacency case.

Corollary 26.

For all n3n\geq 3, let X=Kn\eX=K_{n}\backslash e with non-adjacent vertices uu and vv.

  1. 1.

    If n=3n=3, then perfect state transfer occurs between uu and vv in XX and the degree two vertex of XX is not sedentary.

  2. 2.

    If n=4n=4, then the two degree three vertices of XX admit pretty good state transfer. Moreover, for all n5n\geq 5, each degree n1n-1 vertex of XX is sharply (12n2)(1-\frac{2}{n-2})-sedentary.

  3. 3.

    For all n4n\geq 4, pretty good state transfer occurs between uu and vv in XX.

Proof.

Note that Δ=n2+2n7\Delta=n^{2}+2n-7 is not a perfect square for all n3n\geq 3. Now, (1) follows from PST in P3P_{3} and Example 3. The first statement in (2) follows from [Kirkland2023, Theorem] and the fact that σw(A(X))={1,12(1±17)}\sigma_{w}(A(X))=\{-1,\frac{1}{2}(1\pm\sqrt{17})\} for a degree three vertex ww in XX, while the second statement is obtained by applying Theorem 24(1c) with ||=n2|\mathcal{I}|=n-2 and m=2m=2. Applying Theorem 24(2a) with m=1m=1 yields (3). ∎

6 Threshold graphs

A connected threshold graph is a graph that is either of the form:

  • ((((Om1Km2)Om3)Km4))Km2k:=Γ(m1,,m2k)((((O_{m_{1}}\vee K_{m_{2}})\cup O_{m_{3}})\vee K_{m_{4}})\ldots)\vee K_{m_{2k}}:=\Gamma(m_{1},\ldots,m_{2k})

  • ((((Km1Om2)Km3)Om4))Km2k+1:=Γ(m1,,m2k+1)((((K_{m_{1}}\cup O_{m_{2}})\vee K_{m_{3}})\cup O_{m_{4}})\ldots)\vee K_{m_{2k+1}}:=\Gamma(m_{1},\ldots,m_{2k+1}).

PST and fractional revival in threshold graphs under Laplacian dynamics have been investigated in [Severini] and [Kirkland2020], respectively. Here, we characterize Laplacian sedentariness in this family of graphs.

Theorem 27.

Let X{Γ(m1,,m2k),Γ(m1,,m2k+1)}X\in\{\Gamma(m_{1},\ldots,m_{2k}),\Gamma(m_{1},\ldots,m_{2k+1})\}. If m1=2m_{1}=2, m22m_{2}\equiv 2 (mod 4) and mj0m_{j}\equiv 0 (mod 4) for all j1,2j\neq 1,2, then perfect state transfer occurs between the two vertices of X1{Om1,Km1}X_{1}\in\{O_{m_{1}},K_{m_{1}}\}. Otherwise, each vertex in XX is sedentary.

Proof.

The first statement follows from [Severini, Theorem 2]. The second follows from Corollary 13, L(X)L(X) having integer eigenvalues and the absence of strongly cospectrality in XX [kirkland2023quantum, Corollary 67]. ∎

Next, we provide bounds on the sedentariness of a vertex in a threshold graph.

Theorem 28.

Let X{Γ(m1,,m2k),Γ(m1,,m2k+1)}X\in\{\Gamma(m_{1},\ldots,m_{2k}),\Gamma(m_{1},\ldots,m_{2k+1})\}. Suppose j{1,,2k+1}j\in\{1,\ldots,2k+1\} and uu is a vertex of Yj{Omj,Kmj}Y_{j}\in\{O_{m_{j}},K_{m_{j}}\}. Let αj=r=1jmr\alpha_{j}=\sum_{r=1}^{j}m_{r}. The following hold.

  1. 1.

    Let (a) jj be even, X=Γ(m1,,m2k)X=\Gamma(m_{1},\ldots,m_{2k}) and h=2kh=2k or (b) j3j\geq 3 be odd, X=Γ(m1,,m2k+1)X=\Gamma(m_{1},\ldots,m_{2k+1}) and h=2k+1h=2k+1. For each integer \ell having the same parity as hh such that j+2hj+2\leq\ell\leq h, define β,h:=m+m+2++mh\beta_{\ell,h}:=m_{\ell}+m_{\ell+2}+\ldots+m_{h}. Then

    |UL(X)(t)u,u|12r=jh(1)r+hαrfor all t,\displaystyle|U_{L(X)}(t)_{u,u}|\geq 1-2\sum_{r=j}^{h}\frac{(-1)^{r+h}}{\alpha_{r}}\quad\text{for all $t$},

    with equality if and only if for some tt, eitαh=1e^{it\alpha_{h}}=-1 and eitβ,h=1e^{it\beta_{\ell,h}}=1 for each even j+2hj+2\leq\ell\leq h.

  2. 2.

    Let j=1j=1, m12m_{1}\geq 2 and X=Γ(m1,,m2k+1)X=\Gamma(m_{1},\ldots,m_{2k+1}). For each odd 32k+13\leq\ell\leq 2k+1, define β=r=0(2k+1)/2m+2r=m+m+2++m2k+1\beta_{\ell}=\sum_{r=0}^{(2k+1-\ell)/2}m_{\ell+2r}=m_{\ell}+m_{\ell+2}+\ldots+m_{2k+1}. Then

    |UL(X)(t)u,u|12/m1for all t,|U_{L(X)}(t)_{u,u}|\geq 1-2/m_{1}\quad\text{for all $t$}, (9)

    with equality if and only if for some tt, eit(m1+β3)=1e^{it(m_{1}+\beta_{3})}=-1, eitα2k+1=eit(α+β+2)=1e^{it\alpha_{2k+1}}=e^{it(\alpha_{\ell}+\beta_{\ell+2})}=1 for each odd 32k13\leq\ell\leq 2k-1, and eitβ+2=1e^{it\beta_{\ell+2}}=1 for each odd 12k11\leq\ell\leq 2k-1.

  3. 3.

    Let (a) jj be odd, X=Γ(m1,,m2k)X=\Gamma(m_{1},\ldots,m_{2k}) and h=2kh=2k or (b) jj be even, X=Γ(m1,,m2k+1)X=\Gamma(m_{1},\ldots,m_{2k+1}) and h=2k+1h=2k+1. For each integer \ell that has the same parity as hh such that jhj\leq\ell\leq h, define β=r=0(2k)/2m+2r=m+m+2++mh\beta_{\ell}=\sum_{r=0}^{(2k-\ell)/2}m_{\ell+2r}=m_{\ell}+m_{\ell+2}+\ldots+m_{h}. Then

    |UL(X)(t)u,u|12/αjfor all t,|U_{L(X)}(t)_{u,u}|\geq 1-2/\alpha_{j}\quad\text{for all $t$}, (10)

    with equality if and only if for some tt, eitβj+1=1e^{it\beta_{j+1}}=-1, eitαh=eitβ=1e^{it\alpha_{h}}=e^{it\beta_{\ell}}=1 for each j+2hj+2\leq\ell\leq h and eit(α+β+2)=1e^{it(\alpha_{\ell}+\beta_{\ell+2})}=1 for each jhj\leq\ell\leq h, where \ell has the same parity as hh.

Proof.

We first prove (1). For each even jh2kj\leq h\leq 2k, let Xh=((((Om1Km2)Om3)Km4))KmhX_{h}=((((O_{m_{1}}\vee K_{m_{2}})\cup O_{m_{3}})\vee K_{m_{4}})\ldots)\vee K_{m_{h}} (so that X2k=XX_{2k}=X). Since σu(L(Kmj))={0,mj}\sigma_{u}(L(K_{m_{j}}))=\{0,m_{j}\}, repeatedly applying [kirkland2023quantum, Lemma 3(2)] (hj)/2(h-j)/2 times starting from KmjK_{m_{j}} to XjX_{j} all the way to XhX_{h} gives us

σu(L(Xh))={β,h:j+2h}{0,αh}.\sigma_{u}(L(X_{h}))=\{\beta_{\ell,h}:j+2\leq\ell\leq h\}\cup\{0,\alpha_{h}\}. (11)

Since [mh+21,αh1]T[m_{h+2}\textbf{1},-\alpha_{h}\textbf{1}]^{T} is an eigenvector for L(Xh+2)L(X_{h+2}) associated with αh+2\alpha_{h+2}, the multiplicity of αh+2\alpha_{h+2} is equal to that of αh\alpha_{h} plus one. Thus, (Eαh+2)u,u=(Eαh)u,u+(1αh+11αh+2)(E_{\alpha_{h+2}})_{u,u}=(E_{\alpha_{h}})_{u,u}+(\frac{1}{\alpha_{h+1}}-\frac{1}{\alpha_{h+2}}). Arguing inductively, we get (Eα2k)u,u=1r=j2k(1)rαr(E_{\alpha_{2k}})_{u,u}=1-\sum_{r=j}^{2k}\frac{(-1)^{r}}{\alpha_{r}} in L(X)L(X). Taking S={α2k}S=\{\alpha_{2k}\} and a=(Eα2k)u,ua=(E_{\alpha_{2k}})_{u,u} in Theorem 4 yields (1a). For (1b), the same argument applied to Xh=((((Km1Om2)Km3)Om4))KmhX_{h}=((((K_{m_{1}}\cup O_{m_{2}})\vee K_{m_{3}})\cup O_{m_{4}})\ldots)\vee K_{m_{h}} for each odd jh2k+1j\leq h\leq 2k+1 yields the desired result.

To prove (2), note that α1=m1\alpha_{1}=m_{1}. The same argument used in (11) yields

σu(L(X))={β:32k+1is odd}{α+β+2:12k1is odd}{0,α2k+1}\sigma_{u}(L(X))=\{\beta_{\ell}:3\leq\ell\leq 2k+1\ \text{is odd}\}\cup\{\alpha_{\ell}+\beta_{\ell+2}:1\leq\ell\leq 2k-1\ \text{is odd}\}\cup\{0,\alpha_{2k+1}\}.

Note that σu(L(Km1))={0,α1}\sigma_{u}(L(K_{m_{1}}))=\{0,\alpha_{1}\} and α1+β3,hσu(L(Xh))\alpha_{1}+\beta_{3,h}\in\sigma_{u}(L(X_{h})) for each odd 3h2k+13\leq h\leq 2k+1, where β3,h\beta_{3,h} is defined in (1) and Xh=((((Km1Om2)Km3)Om4))KmhX_{h}=((((K_{m_{1}}\cup O_{m_{2}})\vee K_{m_{3}})\cup O_{m_{4}})\ldots)\vee K_{m_{h}}. Since Eα1=11α1E_{\alpha_{1}}=1-\frac{1}{\alpha_{1}} and the multiplicity of α1+β3,hσu(L(Xh))\alpha_{1}+\beta_{3,h}\in\sigma_{u}(L(X_{h})) is equal to that of α1\alpha_{1}, we get (Eα1+β3)u,u=Eα1=11α1(E_{\alpha_{1}+\beta_{3}})_{u,u}=E_{\alpha_{1}}=1-\frac{1}{\alpha_{1}}. Taking S={α1+β3}S=\{\alpha_{1}+\beta_{3}\} and a=(Eα1+β3)u,u=11α1a=(E_{\alpha_{1}+\beta_{3}})_{u,u}=1-\frac{1}{\alpha_{1}} in Theorem 4 yields (2).

To prove (3a), suppose jj is odd and X=Γ(m1,,m2k)X=\Gamma(m_{1},\ldots,m_{2k}). The same argument used in (11) yields

σu(L(X))={β:j+12kis even}{α+β+2:j+12k2is even}{0,α2k}.\sigma_{u}(L(X))=\{\beta_{\ell}:j+1\leq\ell\leq 2k\ \text{is even}\}\cup\{\alpha_{\ell}+\beta_{\ell+2}:j+1\leq\ell\leq 2k-2\ \text{is even}\}\cup\{0,\alpha_{2k}\}.

For each odd jh2kj\leq h\leq 2k, let Xh=(((((Om1Km2)Om3)Km4))Omh)Kmh+1X_{h}=(((((O_{m_{1}}\vee K_{m_{2}})\cup O_{m_{3}})\vee K_{m_{4}})\ldots)\cup O_{m_{h}})\vee K_{m_{h+1}}. Note that σu(L(Omj))={0}\sigma_{u}(L(O_{m_{j}}))=\{0\} and βj+1,hσu(L(Xh))\beta_{j+1,h}\in\sigma_{u}(L(X_{h})) for each odd jh2kj\leq h\leq 2k, where βj+1,h\beta_{j+1,h} is defined in (1). From the form of the XhX_{h}’s, we deduce that the multiplicity of βj+1,h+2σu(L(Xh+2))\beta_{j+1,h+2}\in\sigma_{u}(L(X_{h+2})) is equal to that of βj+1,hσu(L(Xh))\beta_{j+1,h}\in\sigma_{u}(L(X_{h})). Since (Eβj+1,j)u,u=(11mj)+(1mj1αj)=11αj(E_{\beta_{j+1,j}})_{u,u}=\left(1-\frac{1}{m_{j}}\right)+\left(\frac{1}{m_{j}}-\frac{1}{\alpha_{j}}\right)=1-\frac{1}{\alpha_{j}} in L(Xj)L(X_{j}), we get (Eβj+1,h)u,u=11αj(E_{\beta_{j+1,h}})_{u,u}=1-\frac{1}{\alpha_{j}} in L(Xh)L(X_{h}) for each odd jh2kj\leq h\leq 2k. Letting h=2k1h=2k-1, S={βj+1}S=\{\beta_{j+1}\} and a=(Eβj+1)u,u=11αja=(E_{\beta_{j+1}})_{u,u}=1-\frac{1}{\alpha_{j}} in Theorem 4 yields the desired conclusion in (3a). The same argument works for (3b). ∎

Remark 29.

If m1=1m_{1}=1 and X=Γ(m1,,m2k)X=\Gamma(m_{1},\ldots,m_{2k}), then X=Γ(m1,,m2k1)X=\Gamma(m_{1}^{\prime},\ldots,m_{2k-1}^{\prime}), where m1=m2+1m_{1}^{\prime}=m_{2}+1 and mj=mj+1m_{j}^{\prime}=m_{j+1} for all j1j\geq 1. In this case, Theorem 28(2) applies. Similarly, if m1=1m_{1}=1 and X=Γ(m1,,m2k+1)X=\Gamma(m_{1},\ldots,m_{2k+1}), then Theorem 28(3) applies.

Remark 30.

Let m1=2m_{1}=2 so that α1=2\alpha_{1}=2 in Theorem 28(2-3). If m22m_{2}\equiv 2 and mr0m_{r}\equiv 0 (mod 4) for all r1,2r\neq 1,2, then (9) and (10) yields UL(X)(π2)u,u=0U_{L(X)}(\frac{\pi}{2})_{u,u}=0 for X{Γ(m1,,m2k+1),Γ(m1,,m2k)}X\in\{\Gamma(m_{1},\ldots,m_{2k+1}),\Gamma(m_{1},\ldots,m_{2k})\}. This is consistent with the fact that uu is involved in PST at t=π2t=\frac{\pi}{2} by Theorem 27. However, if m22m_{2}\not\equiv 2 or mr0m_{r}\not\equiv 0 (mod 4) for some r1,2r\neq 1,2, then there is no time tt such that the lower bounds in (9) and (10) are attained respectively. In this case, uu is sedentary in XX.

Lastly, we have the following asymptotic result that applies to MqM_{q} for all qq.

Corollary 31.

Almost all complete multipartite graphs and threshold graphs contain a sedentary vertex.

Proof.

The number of non-isomorphic complete bipartite graphs on nn vertices is equal to P(n)P(n), the number of partitions of nn (where the partitions represent the sizes of the partite sets). Similarly, the number of non-isomorphic complete bipartite graphs on nn vertices which contain a partite set of size three is equal to P(n3)P(n-3). From Theorem 15(2), we know that each vertex in a partite set of size three is sedentary. Using the famous asymptotic formula P(n)14n3eπ2n/3P(n)\sim\frac{1}{4n\sqrt{3}}e^{\pi\sqrt{2n/3}} as nn\rightarrow\infty due to Hardy and Ramanujan [HardyAsymptoticFI], it follows that P(n3)/P(n)1P(n-3)/P(n)\rightarrow 1 as nn\rightarrow\infty. The same argument applies to threshold graphs. ∎

7 Direct Product

In [Monterde2023, Theorem 4], it was shown that the Cartesian product preserves sedentariness. In this section, we investigate whether the direct product also preserves sedentariness relative to the adjacency matrix.

The direct product X×YX\times Y of XX and YY is the graph with vertex set V(X)×V(Y)V(X)\times V(Y) and adjacency matrix

A(X)A(Y).A(X)\otimes A(Y).

If A(X)=jλjEjA(X)=\sum_{j}\lambda_{j}E_{j} is a spectral decomposition of A(X)A(X), then it was shown in [coutinho2016perfect, Lemma 4.2] that

UA(X×Y)(t)=jEjUA(Y)(λjt).U_{A(X\times Y)}(t)=\sum_{j}E_{j}\otimes U_{A(Y)}(\lambda_{j}t). (12)

Since A(Km)=(m1)(1mJm)+(1)(Im1mJm)A(K_{m})=(m-1)(\frac{1}{m}\textbf{J}_{m})+(-1)(I_{m}-\frac{1}{m}\textbf{J}_{m}), (12) yields the following transition matrix for Km×YK_{m}\times Y

UA(Km×Y)(t)=1mJmUA(Y)((m1)t)+(Im1mJm)UA(Y)(t).\begin{split}U_{A(K_{m}\times Y)}(t)&=\frac{1}{m}\textbf{J}_{m}\otimes U_{A(Y)}((m-1)t)+(I_{m}-\frac{1}{m}\textbf{J}_{m})\otimes U_{A(Y)}(-t).\end{split} (13)

Recall that K2×YK_{2}\times Y is called the bipartite double of YY, and this graph is connected if and only if YY is non-bipartite. Letting m=2m=2 and taking the diagonal entries in (13) gives us

|UA(K2×Y)(t)((u,v),(u,v)|=|Re(UA(Y)(t)v,v)|.\begin{split}\left|U_{A(K_{2}\times Y)}(t)_{((u,v),(u,v)}\right|=|\operatorname{Re}(U_{A(Y)}(t)_{v,v})|.\end{split} (14)
Theorem 32.

Let uV(Km)u\in V(K_{m}) and vV(Y)v\in V(Y). The following hold.

  1. 1.

    If m3m\geq 3 and YY is CC-sedentary at vertex vv, where C>1m1C>\frac{1}{m-1}, then KmYK_{m}\otimes Y is (CC+1m)(C-\frac{C+1}{m})-sedentary at vertex (u,v)(u,v) for any vertex uu of KmK_{m}.

  2. 2.

    K2×YK_{2}\times Y is CC-sedentary at (u,v)(u,v) if and only if |Re(UA(Y)(t)v,v)|C|\operatorname{Re}(U_{A(Y)}(t)_{v,v})|\geq C for all tt.

Proof.

Taking the ((u,v),(u,v))((u,v),(u,v))-entry of (13) and applying triangle inequality yields

|UA(Km×Y)(t)((u,v),(u,v)|=|1mUA(Y)((m1)t)v,v+m1mUA(Y)(t)v,v|m1m|UA(Y)(t)v,v|1m|UA(Y)((m1)t)v,v|m1m|UA(Y)(t)v,v|1m.\begin{split}\left|U_{A(K_{m}\times Y)}(t)_{((u,v),(u,v)}\right|&=\left|\frac{1}{m}U_{A(Y)}((m-1)t)_{v,v}+\frac{m-1}{m}U_{A(Y)}(-t)_{v,v}\right|\\ &\geq\frac{m-1}{m}\left|U_{A(Y)}(-t)_{v,v}\right|-\frac{1}{m}\left|U_{A(Y)}((m-1)t)_{v,v}\right|\\ &\geq\frac{m-1}{m}\left|U_{A(Y)}(t)_{v,v}\right|-\frac{1}{m}.\end{split}

Therefore, if |UA(Y)(t)v,v|C\left|U_{A(Y)}(t)_{v,v}\right|\geq C for all tt, then |UA(KmY)(t)((u,v),(u,v)|C(m1)m1m=CC+1m\left|U_{A(K_{m}\vee Y)}(t)_{((u,v),(u,v)}\right|\geq\frac{C(m-1)}{m}-\frac{1}{m}=C-\frac{C+1}{m}. From this, (1) is immediate, and (2) follows from (14). ∎

Example 33.

Let TT be a twin set in YY. If |T|3|T|\geq 3 and C=12|T|>1m1C=1-\frac{2}{|T|}>\frac{1}{m-1}, then Theorems 15(2) and 32(1) imply that KmYK_{m}\otimes Y is (CC+1m)(C-\frac{C+1}{m})-sedentary at (u,v)(u,v) for each uV(Km)u\in V(K_{m}) and each vTv\in T.

Next, we examine direct products of complete graphs. Let [n]={1,2,,n}[n]=\{1,2,\ldots,n\}. Denote Y1×Y2××YnY_{1}\times Y_{2}\times\ldots\times Y_{n} by j[n]Yj\bigtimes_{j\in[n]}Y_{j} and the direct product of nn copies of YY by Y×nY^{\times n}. For the graph Y1×Y2××YnY_{1}\times Y_{2}\times\ldots\times Y_{n}, the ordering of indices 1,2,,n1,2,\ldots,n does not matter since the direct product is commutative up to isomorphism. In our next result, we determine the diagonal entries of the transition matrix of j[n]Kmj\bigtimes_{j\in[n]}K_{m_{j}}.

Theorem 34.

Let n2n\geq 2. For each j[n]j\in[n], let mj2m_{j}\geq 2 be an integer and consider X=j[n]KmjX=\bigtimes_{j\in[n]}K_{m_{j}}. Then

UX(t)((v1,,vn),(v1,,vn))=1j[n]mjS[n](jS(mj1))eit(1)|S|jS(mj1),U_{X}(t)_{((v_{1},\ldots,v_{n}),(v_{1},\ldots,v_{n}))}=\frac{1}{\prod_{j\in[n]}m_{j}}\displaystyle\sum_{S\subseteq[n]}\left(\prod_{j\in S}(m_{j}-1)\right)e^{it(-1)^{|S|}\prod_{j\notin S}(m_{j}-1)}, (15)

where jS(mj1)=1\prod_{j\in S}(m_{j}-1)=1 whenever S=S=\varnothing. Moreover, the following hold.

  1. 1.

    If mj=2m_{j}=2 for some j[n]j\in[n], then

    UX(t)((v1,,vn),(v1,,vn))=1j[n]\{j}mjS[n]\{j}(jS(mj1))cos(tjS(mj1)).U_{X}(t)_{((v_{1},\ldots,v_{n}),(v_{1},\ldots,v_{n}))}=\frac{1}{\prod_{j\in[n]\backslash\{j\}}m_{j}}\displaystyle\sum_{S\subseteq[n]\backslash\{j\}}\left(\prod_{j\in S}(m_{j}-1)\right)\cos\left(t\prod_{j\notin S}(m_{j}-1)\right). (16)

    Thus, if there is a time t1t_{1} such that cos(t1jS(mj1))<0\cos\left(t_{1}\prod_{j\notin S}(m_{j}-1)\right)<0 for each S[n]\{j}S\subseteq[n]\backslash\{j\}, then each vertex of XX is not sedentary. In particular, if each mjm_{j} is even, then each vertex of XX is not sedentary.

  2. 2.

    If j[n](mj1)12j[n]mj\displaystyle\prod_{j\in[n]}(m_{j}-1)\neq\frac{1}{2}\prod_{j\in[n]}m_{j}, then each vertex of XX is 2j[n]mj|j[n](mj1)12j[n]mj|\displaystyle\frac{2}{\prod_{j\in[n]}m_{j}}\bigg{|}\prod_{j\in[n]}(m_{j}-1)-\frac{1}{2}\prod_{j\in[n]}m_{j}\bigg{|}-sedentary, and this is tight at time t=πt=\pi whenever each mjm_{j} is odd. In particular, if m2m\geq 2, then each vertex of Km×nK_{m}^{\times n} is 2mn|(m1)n12mn|\displaystyle\frac{2}{m^{n}}\bigg{|}(m-1)^{n}-\frac{1}{2}m^{n}\bigg{|}-sedentary.

Proof.

We prove (15) by induction. The base case follows from the fact that UA(Kn)(t)u,u=eit(n1)(1nJn)+eit(In1nJn)U_{A(K_{n})}(t)_{u,u}=e^{it(n-1)}(\frac{1}{n}\textbf{J}_{n})+e^{-it}(I_{n}-\frac{1}{n}\textbf{J}_{n}). For the induction step, let k1k\geq 1 be an integer and suppose (15) holds whenever n=kn=k. Let Y=Kmk+1×j[k]KmjY=K_{m_{k+1}}\times\bigtimes_{j\in[k]}K_{m_{j}}. Applying (13) to YY, and simplifying the resulting transition matrix yields

UY(t)((v1,,vn),(v1,,vn))=1mk+1[1j[k]mjS[k](jS(mj1))eit(1)|S|jS(mj1)(mk+11)]+mk+11mk+1[1j[k]mjS[k](jS(mj1))eit(1)|S|+1jS(mj1)]=1j[k+1]mjS[k](jS(mj1))eit(1)|S|jS(mj1)(mk+11)+1j[k+1]mjS{k+1}[k+1](jS{k+1}(mj1))eit(1)|S|+1jS(mj1).\begin{split}U_{Y}(t)_{((v_{1},\ldots,v_{n}),(v_{1},\ldots,v_{n}))}&=\frac{1}{m_{k+1}}\left[\frac{1}{\prod_{j\in[k]}m_{j}}\displaystyle\sum_{S\subseteq[k]}\left(\prod_{j\in S}(m_{j}-1)\right)e^{it(-1)^{|S|}\prod_{j\notin S}(m_{j}-1)(m_{k+1}-1)}\right]\\ &+\frac{m_{k+1}-1}{m_{k+1}}\left[\frac{1}{\prod_{j\in[k]}m_{j}}\displaystyle\sum_{S\subseteq[k]}\left(\prod_{j\in S}(m_{j}-1)\right)e^{it(-1)^{|S|+1}\prod_{j\notin S}(m_{j}-1)}\right]\\ &=\frac{1}{\prod_{j\in[k+1]}m_{j}}\displaystyle\sum_{S\subseteq[k]}\left(\prod_{j\in S}(m_{j}-1)\right)e^{it(-1)^{|S|}\prod_{j\notin S}(m_{j}-1)(m_{k+1}-1)}\\ &+\frac{1}{\prod_{j\in[k+1]}m_{j}}\displaystyle\sum_{S\cup\{k+1\}\subseteq[k+1]}\left(\prod_{j\in S\cup\{k+1\}}(m_{j}-1)\right)e^{it(-1)^{|S|+1}\prod_{j\notin S}(m_{j}-1)}.\end{split}

Observe that the first summand in the second equality above is a summation that runs over all subsets of [k+1][k+1] that do not contain k+1k+1, whereas the second summand runs over all subsets of [k+1][k+1] that contain k+1k+1. Thus, combining the two summands yields (15) with n=k+1n=k+1. This establishes (15) for all n1n\geq 1.

To prove (1), we may without loss of generality assume that mn=2m_{n}=2. Making use of the last equality above with n=k+1n=k+1 yields

UX(t)((v1,,vn),(v1,,vn))=12j[n1]mjS[n1](jS(mj1))eit(1)|S|jS(mj1)+12j[n1]mjS[n1](jS(mj1))eit(1)|S|+1jS(mj1)\begin{split}U_{X}(t)_{((v_{1},\ldots,v_{n}),(v_{1},\ldots,v_{n}))}&=\frac{1}{2\prod_{j\in[n-1]}m_{j}}\displaystyle\sum_{S\subseteq[n-1]}\left(\prod_{j\in S}(m_{j}-1)\right)e^{it(-1)^{|S|}\prod_{j\notin S}(m_{j}-1)}\\ &+\frac{1}{2\prod_{j\in[n-1]}m_{j}}\displaystyle\sum_{S\subseteq[n-1]}\left(\prod_{j\in S}(m_{j}-1)\right)e^{it(-1)^{|S|+1}\prod_{j\notin S}(m_{j}-1)}\\ \end{split}

Adding the two summands above yields (16). Now, note that UX(0)((v1,,vn),(v1,,vn))>0U_{X}(0)_{((v_{1},\ldots,v_{n}),(v_{1},\ldots,v_{n}))}>0 from (16). Thus, if cos(t1jS(mj1))<0\cos\left(t_{1}\prod_{j\notin S}(m_{j}-1)\right)<0 for each S[n1]S\subseteq[n-1], then UX(t1)((v1,,vn),(v1,,vn))<0U_{X}(t_{1})_{((v_{1},\ldots,v_{n}),(v_{1},\ldots,v_{n}))}<0. Invoking the intermediate value theorem, there exists t0>0t_{0}>0 such that UY(t0)((v1,,vn),(v1,,vn))=0U_{Y}(t_{0})_{((v_{1},\ldots,v_{n}),(v_{1},\ldots,v_{n}))}=0, and so each vertex of XX is not sedentary by Theorem 32(2). In particular, if mjm_{j} is even for each j[n]j\in[n], then we may take t1=πt_{1}=\pi, and so the conclusion applies.

Finally, we prove (2). Rewriting (15) and applying triangle inequality gives us

|UX(t)((v1,,vn),(v1,,vn))|=1j[n]mj|j[n](mj1)eit(1)n+S[n](jS(mj1))eit(1)|S|jS(mj1)|=1j[n]mj|j[n](mj1)+S[n](jS(mj1))eit(1)|S|(jS(mj1)(1)n|S|)|1j[n]mj|j[n](mj1)S[n](jS(mj1))|=2j[n]mj|j[n](mj1)12j[n]mj|\begin{split}|U_{X}(t)_{((v_{1},\ldots,v_{n}),(v_{1},\ldots,v_{n}))}|&=\frac{1}{\prod_{j\in[n]}m_{j}}\left|\prod_{j\in[n]}(m_{j}-1)e^{it(-1)^{n}}+\displaystyle\sum_{S\subsetneq[n]}\left(\prod_{j\in S}(m_{j}-1)\right)e^{it(-1)^{|S|}\prod_{j\notin S}(m_{j}-1)}\right|\\ &=\frac{1}{\prod_{j\in[n]}m_{j}}\left|\prod_{j\in[n]}(m_{j}-1)+\displaystyle\sum_{S\subsetneq[n]}\left(\prod_{j\in S}(m_{j}-1)\right)e^{it(-1)^{|S|}\left(\prod_{j\notin S}(m_{j}-1)-(-1)^{n-|S|}\right)}\right|\\ &\geq\frac{1}{\prod_{j\in[n]}m_{j}}\left|\prod_{j\in[n]}(m_{j}-1)-\displaystyle\sum_{S\subsetneq[n]}\left(\prod_{j\in S}(m_{j}-1)\right)\right|\\ &=\frac{2}{\prod_{j\in[n]}m_{j}}\left|\prod_{j\in[n]}(m_{j}-1)-\frac{1}{2}\displaystyle\prod_{j\in[n]}m_{j}\right|\end{split}

for all tt, where the last equality above uses the fact that j[n]mj=j[n](mj1)+S[n](jS(mj1))\displaystyle\prod_{j\in[n]}m_{j}=\prod_{j\in[n]}(m_{j}-1)+\sum_{S\subsetneq[n]}\left(\prod_{j\in S}(m_{j}-1)\right). Moreover, the above inequality is an equality at time t=πt=\pi whenever mjm_{j} is odd for each j[n]j\in[n]. From these considerations, statement (2) is immediate. ∎

Corollary 35.

Let uV(Km)u\in V(K_{m}) and vV(Kn)v\in V(K_{n}). The following hold in Km×KnK_{m}\times K_{n}.

  1. 1.

    If m=2m=2 or n=2n=2, then (u,v)(u,v) is not sedentary.

  2. 2.

    If m,n3m,n\geq 3, then (u,v)(u,v) is sedentary. In particular, if (m,n)(3,4)(m,n)\neq(3,4), then (u,v)(u,v) is (|mn2(m+n1)mn|)(|\frac{mn-2(m+n-1)}{mn}|)-sedentary and this is tight whenever mm and nn are odd. Otherwise, (u,v)(u,v) is tightly 0.1420.142-sedentary.

Proof.

Let m=2m=2. By (16), UA(K2×Kn)(t)((u,v),(u,v)=1ncos(t(n1))+n1ncostU_{A(K_{2}\times K_{n})}(t)_{((u,v),(u,v)}=\frac{1}{n}\cos(t(n-1))+\frac{n-1}{n}\cos t. If n=2n=2, then K2×K2K_{2}\times K_{2} is disconnected and each vertex pairs up with another to admit PST. Now, suppose n3n\geq 3. Since 1ncos(t(n1))+n1ncost<0\frac{1}{n}\cos(t(n-1))+\frac{n-1}{n}\cos t<0 at t=πt=\pi, we conclude that each vertex of K2×KnK_{2}\times K_{n} is not sedentary by Theorem 34(1). Thus, (1) holds. To prove (2), note that the case (m,n)(3,4)(m,n)\neq(3,4) is immediate from Theorem 34(2). Now, if (m,n)=(3,4)(m,n)=(3,4), then we get |UA(K3×K4)(t)((u,v),(u,v)|=112|e5it+3e3it+2e4it+6|\left|U_{A(K_{3}\times K_{4})}(t)_{((u,v),(u,v)}\right|=\frac{1}{12}\left|e^{5it}+3e^{-3it}+2e^{-4it}+6\right|. Using your favourite computer package, you may verify that |UA(K3×K4)(t)((u,v),(u,v)|0.142\left|U_{A(K_{3}\times K_{4})}(t)_{((u,v),(u,v)}\right|\geq 0.142 with equality whenever t{0.945,2π0.945}t\in\{0.945,2\pi-0.945\}. Thus, each vertex of Km×KnK_{m}\times K_{n} is sedentary for all m,n3m,n\geq 3. ∎

Remark 36.

Consider X=K2×KnX=K_{2}\times K_{n} and let V(K2)={1,2}V(K_{2})=\{1,2\}. By Corollary 35(1), XX is not sedentary at every vertex. Now, for each uV(Kn)u\in V(K_{n}), one checks that vertex (1,u)(1,u) is strongly cospectral only to vertex (2,u)(2,u) in XX. Invoking [Coutinho2014, Theorem 3.3.4], we obtain PST between (1,u)(1,u) and (2,u)(2,u) in XX if and only if nn is even. Consequently, each vertex of K2×KnK_{2}\times K_{n} for each odd nn is periodic, sedentary, and is involved in strong cospectrality but not PST. This also applies to the Laplacian case because XX is regular.

To complement Corollary 35(1), we provide an infinite family of graphs of the form K2×YK_{2}\times Y that contain sedentary vertices, where YY is non-bipartite.

Theorem 37.

Let ZZ be a dd-regular graph on nn vertices such that d>0d>0 is an integer and n=12s(d+s)n=\frac{1}{2}s(d+s) for some even integer ss satisfying ν2(s)ν2(d)\nu_{2}(s)\geq\nu_{2}(d). If vv is an apex of Y=O2ZY=O_{2}\vee Z, then K2×YK_{2}\times Y is tightly CC-sedentary at vertex (u,v)(u,v) for any vertex uu of KmK_{m}, where

C=mink𝒦cos2(s1kπd1+2s1)>0C=\min_{k\in\mathcal{K}}\cos^{2}\left(\frac{s_{1}k\pi}{d_{1}+2s_{1}}\right)>0,

𝒦={k+:j(d1+2s1)4s1k(j+2)(d1+2s1)4s1,js1,j1 (mod 4)}\mathcal{K}=\left\{k\in\mathbb{Z}^{+}:\lceil\frac{j(d_{1}+2s_{1})}{4s_{1}}\rceil\leq k\leq\lfloor\frac{(j+2)(d_{1}+2s_{1})}{4s_{1}}\rfloor,\ j\leq s_{1},\ \text{$j\equiv 1$ (mod 4)}\right\}, d1=dgcd(d,s)d_{1}=\frac{d}{\operatorname{gcd}(d,s)} and s1=sgcd(d,s)s_{1}=\frac{s}{\operatorname{gcd}(d,s)}. In particular, |UA(K2×Y(τ)((u,v),(u,v)|=C\left|U_{A(K_{2}\times Y}(\tau)_{((u,v),(u,v)}\right|=C at time τ=2k0πd+2s\tau=\frac{2k_{0}\pi}{d+2s}, where k0𝒦k_{0}\in\mathcal{K} such that C=cos2(s1k0πd1+2s1)C=\cos^{2}\left(\frac{s_{1}k_{0}\pi}{d_{1}+2s_{1}}\right).

Proof.

From the proof of [Monterde2023, Theorem 36], UA(Y)(t)u,u=12(d+2s)((d+2s)+seit(d+s)+(d+s)eits)U_{A(Y)}(t)_{u,u}=\frac{1}{2(d+2s)}\left((d+2s)+se^{it(d+s)}+(d+s)e^{-its}\right). From this, we obtain

Re(UA(Y)(t)v,v)=12(d+2s)((d+2s)+f(t)),\operatorname{Re}\left(U_{A(Y)}(t)_{v,v}\right)=\frac{1}{2(d+2s)}\left((d+2s)+f(t)\right), (17)

where f(t)=scos(t(d+s))+(d+s)cos(ts)f(t)=s\cos(t(d+s))+(d+s)\cos(ts). Note that f(t)=s(d+s)(sin(t(d+s)+sin(ts))f^{\prime}(t)=-s(d+s)\left(\sin(t(d+s)+\sin(ts)\right) and f′′(t)=s(d+s)((d+s)cos(t(d+s)+scos(ts))f^{\prime\prime}(t)=-s(d+s)\left((d+s)\cos(t(d+s)+s\cos(ts)\right). Using a sum to product identity, we obtain f(t)=2s(d+s)(sin(t(d+2s)/2)cos(td/2))f^{\prime}(t)=-2s(d+s)\left(\sin(t(d+2s)/2)\cos(td/2)\right). Therefore f(t)=0f^{\prime}(t)=0 if and only if either t=2kπd+2st=\frac{2k\pi}{d+2s} for any integer kk or t=π/dt=\ell\pi/d for any odd integer \ell.

  • Let t=2kπd+2st=\frac{2k\pi}{d+2s}, where 0<kd+2s0<k\leq d+2s is an integer. Then cos(ts)=cos(t(d+s))=cos(2s1kπd1+2s1)\cos(ts)=\cos(t(d+s))=\cos(\frac{2s_{1}k\pi}{d_{1}+2s_{1}}), and so f′′(t)=s(d+s)(d+2s)cos(2s1kπd1+2s1)>0f^{\prime\prime}(t)=-s(d+s)(d+2s)\cos(\frac{2s_{1}k\pi}{d_{1}+2s_{1}})>0 if and only if cos(2s1kπd1+2s1)<0\cos(\frac{2s_{1}k\pi}{d_{1}+2s_{1}})<0, i.e., jπ2<2s1kπd1+2s1<(j+2)π2\frac{j\pi}{2}<\frac{2s_{1}k\pi}{d_{1}+2s_{1}}<\frac{(j+2)\pi}{2} for each js1j\leq s_{1} with j1j\equiv 1 (mod 4). Thus, j(d1+2s1)4s1k(j+2)(d1+2s1)4s1\lceil\frac{j(d_{1}+2s_{1})}{4s_{1}}\rceil\leq k\leq\lfloor\frac{(j+2)(d_{1}+2s_{1})}{4s_{1}}\rfloor and f(t)=(d+2s)cos(2s1kπd1+2s1)f(t)=(d+2s)\cos(\frac{2s_{1}k\pi}{d_{1}+2s_{1}}) attains a minima. Using (17), we get Re(UA(Y)(t)v,v)=cos2(s1kπd1+2s1)\operatorname{Re}\left(U_{A(Y)}(t)_{v,v}\right)=\cos^{2}(\frac{s_{1}k\pi}{d_{1}+2s_{1}}), which is positive because d1+2s1d_{1}+2s_{1} is odd by assumption, and so s1kπd1+2s1jπ2\frac{s_{1}k\pi}{d_{1}+2s_{1}}\neq\frac{j\pi}{2} for each any odd jj.

  • Let t=π/dt=\ell\pi/d for odd \ell. Then cos(ts)=cos(s1πd1)=cos(t(d+s))\cos(ts)=\cos(\frac{\ell s_{1}\pi}{d_{1}})=-\cos(t(d+s)), and so f′′(t)=sd(d+s)cos(s1πd1)>0f^{\prime\prime}(t)=sd(d+s)\cos(\frac{\ell s_{1}\pi}{d_{1}})>0 if and only if cos(s1πd1)>0\cos(\frac{\ell s_{1}\pi}{d_{1}})>0. In this case, f(t)=dcos(s1πd1)f(t)=d\cos(\frac{\ell s_{1}\pi}{d_{1}}), and so (17) yields Re(UA(Y)(t)v,v)=12(1+d1cos(s1π/d1)d1+2s1)>12\operatorname{Re}\left(U_{A(Y)}(t)_{v,v}\right)=\frac{1}{2}\big{(}1+\frac{d_{1}\cos(\ell s_{1}\pi/d_{1})}{d_{1}+2s_{1}}\big{)}>\frac{1}{2}.

Combining the above two cases with (14) yields |UA(K2×Y(t)((u,v),(u,v)|C\left|U_{A(K_{2}\times Y}(t)_{((u,v),(u,v)}\right|\geq C as desired. ∎

Example 38.

Let vv be an apex vv of Y=O2ZY=O_{2}\vee Z, where ZZ is a dd-regular graph on n=12s(d+s)n=\frac{1}{2}s(d+s) vertices.

  1. 1.

    Let d=sd=s so that n=s2n=s^{2} and d1=s1=1d_{1}=s_{1}=1. By Theorem 37, C=mink{1,2}cos2(kπ3)=14C=\min_{k\in\{1,2\}}\cos^{2}\left(\frac{k\pi}{3}\right)=\frac{1}{4}. Thus, |UA(K2×Y(τ)((u,v),(u,v)|C\left|U_{A(K_{2}\times Y}(\tau)_{((u,v),(u,v)}\right|\geq C for all tt with equality at time τ=2kπd+2s\tau=\frac{2k\pi}{d+2s}, where k{1,2}k\in\{1,2\}.

  2. 2.

    Let d=2sd=2s so that n=3s22n=\frac{3s^{2}}{2}. In this case, ν2(d)>ν2(s)\nu_{2}(d)>\nu_{2}(s), and so PST occurs between the apexes of YY at some time τ\tau by [Kirkland2023, Theorem 11(1)]. This implies that Re(UA(Y)(τ)v,v)=0\operatorname{Re}\left(U_{A(Y)}(\tau)_{v,v}\right)=0, and so by Theorem 13(2), we get that (u,v)(u,v) is not sedentary in K2YK_{2}\vee Y for any vertex uu of K2K_{2}.

The conditions in Theorem 37 ensure that an apex vv of Y=O2ZY=O_{2}\vee Z is sedentary [Monterde2023, Theorem 36(2)]. Thus, Theorem 37 implies that the bipartite double of a disconnected double cone preserves the sedentariness of its apexes. In contrast, the bipartite double of a complete graph does not preserve the sedentariness of its vertices by Corollary 35(1).

We end this section with a remark about the Laplacian case. Note that the Laplacian matrix of X×YX\times Y is given by L(X×Y)=L(X)D(Y)+D(X)L(Y)L(X)L(Y)L(X\times Y)=L(X)\otimes D(Y)+D(X)\otimes L(Y)-L(X)\otimes L(Y). If XX and YY are both regular, then L(X)D(Y)L(X)\otimes D(Y), D(X)L(Y)D(X)\otimes L(Y) and L(X)L(Y)L(X)\otimes L(Y) pairwise commute, and so UL(X×Y)(t)U_{L(X\times Y)}(t) is simply the product of eit(L(X)D(Y))e^{it(L(X)\otimes D(Y))}, eit(D(X)L(Y))e^{it(D(X)\otimes L(Y))} and eit(L(X)L(Y))e^{it(L(X)\otimes L(Y))}. In this case, X×YX\times Y is also regular and so the quantum walks relative to AA and LL are equivalent. However, if at least one of XX and YY is not regular, then it is not clear how to obtain UL(X×Y)(t)U_{L(X\times Y)}(t). We leave this as an open question.

8 Blow-ups

Throughout this section, we assume that XX is a simple weighted graph. A blow-up of mm copies of XX is the graph with adjacency matrix

JnA(X).\textbf{J}_{n}\otimes A(X).

If (j,u)(j,u) denotes the jjth copy of uV(X)u\in V(X) in 𝑚X\overset{m}{\uplus}\leavevmode\nobreak\ X, then Tu={(j,u):jm}T_{u}=\{(j,u):j\in\mathbb{Z}_{m}\} is a twin set in 𝑚X\overset{m}{\uplus}\leavevmode\nobreak\ X. In particular, if TT is a twin set in XX, then uTTu\bigcup_{u\in T}T_{u} is a twin set in 𝑚X\overset{m}{\uplus}\leavevmode\nobreak\ X. It is known that every vertex of 𝑚X\overset{m}{\uplus}\leavevmode\nobreak\ X is (12m)(1-\frac{2}{m})-sedentary for all m3m\geq 3 [Monterde2023, Theorem 24]. Here, we provide sharper bounds for vertex sedentariness in blow-upss and resolve the case m=2m=2. First, we need to state [bhattacharjya2023quantum, Proposition 2].

Lemma 39.

If uu is a vertex of XX, then

σ(j,u)(A(𝑚X))={mλ:λσu(A(X))}{0}.\sigma_{(j,u)}(A(\overset{m}{\uplus}\leavevmode\nobreak\ X))=\{m\lambda:\lambda\in\sigma_{u}(A(X))\}\cup\{0\}. (18)

Suppose XX has nn vertices. If θ\theta is an eigenvalue of A(X)A(X) with associated orthogonal projection matrix F0F_{0}, then in A(𝑚X)A(\overset{m}{\uplus}\leavevmode\nobreak\ X), [bhattacharjya2023quantum, Equation 2] implies that

E0=Imn[(1/m)Jm(InF0)]E_{0}=I_{mn}-\left[(1/m)\textbf{J}_{m}\otimes(I_{n}-F_{0})\right].

Thus, if 0σu(A(X))0\in\sigma_{u}(A(X)), then (E0)(j,u),(j,u)=11(F0)u,um=m1m+(F0)u,u(E_{0})_{(j,u),(j,u)}=1-\frac{1-(F_{0})_{u,u}}{m}=\frac{m-1}{m}+(F_{0})_{u,u}. Now, if θ\theta is not an eigenvalue of A(X)A(X), then in A(𝑚X)A(\overset{m}{\uplus}\leavevmode\nobreak\ X), [bhattacharjya2023quantum, Equation 5] implies that

E0=Imn[(1/m)JmIn]E_{0}=I_{mn}-\left[(1/m)\textbf{J}_{m}\otimes I_{n}\right].

Thus, if 0σu(A(X))0\notin\sigma_{u}(A(X)), then (E0)(j,u),(j,u)=m1m(E_{0})_{(j,u),(j,u)}=\frac{m-1}{m}. In both cases, taking S={0}S=\{0\} and a=(E0)(j,u),(j,u)a=(E_{0})_{(j,u),(j,u)} in Theorem 15 yields the following result.

Theorem 40.

Let m2m\geq 2 and XX be a simple weighted graph. The following hold relative to A=A(𝑚X)A=A(\overset{m}{\uplus}\leavevmode\nobreak\ X).

  1. 1.

    If 0σu(A(X))0\in\sigma_{u}(A(X)), then |UA(t)(j,u),(j,u)|12m+2(F0)u,u|U_{A}(t)_{(j,u),(j,u)}|\geq 1-\frac{2}{m}+2(F_{0})_{u,u} for all tt and for each jmj\in\mathbb{Z}_{m}, where F0F_{0} is the orthogonal projection matrix associated with the eigenvalue 0 of A(X)A(X).

  2. 2.

    If 0σu(A(X))0\notin\sigma_{u}(A(X)), then |UA(t)(j,u),(j,u)|12m|U_{A}(t)_{(j,u),(j,u)}|\geq 1-\frac{2}{m} for all tt and for each jmj\in\mathbb{Z}_{m}.

In both cases, equality holds if and only if there is a time t1t_{1} such that eit1λ=1e^{it_{1}\lambda}=-1 for all 0λσ(j,u)(A)0\neq\lambda\in\sigma_{(j,u)}(A).

If m=2m=2 and 0σu(A(X))0\notin\sigma_{u}(A(X)), Theorem 40(2) implies that vertex (j,u)(j,u) may or may not be sedentary in 2X\overset{2}{\uplus}\leavevmode\nobreak\ X. We resolve this in the next corollary, which is obtained by a direct application of Lemma 9.

Corollary 41.

If m=2m=2 and 0σu(A(X))0\notin\sigma_{u}(A(X)), then (j,u)(j,u) is adjacency sedentary in 𝑚X\overset{m}{\uplus}\leavevmode\nobreak\ X for each jmj\in\mathbb{Z}_{m} if and only if there are integers mjm_{j} such that λjσu(A(X))mjλj=0\sum_{\lambda_{j}\in\sigma_{u}(A(X))}m_{j}\lambda_{j}=0 and λjσu(A(X))mj\sum_{\lambda_{j}\in\sigma_{u}(A(X))}m_{j} is odd. If we add that ϕ(A(X),t)[t]\phi(A(X),t)\in\mathbb{Z}[t] and uu is periodic, then the latter statement is equivalent to each eigenvalue λjσu(A(X))\lambda_{j}\in\sigma_{u}(A(X)) is of the form λj=bjΔ\lambda_{j}=b_{j}\sqrt{\Delta}, where bjb_{j} is an integer and either Δ=1\Delta=1 or Δ>1\Delta>1 is a square-free integer and the ν2(bj)\nu_{2}(b_{j})’s are not all equal. In this case, uu is tightly sedentary.

We close this section by providing families of graphs whose blow-ups contain sedentary vertices.

Example 42.

Consider the cycle CnC_{n} for all n3n\geq 3. By [bhattacharjya2023quantum, Theorem 8], 𝑚Cn\overset{m}{\uplus}\leavevmode\nobreak\ C_{n} does not admit PGST for all m2m\geq 2. Invoking Corollary 12, we get that each vertex of 𝑚Cn\overset{m}{\uplus}\leavevmode\nobreak\ C_{n} is sedentary for all m2m\geq 2.

Example 43.

Consider the path PnP_{n} for all n3n\geq 3. Making use of [bhattacharjya2023quantum, Theorem 8] and Corollary 12, we get that vertex (j,u)(j,u) in 𝑚Pn\overset{m}{\uplus}\leavevmode\nobreak\ P_{n} is sedentary for all m2m\geq 2 and for each all j𝐙mj\in\mathbf{Z}_{m} whenever

  • n=2tr1,n=2^{t}r-1, t0t\geq 0 and rr is an odd composite number; and

  • n=2tr1,n=2^{t}r-1, t>0t>0, rr is an odd prime number and uu is not a multiple of 2t12^{t-1}.

9 Joins

We now examine sedentariness under the join operation relative to M{A,L}M\in\{A,L\}. Here, we require the underlying graphs to be regular when dealing with M=AM=A. To do this, we first need the following result due to Kirkland and Monterde [kirkland2023quantum, Corollaries 70 and 72].

Lemma 44.

Let M{A,L}M\in\{A,L\}. For all u,vV(X)u,v\in V(X) and for all tt, we have

||UM(XY)(t)u,v||UM(X)(t)u,v||2/|V(X)|.\left|\ |U_{M(X\vee Y)}(t)_{u,v}|-|U_{M(X)}(t)_{u,v}|\ \right|\leq 2/|V(X)|.
Theorem 45.

If uu is CC-sedentary in XX with C>2|V(X)|C>\frac{2}{|V(X)|}, then uu is (C2|V(X)|)\left(C-\frac{2}{|V(X)|}\right)-sedentary in XYX\vee Y for any graph YY, where we require that XX and YY are both regular whenever M=AM=A.

Proof.

By assumption, we have |UX(t)u,u|2/|V(X)|C2/|V(X)|>0|U_{X}(t)_{u,u}|-2/|V(X)|\geq C-2/|V(X)|>0 for all tt. Thus, Lemma 44 yields |UM(t)u,u||UX(t)u,u|2/|V(X)|C2/|V(X)|>0|U_{M}(t)_{u,u}|\geq|U_{X}(t)_{u,u}|-2/|V(X)|\geq C-2/|V(X)|>0 for all tt. ∎

From Theorem 45, if CC is large enough, then CC-sedentariness in XX is preserved in XYX\vee Y.

Example 46.

Let m,nm,n be integers such that mnm+n>2\frac{mn}{m+n}>2. By Corolllary 35(2), each vertex of X=Km×KnX=K_{m}\times K_{n} is CC-sedentary, where C=mn2(m+n1)mnC=\frac{mn-2(m+n-1)}{mn}. By assumption, C>2mn=2|V(X)|C>\frac{2}{mn}=\frac{2}{|V(X)|}, and so each vertex of XX is (mn2(m+n)mn)(\frac{mn-2(m+n)}{mn})-sedentary in XYX\vee Y for any graph YY by Theorem 45. Moreover, this holds for M{A,L}M\in\{A,L\}, where we require that YY is regular whenever M=AM=A.

If C=2/|V(X)|C=2/|V(X)|, then the conclusion in Theorem 45 may not hold for any graph YY.

Example 47.

Let kk be odd and consider X=CP(2k)X=CP(2k) with non-adjacent vertices uu and vv. By Corollary 20, every vertex in XX is tightly CC-sedentary, where C=1kC=\frac{1}{k}. Let YY be a graph on nn vertices that is \ell-regular whenever M=AM=A. Note that uu and vv are twins in XX and in XYX\vee Y.

  1. 1.

    Let M=LM=L. From [kirkland2023quantum, Example 41], PST occurs between uu and vv in XYX\vee Y if and only if n2n\equiv 2 (mod 4). Since each vertex in XYX\vee Y is periodic, PST and PGST are equivalent in XYX\vee Y. Hence, by Corollary 12, each vertex of XX is sedentary in XYX\vee Y if and only if n2n\not\equiv 2 (mod 4).

  2. 2.

    Let M=AM=A. From [kirkland2023quantum, Example 47], PST occurs between uu and vv in XYX\vee Y if and only if n=s(2k2+s)2kn=\frac{s(2k-\ell-2+s)}{2k} and ν2()>ν2(s)=1\nu_{2}(\ell)>\nu_{2}(s)=1. Moreover, PGST occurs between uu and vv in XYX\vee Y if and only if (m+2)2+4mn(m+\ell-2)^{2}+4mn is not a perfect square. Invoking Corollary 12, we get that each vertex of XX is sedentary in XYX\vee Y if and only if n=s(2k2+s)2kn=\frac{s(2k-\ell-2+s)}{2k} and ν2()<ν2(s)\nu_{2}(\ell)<\nu_{2}(s) or ν2(s)1\nu_{2}(s)\neq 1.

We end this section by showing a vertex in XX can be sedentary in XYX\vee Y but not in XX.

Example 48.

Let X=CP(2k)X=CP(2k), where kk is even. By Corollary 20, no vertex in XX is sedentary since each is involved in PST with minimum time π2\frac{\pi}{2}. The same argument in Example 47 yields the following:

  1. 1.

    Let M=LM=L. From [kirkland2023quantum, Corollary 28], PST occurs between uu and vv in XYX\vee Y if and only if n0n\equiv 0 (mod 4). Thus, each vertex of XX is sedentary in XYX\vee Y if and only if n0n\not\equiv 0 (mod 4).

  2. 2.

    Let M=AM=A. Applying [kirkland2023quantum, Lemma 3(2)] yields σu(A(XY)={0,2,12(2k2+±Δ)}\sigma_{u}(A(X\vee Y)=\{0,-2,\frac{1}{2}(2k-2+\ell\pm\sqrt{\Delta})\}, where Δ=(2k2)2+8kn\Delta=(2k-\ell-2)^{2}+8kn. If Δ\Delta is a perfect square, then uu is involved in PST if and only if ν2(2k2+±Δ)=1\nu_{2}(2k-2+\ell\pm\sqrt{\Delta})=1 [Kirkland2023, Theorem 10]. But if Δ\Delta is not a perfect square, then uu is involved in PGST if and only if 2k2+02k-2+\ell\equiv 0 (mod 4) [Kirkland2023, Theorem 13]. Thus, each vertex of XX is sedentary in XYX\vee Y if and only if either (i) Δ\Delta is a perfect square and ν2(λ)1\nu_{2}(\lambda)\neq 1 for some λ{2k2+±Δ}\lambda\in\{2k-2+\ell\pm\sqrt{\Delta}\} or (ii) 2k2+02k-2+\ell\not\equiv 0 (mod 4).

10 Open questions

In this paper, we proved new results on sedentariness. We showed that sedentariness and PGST are dichotomous amongst twin vertices. This allowed us to characterize sedentary vertices in complete multipartite graphs and threshold graphs and provide sharp bounds on their sedentariness. We also investigated sedentariness under direct products, blow-ups and joins. Our results can be used to constructs new examples of graphs with sedentary vertices. We now present some open questions that deserve further exploration.

  1. 1.

    Characterize sedentariness in trees, Cayley graphs and distance-regular graphs.

  2. 2.

    Under what conditions is a vertex in a graph neither sedentary nor involved in PGST? Are there other types of vertices that admit the dichotomous property similar to that of twin vertices?

  3. 3.

    Is the addition of a weighted loop or an attachment of a pendent path to a vertex induce or preserve sedentariness? If yes, then we ask: which weights of loops or lengths of paths achieve this task?

  4. 4.

    Determine other graph operations (such as the rooted product, lexicographic product, strong product and corona product) that induce and/or preserve sedentariness.

  5. 5.

    Characterize sedentary vertices in threshold graphs relative to the adjacency matrix, and provide tight bounds on their sedentariness.

Acknowledgements

I thank the University of Manitoba Faculty of Science and Faculty of Graduate Studies for the support. I also thank Steve Kirkland and Sarah Plosker for the guidance.

References