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

Quantum walks on join graphs

Steve Kirkland,​​1 Hermie Monterde ​​1
Abstract

The join XYX\vee Y of two graphs XX and YY is the graph obtained by joining each vertex of XX to each vertex of YY. We explore the behaviour of a continuous quantum walk on a weighted join graph having the adjacency matrix or Laplacian matrix as its associated Hamiltonian. We characterize strong cospectrality, periodicity and perfect state transfer (PST) in a join graph. We also determine conditions in which strong cospectrality, periodicity and PST are preserved in the join. Under certain conditions, we show that there are graphs with no PST that exhibits PST when joined by another graph. This suggests that the join operation is promising in producing new graphs with PST. Moreover, for a periodic vertex in XX and XYX\vee Y, we give an expression that relates its minimum periods in XX and XYX\vee Y. While the join operation need not preserve periodicity and PST, we show that ||UM(XY,t)u,v||UM(X,t)u,v||2|V(X)|\big{|}|U_{M}(X\vee Y,t)_{u,v}|-|U_{M}(X,t)_{u,v}|\big{|}\leq\frac{2}{|V(X)|} for all vertices uu and vv of XX, where UM(XY,t)U_{M}(X\vee Y,t) and UM(X,t)U_{M}(X,t) denote the transition matrices of XYX\vee Y and XX respectively relative to either the adjacency or Laplacian matrix. We demonstrate that the bound 2|V(X)|\frac{2}{|V(X)|} is tight for infinite families of graphs.

Keywords: quantum walk, join graph, perfect state transfer, strong cospectrality, adjacency matrix, Laplacian matrix

MSC2010 Classification: 05C50; 05C76; 05C22; 15A16; 15A18; 81P45;

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

1 Introduction

The graphs K2K_{2}, C4C_{4} P3P_{3}, and K4\eK_{4}\backslash e are well-known examples of small graphs that admit adjacency or Laplacian perfect state transfer. Kirkland et al. noticed that in these small graphs, the vertices involved in perfect state transfer share the same neighbours, an observation that prompted them to examine state transfer between twins [Kirkland2023]. However, it can also be observed that these small graphs are in fact join graphs. This motivates our investigation on quantum walks on graphs built using the join operation.

The first infinite family of join graphs revealed to admit Laplacian perfect state transfer was Kn\e:=O2Kn2K_{n}\backslash e:=O_{2}\vee K_{n-2} (the complete graph on nn vertices minus an edge) with n0n\equiv 0 (mod 4) [Bose2008]. Motivated by this result, Angeles-Canul et al. gave sufficient conditions for adjacency perfect state transfer to occur between the vertices of X{O2,K2}X\in\{O_{2},K_{2}\} in the unweighted graph XYX\vee Y, where YY is a regular graph [Angeles-Canul2010]. Such a graph is called a double cone on YY, and the vertices of X{O2,K2}X\in\{O_{2},K_{2}\} are called the apexes of the double cone. They also determined sufficient conditions such that adjacency perfect state transfer in XX is preserved under joins of copies of XX. In a subsequent paper, Angeles-Canul et al. investigated perfect state transfer in weighted join graphs, and found that the apexes of a double cone on a kk-regular graph admit adjacency perfect state transfer by appropriate choice of weights of an edge between the apexes and/or loops on the apexes [Angeles-Canul2009]. More recently, Kirkland et al. fully characterized unweighted double cones that admit adjacency perfect state transfer between apexes [Kirkland2023]. For the Laplacian case, Alvir et al. showed that the apexes of unweighted O2YO_{2}\vee Y admit perfect state transfer if and only if |V(Y)|2|V(Y)|\equiv 2 (mod 4), while the apexes of unweighted K2YK_{2}\vee Y do not admit perfect state transfer [Alvir2016]. Joins have also been investigated in graphs with well-structured eigenbases [Johnston2017, mclaren2023weak], and in the contexts of fractional revival [Monterde2023a, Chan2021], sedentariness [Monterde2023] and strong cospectrality [Monterde2022].

Despite its widespread presence in literature, state transfer on join graphs remains largely unexplored. In this paper, we provide a systematic study of quantum walks on weighted join graphs having the adjacency and Laplacian matrices as their associated Hamiltonian. In Section 3, we provide known results about transition matrices and eigenvalue supports of vertices in join graphs. In Section 4, we characterize periodic vertices in a join (Theorem 4), and determine the conditions in which periodicity of a vertex in XX and XYX\vee Y are equivalent (Corollaries 7 and 9). We also use the join operation to provide infinite families of graphs that are not integral (resp., Laplacian integral) whereby each member graph contains periodic vertices (Corollary 12). We devote Section 5 to finding a relationship between the minimum periods ρX\rho_{X} and ρXY\rho_{X\vee Y} of a vertex that is periodic in both XX and XYX\vee Y, resp. (Theorems 14 and 15). We find that ρXY\rho_{X\vee Y} is a rational multiple of ρX\rho_{X}, and if we add the extra hypothesis that XX is disconnected, then ρXY\rho_{X\vee Y} is an integer multiple of ρX\rho_{X}. Section 6 provides a characterization of strong cospectrality in joins (Theorem 20). It turns out that strong cospectrality between two vertices of XX in XYX\vee Y requires strong cospectrality in XX, except for the case when XX is an empty graph on two vertices. In fact, for the unweighted case, strong cospectrality in XX and XYX\vee Y are equivalent with a few exceptions (Corollary 20). In Section 7, we characterize perfect state transfer in joins (Theorems 23 and 26). Unlike strong cospectrality, perfect state transfer between two vertices of XX in XYX\vee Y does not require perfect state transfer between them in XX. This motivates us in Section 8 to determine necessary and sufficient conditions such that perfect state transfer in XX is preserved in XYX\vee Y (Corollaries 28 and 35), and in Section 9 to determine conditions such that perfect state transfer occurs in XYX\vee Y given that it does not occur in XX (Theorems 40 and 42, and Corollary 46). In Section 10, we characterize strong cospectrality and PST in self-joins and iterated join graphs. The latter result generalizes a previously known characterization of perfect state transfer in threshold graphs. While the join operation does not preserve periodicity and PST in general, we show in Section 11 that ||UM(XY,t)u,v||UM(X,t)u,v||2|V(X)|\big{|}|U_{M}(X\vee Y,t)_{u,v}|-|U_{M}(X,t)_{u,v}|\big{|}\leq\frac{2}{|V(X)|} for all tt and for all vertices uu and vv of XX (Corollaries 70 and 72). Consequently, the behaviour of vertices in XX in the quantum walk on XYX\vee Y mimics the behaviour of the quantum walk on XX as |V(X)||V(X)|\rightarrow\infty (Corollary 73). We also show that this upper bound is tight for infinite families of graphs (Examples 75 and 76). Finally, open problems are presented in Section 12.

2 Preliminaries

Throughout, we assume that a graph XX is weighted and undirected with possible loops but no multiple edges. We denote the vertex set of XX by V(X)V(X), and we allow the edges of XX to have positive real weights. We say that XX is simple if XX has no loops, and XX is unweighted if all edges of XX have weight one. For uV(X)u\in V(X), we denote the the characteristic vector of uu as eu\textbf{e}_{u}, which is a vector with a 11 on the entry indexed by uu and zeros elsewhere. 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}, and if the context is clear, then we simply write these matrices as 1, 0, J and II, respectively. We also represent the transpose of the matrix HH by HTH^{T}, and the characteristic polynomial of HH in the variable tt by ϕ(H,t)\phi(H,t). Lastly, we denote the simple unweighted empty, cycle, complete, and path graphs on nn vertices as OnO_{n}, CnC_{n}, KnK_{n}, and PnP_{n}, respectively. We also denote the hypercube on 2p2^{p} vertices by QpQ_{p}, and the cocktail party graph on mm vertices by CP(m)CP(m), where mm is even.

The adjacency matrix A(X)A(X) of XX is the matrix 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). The Laplacian matrix of XX is the matrix L(X)=D(X)A(X)L(X)=D(X)-A(X). We use M(X)M(X) to denote A(X)A(X) or L(X)L(X). If the context is clear, then we write M(X)M(X), A(X)A(X), L(X)L(X) and D(X)D(X) resp. as MM, AA, LL and DD. We say that XX is integral (resp., Laplacian integral) if all eigenvalues of A(X)A(X) (resp., L(X)L(X)) are integers. 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), i.e., the row sums of A(X)A(X) are a constant kk. If XX is simple, then XX is weighted kk-regular if and only if D(X)=kID(X)=kI.

A (continuous-time) quantum walk on XX with respect to HH is determined by the unitary matrix

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

The matrix HH is called the Hamiltonian of the quantum walk. Typically, HH is taken to be AA or LL, but any Hermitian matrix HH that respects the adjacencies of XX works in general. Since M{A,L}M\in\{A,L\} is real symmetric, MM 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}. If the eigenvalues are not indexed, then we also denote by EλE_{\lambda} the orthogonal projection matrix corresponding to the eigenvalue λ\lambda of MM. Now, (2) allows us to write (1) as

UM(X,t)=jeitλjEj.U_{M}(X,t)=\sum_{j}e^{it\lambda_{j}}E_{j}.

We say that perfect state transfer (PST) occurs between two vertices uu and vv if |UM(X,τ)u,v|=1|U_{M}(X,\tau)_{u,v}|=1 for some time τ\tau. The minimum τ>0\tau>0 such that |UM(X,τ)u,v|=1|U_{M}(X,\tau)_{u,v}|=1 is called the minimum PST time between uu and vv. We say that uu is periodic if |UM(X,τ)u,u|=1|U_{M}(X,\tau)_{u,u}|=1 for some time τ\tau. The minimum τ>0\tau>0 such that |UM(X,τ)u,u|=1|U_{M}(X,\tau)_{u,u}|=1 is called the minimum period of uu. These properties depend on the matrix MM, so we sometimes say adjacency PST (resp., periodicity) when M=AM=A; similar language applies when M=LM=L. Further, if XX is simple, weighted and kk-regular, then UL(X,t)=eitL=eit(kIA)=eitkeitA=eitkUA(X,t)U_{L}(X,t)=e^{itL}=e^{it(kI-A)}=e^{itk}e^{-itA}=e^{itk}U_{A}(X,-t). Hence, |UL(X,τ)u,v|2=|UA(X,τ)u,v|2\left|U_{L}(X,\tau)_{u,v}\right|^{2}=\left|U_{A}(X,\tau)_{u,v}\right|^{2} for any u,vV(X)u,v\in V(X), so the quantum walks on AA and LL exhibit the same state transfer properties.

3 Joins

The join XYX\vee Y of weighted graphs XX and YY is the graph obtained by joining every vertex of XX with every vertex of YY with an edge of weight one. Unless otherwise stated, we make the following assumptions:

  • when dealing with M=L,M=L, we take XX and YY to be simple;

  • when dealing with M=A,M=A, we take XX and YY to be weighted kk- and \ell-regular resp. (possibly with loops).

Denote the transition matrices of XYX\vee Y and XX by UM(XY,t)U_{M}(X\vee Y,t) and UM(X,t)U_{M}(X,t) respectively. The transition matrix of a join with respect to LL and AA are known (see [Alvir2016, Fact 8] and [Coutinho2021, Lemma 4.3.1], respectively). Throughout, mm and nn denote the number of vertices of XX and YY, respectively.

Lemma 1.

If u,vV(X)u,v\in V(X) and wV(Y)w\in V(Y), then

UL(XY,t)u,w=1m+n(1eit(m+n))U_{L}(X\vee Y,t)_{u,w}=\frac{1}{m+n}\left(1-e^{it(m+n)}\right) (3)

and

UL(XY,t)u,v=eitnUL(X,t)u,v+1m+n+neit(m+n)m(m+n)eitnm.U_{L}(X\vee Y,t)_{u,v}=e^{itn}U_{L}(X,t)_{u,v}+\frac{1}{m+n}+\frac{ne^{it(m+n)}}{m(m+n)}-\frac{e^{itn}}{m}. (4)
Lemma 2.

Let D=(k)2+4mnD=(k-\ell)^{2}+4mn and λ±=12(k+±D)\lambda^{\pm}=\frac{1}{2}\left(k+\ell\pm\sqrt{D}\right). If u,vV(X)u,v\in V(X) and wV(Y)w\in V(Y), then

UA(XY,t)u,w=1D(eitλ+eitλ)U_{A}(X\vee Y,t)_{u,w}=\frac{1}{\sqrt{D}}(e^{it\lambda^{+}}-e^{it\lambda^{-}}) (5)

and

UA(XY,t)u,v=UA(X,t)u,v+eitλ+(kλ)mDeitλ(kλ+)mDeitkm.U_{A}(X\vee Y,t)_{u,v}=U_{A}(X,t)_{u,v}+\frac{e^{it\lambda^{+}}(k-\lambda^{-})}{m\sqrt{D}}-\frac{e^{it\lambda^{-}}(k-\lambda^{+})}{m\sqrt{D}}-\frac{e^{itk}}{m}. (6)

The eigenvalue support of uu with respect to MM is the set

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

The following result determines the elements of the eigenvalue support of a vertex in a join graph. In what follows, σu(M(X))\sigma_{u}(M(X)) and σu(M)\sigma_{u}(M) denote the eigenvalue supports of vertex uu in XX and XYX\vee Y, respectively.

Lemma 3.

Let uV(X)u\in V(X), and define 𝒮:={λ+n:λσu(L(X))\{0}}\mathcal{S}:=\{\lambda+n:\lambda\in\sigma_{u}(L(X))\backslash\{0\}\}.

  1. 1.

    We have σu(L)=𝒮\sigma_{u}(L)=\mathcal{S}\cup\mathcal{R}, where ={0,m+n}\mathcal{R}=\{0,m+n\} if XX is connected and ={0,m+n,n}\mathcal{R}=\{0,m+n,n\} otherwise.

  2. 2.

    We have σu(A)=σu(A(X))\{k}\sigma_{u}(A)=\sigma_{u}(A(X))\backslash\{k\}\cup\mathcal{R}, where ={λ±}\mathcal{R}=\{\lambda^{\pm}\} if XX is connected and ={λ±,k}\mathcal{R}=\{\lambda^{\pm},k\} otherwise.

4 The ratio condition

A set SS\subseteq\mathbb{R} with at least two elements satisfies the ratio condition if λθμη\frac{\lambda-\theta}{\mu-\eta}\in\mathbb{Q} for all λ,θ,μ,ηS\lambda,\theta,\mu,\eta\in S with μη\mu\neq\eta. A vertex uu is periodic if and only if σu(M)\sigma_{u}(M) satisfies the ratio condition [Coutinho2021, Corollary 8.3.1]. Combining this fact with Lemma 3 yields a characterization of periodicity in joins.

Theorem 4.

Let XX and YY be weighted graphs and uV(X)u\in V(X).

  1. 1.

    Vertex uu is Laplacian periodic in XYX\vee Y if and only if all eigenvalues in σu(L(X))\sigma_{u}(L(X)) are rational. Moreover, if uu is Laplacian periodic in XYX\vee Y, then it is Laplacian periodic in XX.

  2. 2.

    If XX is connected (resp., disconnected), then uu is adjacency periodic if and only if λ+λD\frac{\lambda^{+}-\lambda}{\sqrt{D}}\in\mathbb{Q} for all λσu(A(X))\{k}\lambda\in\sigma_{u}(A(X))\backslash\{k\} (resp., λσu(A(X))\lambda\in\sigma_{u}(A(X))). If we assume further that k,,Dk,\ell,\sqrt{D}\in\mathbb{Q}, then uu is adjacency periodic in XYX\vee Y if and only if all eigenvalues in σu(A(X))\sigma_{u}(A(X)) are rational.

The following example illustrates that the converse of the last statement in Theorem 4(1) does not hold.

Example 5.

Consider XO1X\vee O_{1}, where XX is a weighted K2K_{2} with positive edge weight η\eta\notin\mathbb{Q}. Then uV(X)u\in V(X) is periodic in XX and σu(L)={0,3,1+2η}\sigma_{u}(L)=\{0,3,1+2\eta\}. Since 1+2η030\frac{1+2\eta-0}{3-0}\notin\mathbb{Q}, we get that uu is not periodic in XO1X\vee O_{1}.

If ϕ(M(X),t)[t]\phi(M(X),t)\in\mathbb{Z}[t], then we obtain another characterization of periodicity [Coutinho2021, Corollary 7.6.2].

Lemma 6.

Let SS be a set of real algebraic integers that is closed under taking algebraic conjugates. If SS satisfies the ratio condition, then either (i) SS\subseteq\mathbb{Z} or (ii) each λj\lambda_{j} in SS is a quadratic integer of the form λj=12(a+bjΔ)\lambda_{j}=\frac{1}{2}(a+b_{j}\sqrt{\Delta}), where a,bj,Δa,b_{j},\Delta\in\mathbb{Z} and Δ>1\Delta>1 is square-free. In particular, if XX is a weighted graph with possible loops such that ϕ(M(X),t)[t]\phi(M(X),t)\in\mathbb{Z}[t], then vertex uu is periodic in XX if and only if either (i) or (ii) holds for S=σu(M(X))S=\sigma_{u}(M(X)). In this case, ρ=2π/gΔ\rho=2\pi/g\sqrt{\Delta}, where each λjσu(M)\lambda_{j}\in\sigma_{u}(M), g=gcd{(λ1λj)/Δ}g=\operatorname{gcd}\{(\lambda_{1}-\lambda_{j})/\sqrt{\Delta}\} and Δ=1\Delta=1 if (i) holds.

We now show that the converse of the last statement in Theorem 4(1) holds when ϕ(L(X),t)[t]\phi(L(X),t)\in\mathbb{Z}[t].

Corollary 7.

If ϕ(L(X),t)[t]\phi(L(X),t)\in\mathbb{Z}[t], then the following statements are equivalent.

  1. 1.

    Vertex uu is Laplacian periodic in XX.

  2. 2.

    Vertex uu is Laplacian periodic in XYX\vee Y for any simple positively weighted graph YY.

  3. 3.

    The eigenvalues in σu(L(X))\sigma_{u}(L(X)) are all integers.

Proof.

Lemma 3(1) and the ratio condition yields (3) implies (2). The second statement of Theorem 4(1) yields (2) implies (1). Lemma 6(i) and the fact that 0σu(L(X))0\in\sigma_{u}(L(X)) yield (1) implies (3). ∎

Next, we have the following result concerning the adjacency matrix in relation to Theorem 4(2).

Corollary 8.

Let uV(X)u\in V(X), ϕ(A(X),t)[t]\phi(A(X),t)\in\mathbb{Z}[t] and k,k,\ell\in\mathbb{Z}. Consider DD in Lemma 2.

  1. 1.

    If XX is connected, then uu is adjacency periodic in XYX\vee Y if and only if either:

    1. (a)

      The eigenvalues in σu(A(X))\sigma_{u}(A(X)) are all integers and DD is a perfect square (so that λ±\lambda^{\pm}\in\mathbb{Z}). In this case, uu is also adjacency periodic in XX.

    2. (b)

      Each λjσu(A(X))\{k}\lambda_{j}\in\sigma_{u}(A(X))\backslash\{k\} is a quadratic integer of the form 12(k++bjΔ)\frac{1}{2}(k+\ell+b_{j}\sqrt{\Delta}), where Δ>1\Delta>1 is square-free, D/Δ\sqrt{D/\Delta}\in\mathbb{Z} and D/Δ>bj\sqrt{D/\Delta}>b_{j} for each jj. In this case, uu is not periodic in XX.

  2. 2.

    If XX is disconnected, then uu is adjacency periodic in XYX\vee Y if and only if condition (1a) holds.

Proof.

To prove (1), suppose XX is connected. By Lemma 3(2), σu(A)=σu(A(X))\{k}{λ±}\sigma_{u}(A)=\sigma_{u}(A(X))\backslash\{k\}\cup\{\lambda^{\pm}\}, and uu is periodic in XYX\vee Y if and only if σu(A)\sigma_{u}(A) satisfies the ratio condition. Since kk\in\mathbb{Z}, the latter condition implies that σu(A(X))\{k}\sigma_{u}(A(X))\backslash\{k\} satisfies the ratio condition, and as ϕ(A(X),t)[t]\phi(A(X),t)\in\mathbb{Z}[t], we get two cases from Lemma 6. First, let σu(A(X))\{k}\sigma_{u}(A(X))\backslash\{k\}\subseteq\mathbb{Z}. Since λ+=12(k++D)\lambda^{+}=\frac{1}{2}(k+\ell+\sqrt{D}), we get λ+λD\frac{\lambda^{+}-\lambda}{\sqrt{D}}\in\mathbb{Q} if and only if D\sqrt{D}\in\mathbb{Z}. Combining this with Theorem 4(2) yields (1a). Next, suppose each λjσu(A(X))\{k}\lambda_{j}\in\sigma_{u}(A(X))\backslash\{k\} is of the form 12(a+bjΔ)\frac{1}{2}(a+b_{j}\sqrt{\Delta}) so that λ+λj=12((k+a)+(DbjΔ))\lambda^{+}-\lambda_{j}=\frac{1}{2}\left((k+\ell-a)+(\sqrt{D}-b_{j}\sqrt{\Delta})\right). Then λ+λjD\frac{\lambda^{+}-\lambda_{j}}{\sqrt{D}}\in\mathbb{Q} if and only if k+a=0k+\ell-a=0 and Δ/D\sqrt{\Delta/D}\in\mathbb{Q}. Now, Perron-Frobenius Theorem yields λ+=12(a+D)\lambda^{+}=\frac{1}{2}(a+\sqrt{D}) as the largest eigenvalue of A(XY)A(X\vee Y). Thus, D>Δ\sqrt{D}>\sqrt{\Delta}. Since σu(A)\sigma_{u}(A) satisfies the ratio condition, we get D=bΔ\sqrt{D}=b\sqrt{\Delta} for some integer bb, and so D/Δ=b>bj\sqrt{D/\Delta}=b>b_{j} for each jj. This establishes (1b).

To prove (2), suppose XX is disconnected. In this case, kσu(A(X))k\in\sigma_{u}(A(X)) is an integer, and so we apply Lemma 4(2) to conclude that uu is periodic in XYX\vee Y if and only if σu(A(X))\sigma_{u}(A(X))\subseteq\mathbb{Z} and D\sqrt{D}\in\mathbb{Q}. ∎

We now state an analogue of Corollary 7, which is an immediate consequence of Corollary 8.

Corollary 9.

Let ϕ(A(XY),t)[t]\phi(A(X\vee Y),t)\in\mathbb{Z}[t] and k,k,\ell\in\mathbb{Z}. If uu is adjacency periodic in XX, then uu is adjacency periodic in XYX\vee Y if and only if DD is a perfect square. Moreover, if DD is a perfect square, then the following are equivalent.

  1. 1.

    Vertex uu is adjacency periodic in XX.

  2. 2.

    Vertex uu is adjacency periodic in XYX\vee Y for any weighted \ell-regular graph YY.

  3. 3.

    All eigenvalues in σu(A(X))\sigma_{u}(A(X)) are integers.

Remark 10.

For LL, our assumption on XX and YY implies that the join operation preserves periodicity, and periodicity in the join is inherited by the underlying graph. However, this does not hold for AA given our assumption on XX and YY. One can have σu(A(X))\sigma_{u}(A(X))\subseteq\mathbb{Z} but D\sqrt{D}\notin\mathbb{Z}, and so uu is periodic in XX but not in XYX\vee Y. Conversely, periodicity in XYX\vee Y is not necessarily inherited by XX by Corollary 4.

We say that XX is periodic if there is a τ>0\tau>0 such that U(τ)=γIU(\tau)=\gamma I for some unit γ\gamma\in\mathbb{C}. Equivalently, each uV(X)u\in V(X) is periodic at time τ\tau. We characterize periodic join graphs under mild conditions.

Theorem 11.

Let XX and YY be weighted graphs and ϕ(M(XY),t)[t]\phi(M(X\vee Y),t)\in\mathbb{Z}[t].

  1. 1.

    If XX and YY are simple, then XYX\vee Y is Laplacian periodic if and only if XYX\vee Y is Laplacian integral.

  2. 2.

    If XX and YY are regular, then XYX\vee Y is adjacency periodic if and only if XYX\vee Y is integral.

Proof.

Note that (1) follows from the ratio condition and the fact that L(XY)L(X\vee Y) is positive semidefinite with 0 as an eigenvalue. Next, let XX and YY be kk- and \ell-regular. By way of contradiction, suppose each eigenvalue λj\lambda_{j} of A(XY)A(X\vee Y) is a quadratic integer. Since λ±=12(k+±D)\lambda^{\pm}=\frac{1}{2}(k+\ell\pm\sqrt{D}) is an eigenvalue of A(XY)A(X\vee Y), each λj\lambda_{j} must be of the form 12(k++bjΔ)\frac{1}{2}(k+\ell+b_{j}\sqrt{\Delta}). Without loss of generality, let k>0k\geq\ell>0. For each eigenvalue λj\lambda_{j}\neq\ell of A(Y)A(Y), we get λj=12(k++bjΔ)+12bjΔ\lambda_{j}=\frac{1}{2}(k+\ell+b_{j}\sqrt{\Delta})\geq\ell+\frac{1}{2}b_{j}\sqrt{\Delta} for all each jj. Similarly, λj¯12bjΔ\overline{\lambda_{j}}\geq\ell-\frac{1}{2}b_{j}\sqrt{\Delta} for all each jj. Thus, λj>\lambda_{j}>\ell or λj¯>\overline{\lambda_{j}}>\ell, which cannot happen because \ell is the largest eigenvalue of A(Y)A(Y) by the Perron-Frobenius Theorem. Invoking Lemma 6 establishes (2). ∎

We close this section by providing infinite families of graphs that are not (Laplacian) integral but contain periodic vertices. The following result is immediate from Lemma 3 and the ratio condition.

Corollary 12.

Let XX and YY be weighted graphs, and let uV(X)u\in V(X).

  1. 1.

    If XX is Laplacian integral but YY is not, then XYX\vee Y is not Laplacian integral, the eigenvalues in σu(L(X))\sigma_{u}(L(X)) are all integers, and uu is periodic in XYX\vee Y.

  2. 2.

    If XX is integral but YY is not, then XYX\vee Y is not integral. Moreover, if k,k,\ell\in\mathbb{Z} and DD in Lemma 2 is a perfect square, then the eigenvalues in σu(A(X))\sigma_{u}(A(X)) are all integers and uu is periodic in XYX\vee Y.

5 Minimum periods

Denote the minimum periods of uu in XYX\vee Y and XX by ρXY\rho_{X\vee Y} and ρX\rho_{X} respectively. Motivated by Corollaries 7 and 9, we ask: if uV(X)u\in V(X) is periodic in XX and XYX\vee Y, then how are ρXY\rho_{X\vee Y} and ρX\rho_{X} related? To answer this question, we state a result due to Kirkland et al. [Kirkland2023, Theorem 4].

Theorem 13.

Let uu and vv be vertices in XX, and suppose σu(M)={λ1,,λr}\sigma_{u}(M)=\{\lambda_{1},\ldots,\lambda_{r}\}. If uu is periodic, then ρ=2πqλ1λ2\rho=\frac{2\pi q}{\lambda_{1}-\lambda_{2}}, where q=lcm(q2,,qr)q=\operatorname{lcm}(q_{2},\ldots,q_{r}) and each qjq_{j} is a positive integer such that λ1λjλ1λ2=pjqj\frac{\lambda_{1}-\lambda_{j}}{\lambda_{1}-\lambda_{2}}=\frac{p_{j}}{q_{j}} for some integer pjp_{j} satisfying gcd(pj,qj)=1\text{gcd}(p_{j},q_{j})=1. In particular, if r=2r=2, then uu is periodic with ρ=2πλ1λ2\rho=\frac{2\pi}{\lambda_{1}-\lambda_{2}}.

We now address our above question by finding a constant c>0c>0 such that ρXY=cρX\rho_{X\vee Y}=c\rho_{X}.

Theorem 14.

Let uu be a non-isolated and periodic vertex in XX and XYX\vee Y with σu(L(X))={λ1,,λr,λr+1}\sigma_{u}(L(X))=\{\lambda_{1},\ldots,\lambda_{r},\lambda_{r+1}\}, where λr+1=0\lambda_{r+1}=0. Let λr+2:=m\lambda_{r+2}:=m and λr+3:=n\lambda_{r+3}:=-n.

  1. 1.

    Suppose XX is connected.

    1. (a)

      If r=1r=1, then ρXY=(λ1qλ1+n)ρX\rho_{X\vee Y}=\left(\frac{\lambda_{1}q}{\lambda_{1}+n}\right)\rho_{X}, where qq is a positive integer such that m+nλ1+n=pq\frac{m+n}{\lambda_{1}+n}=\frac{p}{q} for some integer pp with gcd(p,q)=1\operatorname{gcd}(p,q)=1. In particular, if λ1=m\lambda_{1}=m, then q=1q=1.

    2. (b)

      If r=2r=2 and λ2=m\lambda_{2}=m, then ρXY=(λ1qq(λ1+n))ρX\rho_{X\vee Y}=\left(\frac{\lambda_{1}q}{q^{\prime}(\lambda_{1}+n)}\right)\rho_{X}, where qq is a positive integer defined in (a) and qq^{\prime} is a positive integer such that m/λ1=p/qm/\lambda_{1}=p^{\prime}/q^{\prime} for some integer pp^{\prime} with gcd(p,q)=1\operatorname{gcd}(p^{\prime},q^{\prime})=1.

    3. (c)

      If r2r\geq 2 and λjm\lambda_{j}\neq m for each j{1,,r}j\in\{1,\ldots,r\}, then

      ρXY=(qr+2qr+3gcd(R1,qr+1)qr+1gcd(R1,qr+2)gcd(R2,qr+3))ρX,\rho_{X\vee Y}=\left(\frac{q_{r+2}q_{r+3}\operatorname{gcd}(R_{1},q_{r+1})}{q_{r+1}\operatorname{gcd}(R_{1},q_{r+2})\operatorname{gcd}(R_{2},q_{r+3})}\right)\rho_{X}, (7)

      where R1=lcm(q3,,qr)R_{1}=\operatorname{lcm}(q_{3},\ldots,q_{r}), R2=lcm(q2,,qr,qr+2)R_{2}=\operatorname{lcm}(q_{2},\ldots,q_{r},q_{r+2}), the qjq_{j}’s are positive integers such that λ1λjλ1λ2=pjqj\frac{\lambda_{1}-\lambda_{j}}{\lambda_{1}-\lambda_{2}}=\frac{p_{j}}{q_{j}} for some integers pjp_{j} with gcd(pj,qj)=1\operatorname{gcd}(p_{j},q_{j})=1 and λ1>λ2\lambda_{1}>\lambda_{2}.

    4. (d)

      If r3r\geq 3 and λr=m\lambda_{r}=m, then (7) holds with R1=lcm(q3,,qr1)R_{1}=\operatorname{lcm}(q_{3},\ldots,q_{r-1}), R2=lcm(q3,,qr1,qr+2)R_{2}=\operatorname{lcm}(q_{3},\ldots,q_{r-1},q_{r+2}).

  2. 2.

    Suppose XX is disconnected.

    1. (a)

      If r=1r=1, then ρXY=qρX\rho_{X\vee Y}=q\rho_{X}, where q=lcm(q3,q4)q=\operatorname{lcm}(q_{3},q_{4}) and the qjq_{j}’s are positive integers such that λ1λjλ1λ2=pjqj\frac{\lambda_{1}-\lambda_{j}}{\lambda_{1}-\lambda_{2}}=\frac{p_{j}}{q_{j}} for some pjp_{j} with gcd(pj,qj)=1\operatorname{gcd}(p_{j},q_{j})=1. In particular, if λ1=m\lambda_{1}=m, then q=mgcd(m,n)q=\frac{m}{\operatorname{gcd}(m,n)}.

    2. (b)

      If r=2r=2 and λ2=m\lambda_{2}=m, then ρXY=(q/q4)ρX\rho_{X\vee Y}=(q/q_{4})\rho_{X}, where q=lcm(q4,q5)q=\operatorname{lcm}(q_{4},q_{5}) and the qjq_{j}’s are positive integers such that λ1λjλ1=pjqj\frac{\lambda_{1}-\lambda_{j}}{\lambda_{1}}=\frac{p_{j}}{q_{j}} for some integers pjp_{j} with gcd(pj,qj)=1\operatorname{gcd}(p_{j},q_{j})=1.

    3. (c)

      If r2r\geq 2 and λjm\lambda_{j}\neq m for each j{1,,r}j\in\{1,\ldots,r\}, then

      ρXY=(qr+2qr+3gcd(q,qr+2)gcd(R,qr+3))ρX,\rho_{X\vee Y}=\left(\frac{q_{r+2}q_{r+3}}{\operatorname{gcd}(q,q_{r+2})\operatorname{gcd}(R,q_{r+3})}\right)\rho_{X}, (8)

      where R=lcm(q3,,qr+2)R=\operatorname{lcm}(q_{3},\ldots,q_{r+2}), the qjq_{j}’s are positive integers such that λ1λjλ1λ2=pjqj\frac{\lambda_{1}-\lambda_{j}}{\lambda_{1}-\lambda_{2}}=\frac{p_{j}}{q_{j}} for some integers pjp_{j} with gcd(pj,qj)=1\operatorname{gcd}(p_{j},q_{j})=1 and λ1>λ2\lambda_{1}>\lambda_{2}. Further, the coefficient of ρX\rho_{X} in (8) is an integer.

    4. (d)

      If r3r\geq 3 and λr=m\lambda_{r}=m, then

      ρXY=(Q/q)ρX,\rho_{X\vee Y}=(Q/q)\rho_{X}, (9)

      where q=lcm(q3,,qr,qr+1)q=\operatorname{lcm}(q_{3},\ldots,q_{r},q_{r+1}), Q=lcm(q3,,qr1,qr+1,qr+2,qr+3)Q=\operatorname{lcm}(q_{3},\ldots,q_{r-1},q_{r+1},q_{r+2},q_{r+3}) and the qjq_{j}’s are positive integers satisfying λ1λjλ1=pjqj\frac{\lambda_{1}-\lambda_{j}}{\lambda_{1}}=\frac{p_{j}}{q_{j}} and gcd(pj,qj)=1\operatorname{gcd}(p_{j},q_{j})=1. In this case, Q/qQ/q is an integer.

Proof.

Let XX be connected. By Lemma 3(1), σu(L)={λ1+n,,λr+n}{λr+2+n,λr+3+n}\sigma_{u}(L)=\{\lambda_{1}+n,\ldots,\lambda_{r}+n\}\cup\{\lambda_{r+2}+n,\lambda_{r+3}+n\}. Note that λjλr+3\lambda_{j}\neq\lambda_{r+3} for each j{1,,r}j\in\{1,\ldots,r\}. However, it is possible that λj=λr+2=m\lambda_{j}=\lambda_{r+2}=m for some j{1,,r}j\in\{1,\ldots,r\}. In this case, we may instead assume that λr=m\lambda_{r}=m. We have the following cases.

  • Let r=1r=1. Then σu(L(X))={λ1,0}\sigma_{u}(L(X))=\{\lambda_{1},0\} and σu(L)={λ1+n,m+n,0}\sigma_{u}(L)=\{\lambda_{1}+n,m+n,0\}, and so Theorem 13 yields ρX=2π/λ1\rho_{X}=2\pi/\lambda_{1} and ρXY=2πq/(λ1+n)\rho_{X\vee Y}=2\pi q/(\lambda_{1}+n), where q=1q=1 whenever λ1=m\lambda_{1}=m.

  • Let r=2r=2 and λr=m\lambda_{r}=m. Then σu(L)={λ1+n,m+n,0}\sigma_{u}(L)=\{\lambda_{1}+n,m+n,0\}, and so ρXY=2πq/(λ1+n)\rho_{X\vee Y}=2\pi q/(\lambda_{1}+n) as in the previous case. Since σu(L(X))={λ1,m,0}\sigma_{u}(L(X))=\{\lambda_{1},m,0\}, Theorem 13 yields ρX=2πq/λ1\rho_{X}=2\pi q^{\prime}/\lambda_{1}.

  • Let r2r\geq 2 and λjm\lambda_{j}\neq m for each jj. As σu(L)={λ1+n,,λr+n,λr+2+n,λr+3+n}\sigma_{u}(L)=\{\lambda_{1}+n,\ldots,\lambda_{r}+n,\lambda_{r+2}+n,\lambda_{r+3}+n\} and uu is periodic in XX and XYX\vee Y, Theorem 13 yields ρX=2πq/(λ1λ2)\rho_{X}=2\pi q/(\lambda_{1}-\lambda_{2}) and ρXY=2πQ/(λ1λ2)\rho_{X\vee Y}=2\pi Q/(\lambda_{1}-\lambda_{2}), where q=lcm(q3,,qr+1)q=\operatorname{lcm}(q_{3},\ldots,q_{r+1}), Q=lcm(q3,,qr,qr+2,qr+3)Q=\operatorname{lcm}(q_{3},\ldots,q_{r},q_{r+2},q_{r+3}) and λ1>λ2\lambda_{1}>\lambda_{2}. Since lcm(a,b,c)=lcm(lcm(a,b),c)\operatorname{lcm}(a,b,c)=\operatorname{lcm}(\operatorname{lcm}(a,b),c) and lcm(a,b)=abgcd(a,b)\operatorname{lcm}(a,b)=\frac{ab}{\operatorname{gcd}(a,b)}, we may write q=R1qr+1gcd(R1,qr+1)q=\frac{R_{1}q_{r+1}}{\operatorname{gcd}(R_{1},q_{r+1})} and

    Q=R1qr+2qr+3gcd(R1,qr+2)gcd(R2,qr+3)=qc,Q=\frac{R_{1}q_{r+2}q_{r+3}}{\operatorname{gcd}(R_{1},q_{r+2})\operatorname{gcd}(R_{2},q_{r+3})}=qc, (10)

    where cc is the coefficient of ρX\rho_{X} in (7). This implies that ρXY=cρX\rho_{X\vee Y}=c\rho_{X}.

  • Let r3r\geq 3 and λr=m\lambda_{r}=m. Then σu(L)={λ1+n,,λr1+n,λr+2,λr+3}\sigma_{u}(L)=\{\lambda_{1}+n,\ldots,\lambda_{r-1}+n,\lambda_{r+2},\lambda_{r+3}\}, and so the same argument in previous subcase yields (10) with R1=lcm(q3,,qr1)R_{1}=\operatorname{lcm}(q_{3},\ldots,q_{r-1}) and R2=lcm(q3,,qr1,qr+2)R_{2}=\operatorname{lcm}(q_{3},\ldots,q_{r-1},q_{r+2}).

Combining the above cases proves (1).

Next, let XX be disconnected. By Lemma 3(1), σu(L)={λ1+n,,λr+n}{λr+1+n,λr+2+n,λr+3+n}\sigma_{u}(L)=\{\lambda_{1}+n,\ldots,\lambda_{r}+n\}\cup\{\lambda_{r+1}+n,\lambda_{r+2}+n,\lambda_{r+3}+n\}. Moreover, the fact that uu is a non-isolated vertex yields r1r\geq 1. We have the following cases.

  • Let r=1r=1. Then σu(L(X))={λ1,0}\sigma_{u}(L(X))=\{\lambda_{1},0\} and σu(L)={λ1+n,λ2+n,λ3+n,λ4+n}\sigma_{u}(L)=\{\lambda_{1}+n,\lambda_{2}+n,\lambda_{3}+n,\lambda_{4}+n\}. By Theorem 13, ρX=2π/λ1\rho_{X}=2\pi/\lambda_{1} and ρXY=2πq/λ1\rho_{X\vee Y}=2\pi q/\lambda_{1}. In particular, if λ1=m\lambda_{1}=m, then q1=q3q_{1}=q_{3}, and so q=q4q=q_{4}.

  • Let r=2r=2 and λr=m\lambda_{r}=m. Then σu(L)={λ1+n,λ3+n,λ4+n,λ5+n}\sigma_{u}(L)=\{\lambda_{1}+n,\lambda_{3}+n,\lambda_{4}+n,\lambda_{5}+n\} and σu(L(X))={λ1,m,0}\sigma_{u}(L(X))=\{\lambda_{1},m,0\}. By Theorem 13, we get ρXY=2πqλ1λ3=2πqλ1\rho_{X\vee Y}=\frac{2\pi q}{\lambda_{1}-\lambda_{3}}=\frac{2\pi q}{\lambda_{1}} and ρX=2πqλ1\rho_{X}=\frac{2\pi q^{\prime}}{\lambda_{1}}, where p,qp^{\prime},q^{\prime}\in\mathbb{Z} such that mλ1=pq\frac{m}{\lambda_{1}}=\frac{p^{\prime}}{q^{\prime}}\in\mathbb{Q} and gcd(p,q)=1\operatorname{gcd}(p^{\prime},q^{\prime})=1. Since p4q4=λ1λ4λ1λ3=λ1mλ1=1pq\frac{p_{4}}{q_{4}}=\frac{\lambda_{1}-\lambda_{4}}{\lambda_{1}-\lambda_{3}}=\frac{\lambda_{1}-m}{\lambda_{1}}=1-\frac{p^{\prime}}{q^{\prime}}, we get q4=qq_{4}=q^{\prime}.

  • Let r2r\geq 2 and λjm\lambda_{j}\neq m for each jj. Then σu(L)={λ1+n,,λr+n,λr+1+n,λr+2+n,λr+3+n}\sigma_{u}(L)=\{\lambda_{1}+n,\ldots,\lambda_{r}+n,\lambda_{r+1}+n,\lambda_{r+2}+n,\lambda_{r+3}+n\}. By Theorem 13, ρXY=2πQ/(λ1λ2)\rho_{X\vee Y}=2\pi Q/(\lambda_{1}-\lambda_{2}) where Q=lcm(q3,,qr+3)Q=\operatorname{lcm}(q_{3},\ldots,q_{r+3}) and qjq_{j}’s are as defined in (1c). The same argument then yields Q=qqr+2qr+3gcd(q,qr+2)gcd(R,qr+3)=qc,Q=\frac{qq_{r+2}q_{r+3}}{\operatorname{gcd}(q,q_{r+2})\operatorname{gcd}(R,q_{r+3})}=qc, where q=lcm(q3,,qr+1)q=\operatorname{lcm}(q_{3},\ldots,q_{r+1}).

  • Let r3r\geq 3 and λr=m\lambda_{r}=m. Since σu(L)={λ1+n,,λr1+n,λr+1+n,λr+2+n,λr+3+n}\sigma_{u}(L)=\{\lambda_{1}+n,\ldots,\lambda_{r-1}+n,\lambda_{r+1}+n,\lambda_{r+2}+n,\lambda_{r+3}+n\}, we get ρX=2πq/λ1\rho_{X}=2\pi q/\lambda_{1} and ρXY=2πQ/λ1\rho_{X\vee Y}=2\pi Q/\lambda_{1} by Theorem 13. Now, since λ1λrλ1λr+1=λ1mλ1=prqr\frac{\lambda_{1}-\lambda_{r}}{\lambda_{1}-\lambda_{r+1}}=\frac{\lambda_{1}-m}{\lambda_{1}}=\frac{p_{r}}{q_{r}} and λ1λrλ1λr+2=λ1mλ1=pr+2qr+2\frac{\lambda_{1}-\lambda_{r}}{\lambda_{1}-\lambda_{r+2}}=\frac{\lambda_{1}-m}{\lambda_{1}}=\frac{p_{r+2}}{q_{r+2}}, we get qr=qr+2q_{r}=q_{r+2}, and so Q=lcm(q2,,qr+1,qr+3)Q=\operatorname{lcm}(q_{2},\ldots,q_{r+1},q_{r+3}). Thus, Q/qQ/q\in\mathbb{Z}.

Combining the above cases proves (2). ∎

Using the same argument, we get an analogue of Theorem 14 for the adjacency case.

Theorem 15.

Let uu be a non-isolated and periodic vertex in XX and XYX\vee Y with σu(A(X))={λ1,,λr,λr+1}\sigma_{u}(A(X))=\{\lambda_{1},\ldots,\lambda_{r},\lambda_{r+1}\}, where λr+1=k\lambda_{r+1}=k. Let λr+2:=λ\lambda_{r+2}:=\lambda^{-} and λr+3:=λ+\lambda_{r+3}:=\lambda^{+}, where λ±\lambda^{\pm} are defined in Lemma 2.

  1. 1.

    Suppose XX is connected.

    1. (a)

      Let r=1r=1. Then ρXY=(q(kλ1)D)ρX\rho_{X\vee Y}=\left(\frac{q(k-\lambda_{1})}{\sqrt{D}}\right)\rho_{X}, where qq is a positive integer such that λ+λ1D=pq\frac{\lambda^{+}-\lambda_{1}}{\sqrt{D}}=\frac{p}{q} for some integer pp with gcd(p,q)=1\operatorname{gcd}(p,q)=1. In particular, if λ1=λ\lambda_{1}=\lambda^{-}, then q=1q=1.

    2. (b)

      If r=2r=2 and λ2=λ\lambda_{2}=\lambda^{-}, then ρXY=((λ1λ)qqD)ρX\rho_{X\vee Y}=\left(\frac{(\lambda_{1}-\lambda^{-})q}{q^{\prime}\sqrt{D}}\right)\rho_{X}, where qq is an integer in (1b) and qq^{\prime} is a positive integer such that λ1kλ1λ=pq\frac{\lambda_{1}-k}{\lambda_{1}-\lambda^{-}}=\frac{p^{\prime}}{q^{\prime}}\in\mathbb{Q} for some integer pp^{\prime} with gcd(p,q)=1\operatorname{gcd}(p^{\prime},q^{\prime})=1.

    3. (c)

      If r2r\geq 2 and λjλ\lambda_{j}\neq\lambda^{-} for each j{1,,r}j\in\{1,\ldots,r\}, then the conclusion in Theorem 14(1d) holds.

    4. (d)

      If r3r\geq 3 and λr=λ\lambda_{r}=\lambda^{-}, then the conclusion in Theorem 14(1e) holds.

  2. 2.

    Suppose XX is disconnected.

    1. (a)

      Let r=1r=1. Then ρXY=qρX\rho_{X\vee Y}=q\rho_{X}, where q=lcm(q3,q4)q=\operatorname{lcm}(q_{3},q_{4}) and the qjq_{j}’s are positive integers such that kλjkλ1=pjqj\frac{k-\lambda_{j}}{k-\lambda_{1}}=\frac{p_{j}}{q_{j}} for some pjp_{j} with gcd(pj,qj)=1\operatorname{gcd}(p_{j},q_{j})=1. In particular, if λ1=λ\lambda_{1}=\lambda^{-}, then q=q4q=q_{4}.

    2. (b)

      If r=2r=2 and λ2=λ\lambda_{2}=\lambda^{-}, then ρXY=(q/q2)ρX\rho_{X\vee Y}=\left(q/q_{2}\right)\rho_{X}, where q=lcm(q2,q5)q=\operatorname{lcm}(q_{2},q_{5}), the qjq_{j}’s are positive integers such that kλjkλ1=pjqj\frac{k-\lambda_{j}}{k-\lambda_{1}}=\frac{p_{j}}{q_{j}} for some integers pjp_{j} with gcd(pj,qj)=1\operatorname{gcd}(p_{j},q_{j})=1 and λ1>λ2\lambda_{1}>\lambda_{2}.

    3. (c)

      If r2r\geq 2 and λjλ\lambda_{j}\neq\lambda^{-} for each j{1,,r}j\in\{1,\ldots,r\}, then the conclusion in Theorem 14(2d) holds.

    4. (d)

      If r3r\geq 3 and λr=λ\lambda_{r}=\lambda^{-}, then the conclusion in Theorem 14(2e) holds.

Remark 16.

The advantage of Theorems 14-15 is that they apply even if ϕ(M(X),t)[t]\phi(M(X),t)\notin\mathbb{Z}[t]. We also mention that the constant c>0c>0 such that ρXY=cρX\rho_{X\vee Y}=c\rho_{X} in Theorems 14-15 is always rational. In fact, cc is an integer whenever (i) XX is disconnected or (ii) XX is connected and qr+1=1q_{r+1}=1 in Theorems 14-15(1c-d).

For the family of complete graphs, Theorem 14(1a) yields ρXY=mm+nρX\rho_{X\vee Y}=\frac{m}{m+n}\rho_{X}, where X=KmX=K_{m} and Y=KnY=K_{n}. Here, c=mm+n<1c=\frac{m}{m+n}<1 is rational, and this holds for M{A,L}M\in\{A,L\}. We complement this observation by providing an infinite family of connected graphs XX such that ρXY=cρX\rho_{X\vee Y}=c\rho_{X} for some integer c1c\geq 1.

Example 17.

Suppose that mm is even and let X=Km2,m2X=K_{\frac{m}{2},\frac{m}{2}}. Since σu(L(X))={m/2,m,0}\sigma_{u}(L(X))=\{m/2,m,0\}, each uV(X)u\in V(X) is periodic with ρX=4π/m\rho_{X}=4\pi/m. Consider XYX\vee Y, where YY is a simple unweighted graph on nn vertices. Applying Theorem 14(1b), we get ρXY=cρX\rho_{X\vee Y}=c\rho_{X}, where c=m/2gc=m/2g and g=gcd(m/2,n)g=\operatorname{gcd}(m/2,n). Since gg divides m/2m/2, we conclude that cc is an integer.

Example 18.

Let X=KmX=K_{m} and YY be a simple unweighted \ell-regular graph on nn vertices. Each uV(X)u\in V(X) is periodic with ρX=2π/m\rho_{X}=2\pi/m. By Lemma 3(2), σu(A)={λ±,1}\sigma_{u}(A)=\{\lambda^{\pm},-1\}, where λ±=12(+m1±D)\lambda^{\pm}=\frac{1}{2}(\ell+m-1\pm\sqrt{D}). By Corollary 9, uu periodic in XYX\vee Y if and only if D=(m+1)2+4mnD=(\ell-m+1)^{2}+4mn is a perfect square. Equivalently, n=s(m+s+1)mn=\frac{s(\ell-m+s+1)}{m} for some integer ss such that mm divides s(+s+1)s(\ell+s+1). Thus, if uu periodic in XYX\vee Y, then λ+(1)λ+λ=+s+1m+2s+1\frac{\lambda^{+}-(-1)}{\lambda^{+}-\lambda^{-}}=\frac{\ell+s+1}{\ell-m+2s+1}\in\mathbb{Q}, from which we get ρXY=2π/g\rho_{X\vee Y}=2\pi/g, where g=gcd(+s+1,sm)g=\operatorname{gcd}(\ell+s+1,s-m). Hence ρXY=(m/g)ρX\rho_{X\vee Y}=(m/g)\rho_{X}. Taking s=2ms=2m and 0n20\leq\ell\leq n-2 yields g=gcd(+1,m)g=\operatorname{gcd}(\ell+1,m), and so m/gm/g is an integer.

6 Strong cospectrality

We say that two vertices uu and vv in a weighted graph XX are strongly cospectral if Ejeu=±EjevE_{j}\textbf{e}_{u}=\pm E_{j}\textbf{e}_{v} for each jj, in which case we define the sets

σuv+(M)={λj:Ejeu=Ejev0}andσuv(M)={λj:Ejeu=Ejev0}.\sigma_{uv}^{+}(M)=\{\lambda_{j}:E_{j}\textbf{e}_{u}=E_{j}\textbf{e}_{v}\neq\textbf{0}\}\quad\text{and}\quad\sigma_{uv}^{-}(M)=\{\lambda_{j}:E_{j}\textbf{e}_{u}=-E_{j}\textbf{e}_{v}\neq\textbf{0}\}. (11)

Equivalently, vertices uu and vv are strongly cospectral if for each jj, either every eigenvector w associated with λj\lambda_{j} satisfies wTeu=wTev\textbf{w}^{T}\textbf{e}_{u}=\textbf{w}^{T}\textbf{e}_{v} or every eigenvector w associated with λj\lambda_{j} satisfies wTeu=wTev\textbf{w}^{T}\textbf{e}_{u}=-\textbf{w}^{T}\textbf{e}_{v}.

If vertices uu and vv are strongly cospectral, then σu(M)=σuv+(M)σuv(M)\sigma_{u}(M)=\sigma_{uv}^{+}(M)\cup\sigma_{uv}^{-}(M), and uu and vv belong to the same connected component of XX (so neither of them is isolated). In order to avoid confusion, if uu and vv are strongly cospectral in XX and XYX\vee Y, then we write the sets in (11) as σuv±(M(X))\sigma_{uv}^{\pm}(M(X)) and σuv±(M)\sigma_{uv}^{\pm}(M), respectively.

Strong cospectrality is a necessary condition for PST. Hence, in order to investigate PST in joins, we first need to characterize strong cospectrality. In what follows, we let On(k)O_{n}(k) denote the empty graph on nn vertices with a loop of weight kk on each vertex. If k=0k=0, then we simply write On(k)O_{n}(k) as OnO_{n}.

Theorem 19.

Let m2m\geq 2 and u,vV(X)u,v\in V(X) with uvu\neq v.

  1. 1.

    Vertices uu and vv in XX are Laplacian strongly cospectral in XYX\vee Y if and only if either:

    1. (a)

      Vertices uu and vv are Laplacian strongly cospectral in XX and mσuv(L(X))m\notin\sigma_{uv}^{-}(L(X)). In this case,

      σuv+(L)={λ+n:0λσuv+(L(X))}andσuv(L)={μ+n:μσuv(L(X))}\sigma_{uv}^{+}(L)=\{\lambda+n:0\neq\lambda\in\sigma_{uv}^{+}(L(X))\}\cup\mathcal{R}\quad\text{and}\quad\sigma_{uv}^{-}(L)=\{\mu+n:\mu\in\sigma_{uv}^{-}(L(X))\}

      where ={0,m+n}\mathcal{R}=\{0,m+n\} if XX is connected and ={0,m+n,n}\mathcal{R}=\{0,m+n,n\} otherwise.

    2. (b)

      X=O2X=O_{2}. In this case, σuv+(L)={0,n+2}\sigma_{uv}^{+}(L)=\{0,n+2\} and σuv(L)={n}\sigma_{uv}^{-}(L)=\{n\}.

  2. 2.

    Vertices uu and vv in XX are adjacency strongly cospectral in XYX\vee Y if and only if either:

    1. (a)

      Vertices uu and vv are adjacency strongly cospectral in XX and λ±σuv(L(X))\lambda^{\pm}\notin\sigma_{uv}^{-}(L(X)). In this case,

      σuv+(A)=σuv+(A(X))\{k}andσuv(A)=σuv(A(X))\sigma_{uv}^{+}(A)=\sigma_{uv}^{+}(A(X))\backslash\{k\}\cup\mathcal{R}\quad\text{and}\quad\sigma_{uv}^{-}(A)=\sigma_{uv}^{-}(A(X))

      where ={λ±}\mathcal{R}=\{\lambda^{\pm}\} if XX is connected and ={λ±,k}\mathcal{R}=\{\lambda^{\pm},k\} otherwise.

    2. (b)

      X=O2(k)X=O_{2}(k). In this case, σuv+(L)={λ±}\sigma_{uv}^{+}(L)=\{\lambda^{\pm}\} and σuv(L)={k}\sigma_{uv}^{-}(L)=\{k\}.

  3. 3.

    If wV(Y)w\in V(Y), then vertices uu and ww are not strongly cospectral in XYX\vee Y.

Moreover, if (1a) or (2a) holds, then vertices uu and vv belong to the same connected component in XX.

Proof.

Let λ\lambda and μ\mu be nonzero eigenvalues of L(X)L(X) and L(Y)L(Y) with associated eigenvectors yλ\textbf{y}_{\lambda} and zμ\textbf{z}_{\mu}, respectively. Then 0, m+nm+n, λ+n\lambda+n and μ+n\mu+n are eigenvalues of LL with associated eigenvectors

1m+n,u=[n1mm1n],vλ=[yλ0],andwμ=[0zμ]\textbf{1}_{m+n},\quad\textbf{u}=\left[\begin{array}[]{ccccc}n\textbf{1}_{m}\\ -m\textbf{1}_{n}\end{array}\right],\quad\textbf{v}_{\lambda}=\left[\begin{array}[]{cc}\textbf{y}_{\lambda}\\ \textbf{0}\end{array}\right],\quad\text{and}\quad\textbf{w}_{\mu}=\left[\begin{array}[]{cc}\textbf{0}\\ \textbf{z}_{\mu}\end{array}\right] (12)

respectively. Now, let u,vV(X)u,v\in V(X). We have two cases.

  • Let uu and vv be non-isolated in XX. From the form of vλ\textbf{v}_{\lambda}’s in (12), strong cospectrality in XX is required for strong cospectrality in XYX\vee Y. Hence, we assume uu and vv are strongly cospectral in XX. If mσuv(L(X))m\in\sigma_{uv}^{-}(L(X)), then vm\textbf{v}_{m} is another eigenvector for m+nm+n. In this case, uTeu=uTev\textbf{u}^{T}\textbf{e}_{u}=\textbf{u}^{T}\textbf{e}_{v} and vmTeu=vmTev\textbf{v}_{m}^{T}\textbf{e}_{u}=-\textbf{v}_{m}^{T}\textbf{e}_{v}, and so uu and vv are not strongly cospectral in XYX\vee Y. However, if mσuv(L(X))m\notin\sigma_{uv}^{-}(L(X)), then uu and vv are strongly cospectral in XYX\vee Y with σuv+(L)\sigma_{uv}^{+}(L) and σuv(L)\sigma_{uv}^{-}(L) in Theorem 19(1a) as desired.

  • Let uu and vv be isolated in XX. Then v=euev\textbf{v}=\textbf{e}_{u}-\textbf{e}_{v} is an eigenvector for LL associated to the eigenvalue nn. If X=O2X=O_{2}, then [Monterde2022, Corollary 6.9(2)] yields the desired conclusion. However, if XX has a connected component CC other than {u}\{u\} and {v}\{v\}, then the vector w which is constant on each connected component of XX and whose sum of all entries is 0 is also an eigenvector for LL associated to the eigenvalue nn. This vector satisfies wTeu=wTev\textbf{w}^{T}\textbf{e}_{u}=\textbf{w}^{T}\textbf{e}_{v}, and since vTeu=vTev\textbf{v}^{T}\textbf{e}_{u}=-\textbf{v}^{T}\textbf{e}_{v}, we get that uu and vv are not strongly cospectral in XYX\vee Y.

Combining these cases proves (1). Next, let λk\lambda\neq k and μ\mu\neq\ell be eigenvalues of A(X)A(X) and L(Y)L(Y) respectively, with associated eigenvectors yλ\textbf{y}_{\lambda} and zμ\textbf{z}_{\mu}. Then λ±\lambda^{\pm}, λ\lambda and μ\mu are eigenvalues of AA with associated eigenvectors

u=[(kλ)1mm1n],vλ=[yλ0],andwμ=[0zμ]\textbf{u}=\left[\begin{array}[]{ccccc}(k-\lambda^{\mp})\textbf{1}_{m}\\ m\textbf{1}_{n}\end{array}\right],\quad\textbf{v}_{\lambda}=\left[\begin{array}[]{cc}\textbf{y}_{\lambda}\\ \textbf{0}\end{array}\right],\quad\text{and}\quad\textbf{w}_{\mu}=\left[\begin{array}[]{cc}\textbf{0}\\ \textbf{z}_{\mu}\end{array}\right] (13)

respectively. The same argument as the previous case yields the desired conclusion for (2). Finally, the form of the vλ\textbf{v}_{\lambda}’s in (12-13) yields (3) and the last statement. ∎

From Theorem 19(3), we assume henceforth that strongly cospectral vertices in XYX\vee Y belong to XX.

Note that if XX is unweighted. then mm is an eigenvalue of L(X)L(X) if and only if XX is a join. From (12), we get that mσuv(L(Z))m\in\sigma_{uv}^{-}(L(Z)) if and only if Z=K2Z=K_{2}. Combining this with Theorem 19 yields the next result.

Corollary 20.

Let XX be an unweighted graph with vertices uu and vv.

  1. 1.

    Let X{K2,O2}X\notin\{K_{2},O_{2}\}. Then uu and vv in XX are Laplacian strongly cospectral in XYX\vee Y if and only if they are in XX.

  2. 2.

    Let XO2X\neq O_{2} and suppose XYX\vee Y is not a complete graph. Then uu and vv in XX are adjacency strongly cospectral in XYX\vee Y if and only if they are in XX.

Proof.

Let XX be unweighted. Then mm is an eigenvalue of L(Z)L(Z) if and only if ZZ is a join. In particular, mσuv(L(Z))m\in\sigma_{uv}^{-}(L(Z)) if and only if Z=K2Z=K_{2} by (12). Combining this with Theorem 19 yields (1). To prove (2), note that λ=12(k+(k)2+4mn)=1\lambda^{-}=\frac{1}{2}(k+\ell-\sqrt{(k-\ell)^{2}+4mn})=-1 if and only if m=k+1m=k+1 and n=+1n=\ell+1, i.e., XX and YY are complete. ∎

Example 21.

Consider XYX\vee Y, where X{Km,Om}X\in\{K_{m},O_{m}\}. If m3m\geq 3, then the vertices of XX do not admit strongly cospectrality in XX and XYX\vee Y with respect to M{A,L}M\in\{A,L\} [Monterde2022, Corollary 3.10]. If m=2m=2, then the following hold about the apexes uu and vv of the double cone XYX\vee Y.

  1. 1.

    By Theorem 19(1b-2b), uu and vv are Laplacian and adjacency strongly cospectral in O2YO_{2}\vee Y.

  2. 2.

    Since 2σuv(L(X))2\in\sigma_{uv}^{-}(L(X)), Theorem 19(1a) implies that uu and vv are not Laplacian strongly cospectral in K2YK_{2}\vee Y for any graph YY. Since uu and vv are adjacency strongly cospectral in K2K_{2}, Corollary 20(2) implies that uu and vv are strongly cospectral in K2YK_{2}\vee Y if and only if YY is not a complete graph.

7 Perfect state transfer

To determine PST in joins, we make use of a characterization of PST due to Coutinho. Throughout, we denote the largest power of two that divides an integer aa by ν2(a)\nu_{2}(a). We also denote the minimum PST times between uu and vv in XX and XYX\vee Y by τX\tau_{X} and τXY\tau_{X\vee Y}, respectively.

Theorem 22.

Let MM be a real symmetric matrix such that ϕ(M,t)[t]\phi(M,t)\in\mathbb{Z}[t]. Then vertices uu and vv in XX admit perfect state transfer if and only if all of the following conditions hold.

  1. 1.

    Either (i) σu(M)\sigma_{u}(M)\subseteq\mathbb{Z} or (ii) each λjσu(M)\lambda_{j}\in\sigma_{u}(M) is a quadratic integer of the form 12(a+bjΔ)\frac{1}{2}(a+b_{j}\sqrt{\Delta}) for some integers aa, bjb_{j} and Δ\Delta, where Δ>1\Delta>1 is square-free. Here, we let Δ=1\Delta=1 whenever (i) holds.

  2. 2.

    Vertices uu and vv are strongly cospectral in XX.

  3. 3.

    For all λ,ησuv+(M)\lambda,\eta\in\sigma_{uv}^{+}(M) and μ,θσuv(M)\mu,\theta\in\sigma_{uv}^{-}(M), we have ν2(ληΔ)>ν2(λμΔ)=ν2(λθΔ)\nu_{2}\left(\frac{\lambda-\eta}{\sqrt{\Delta}}\right)>\nu_{2}\left(\frac{\lambda-\mu}{\sqrt{\Delta}}\right)=\nu_{2}\left(\frac{\lambda-\theta}{\sqrt{\Delta}}\right).

Further, τX=πgΔ\tau_{X}=\frac{\pi}{g\sqrt{\Delta}}, where g=gcd({λ0λΔ:λσu(M)})g=\operatorname{gcd}\left(\left\{\frac{\lambda_{0}-\lambda}{\sqrt{\Delta}}:\lambda\in\sigma_{u}(M)\right\}\right) for some fixed λ0σuv+(M)\lambda_{0}\in\sigma_{uv}^{+}(M).

To characterize PST in joins, we now combine Theorem 22 with our results on periodicity and strong cospectrality in joins from Sections 4 and 6. Note that it suffices to consider the vertices of XX in checking whether PST occurs in XYX\vee Y by virtue of Theorem 19(3). We begin with the Laplacian case.

7.1 Laplacian case

Theorem 23.

Let m2m\geq 2 and ϕ(L(X),t)[t]\phi(L(X),t)\in\mathbb{Z}[t]. Vertices uu and vv in XX admit Laplacian perfect state transfer in XYX\vee Y if and only if all of the following conditions hold.

  1. 1.

    Either (i) uu and vv are Laplacian strongly cospectral in XX and mσuv(L(X))m\notin\sigma_{uv}^{-}(L(X)) or (ii) X=O2X=O_{2}.

  2. 2.

    The eigenvalues in σu(L(X))\sigma_{u}(L(X)) are all integers.

  3. 3.

    One of the following conditions hold.

    1. (a)

      XX is connected and one of the following conditions hold for all 0λσuv+(L(X)){m}0\neq\lambda\in\sigma_{uv}^{+}(L(X))\cup\{m\} and μ,θσuv(L(X))\mu,\theta\in\sigma_{uv}^{-}(L(X)).

      1. i.

        ν2(λ)>ν2(μ)=ν2(θ)\nu_{2}(\lambda)>\nu_{2}(\mu)=\nu_{2}(\theta) and ν2(n)>ν2(μ)\nu_{2}(n)>\nu_{2}(\mu).

      2. ii.

        ν2(μ)>ν2(λ)=ν2(n)\nu_{2}(\mu)>\nu_{2}(\lambda)=\nu_{2}(n).

      3. iii.

        ν2(λ)=ν2(μ)=ν2(n)\nu_{2}(\lambda)=\nu_{2}(\mu)=\nu_{2}(n) and ν2(λ+n2ν2(n))>ν2(μ+n2ν2(n))=ν2(θ+n2ν2(n))\nu_{2}\left(\frac{\lambda+n}{2^{\nu_{2}(n)}}\right)>\nu_{2}\left(\frac{\mu+n}{2^{\nu_{2}(n)}}\right)=\nu_{2}\left(\frac{\theta+n}{2^{\nu_{2}(n)}}\right).

    2. (b)

      XO2X\neq O_{2} is disconnected and (ai) above holds. Here, uu and vv are in the same connected component in XX.

    3. (c)

      X=O2X=O_{2} and n2n\equiv 2 (mod 4).

Further, τXY=πg\tau_{X\vee Y}=\frac{\pi}{g}, where g=gcd(𝒯)g=\operatorname{gcd}\left(\mathcal{T}\right), 𝒯=𝒮\mathcal{T}=\mathcal{S}\cup\mathcal{R} and the sets 𝒮,\mathcal{S},\mathcal{R} are given in Lemma 3.

Proof.

By Theorem 19(1) and Corollary 7, conditions (1) and (2) are equivalent resp. to strong cospectrality and periodicity of uu and vv in XYX\vee Y. Combining this with Theorem 22, we get PST between uu and vv in XYX\vee Y if and only if Theorem 22(3) holds. To establish (3a), suppose XX is connected. From Theorem 19(1a), we have σuv+(L)={λ+n:0λσuv+(L(X))}{0,m+n}\sigma_{uv}^{+}(L)=\{\lambda+n:0\neq\lambda\in\sigma_{uv}^{+}(L(X))\}\cup\{0,m+n\} and σuv(L)=σuv(L(X))\sigma_{uv}^{-}(L)=\sigma_{uv}^{-}(L(X)). Theorem 22(3) applied to uu and vv is equivalent to

ν2(λ+n)>ν2(μ+n)=ν2(θ+n)\nu_{2}(\lambda+n)>\nu_{2}(\mu+n)=\nu_{2}(\theta+n) (14)

for all 0λσuv+(L(X)){m}0\neq\lambda\in\sigma_{uv}^{+}(L(X))\cup\{m\} and μ,θσuv(L(X))\mu,\theta\in\sigma_{uv}^{-}(L(X)). We have three cases.

  • Let ν2(λ)>ν2(μ)\nu_{2}(\lambda)>\nu_{2}(\mu) for some 0λσuv+(L(X)){m}0\neq\lambda\in\sigma_{uv}^{+}(L(X))\cup\{m\} and μσuv(L(X))\mu\in\sigma_{uv}^{-}(L(X)). Since ν2(λ+n)>ν2(μ+n)\nu_{2}(\lambda+n)>\nu_{2}(\mu+n) in (14), it must be that ν2(n)>ν2(μ)\nu_{2}(n)>\nu_{2}(\mu) for each μσuv(L(X))\mu\in\sigma_{uv}^{-}(L(X)), in which case, ν2(μ+n)=ν2(μ)\nu_{2}(\mu+n)=\nu_{2}(\mu). Now, for the equality in (14) to hold, we need ν2(μ)=ν2(θ)\nu_{2}(\mu)=\nu_{2}(\theta) for all μ,θσuv(L(X))\mu,\theta\in\sigma_{uv}^{-}(L(X)). Hence, ν2(λ)>ν2(μ)\nu_{2}(\lambda)>\nu_{2}(\mu) for all μσuv(L(X))\mu\in\sigma_{uv}^{-}(L(X)). Finally, if ν2(μ)ν2(η)\nu_{2}(\mu)\geq\nu_{2}(\eta) for some 0ησuv+(L(X)){m}0\neq\eta\in\sigma_{uv}^{+}(L(X))\cup\{m\}, then ν2(n)>ν2(η)\nu_{2}(n)>\nu_{2}(\eta), and so ν2(η+n)=ν2(η)ν2(μ)=ν2(μ+n)\nu_{2}(\eta+n)=\nu_{2}(\eta)\leq\nu_{2}(\mu)=\nu_{2}(\mu+n), a contradiction to (14). Thus, our assumption in this case combined with (14) gives us ν2(λ)>ν2(μ)=ν2(θ)\nu_{2}(\lambda)>\nu_{2}(\mu)=\nu_{2}(\theta) and ν2(n)>ν2(μ)\nu_{2}(n)>\nu_{2}(\mu) for all 0λσuv+(L(X)){m}0\neq\lambda\in\sigma_{uv}^{+}(L(X))\cup\{m\} and μ,θσuv(L(X))\mu,\theta\in\sigma_{uv}^{-}(L(X)), which proves (3ai).

  • Let ν2(μ)>ν2(λ)\nu_{2}(\mu)>\nu_{2}(\lambda) for some 0λσuv+(L(X)){m}0\neq\lambda\in\sigma_{uv}^{+}(L(X))\cup\{m\} and μσuv(L(X))\mu\in\sigma_{uv}^{-}(L(X)). If ν2(λ)ν2(n)\nu_{2}(\lambda)\neq\nu_{2}(n), then ν2(μ+n)ν2(λ+n)\nu_{2}(\mu+n)\geq\nu_{2}(\lambda+n), a contradiction to (14). Hence, we must have ν2(λ)=ν2(n)\nu_{2}(\lambda)=\nu_{2}(n) for each 0λσuv+(L(X)){m}0\neq\lambda\in\sigma_{uv}^{+}(L(X))\cup\{m\}. Finally, if ν2(n)ν2(θ)\nu_{2}(n)\geq\nu_{2}(\theta) for some θσuv(L(X))\theta\in\sigma_{uv}^{-}(L(X)), then we have ν2(θ+n)ν2(n)=ν2(μ+n)\nu_{2}(\theta+n)\neq\nu_{2}(n)=\nu_{2}(\mu+n), which again contradicts (14). Thus, ν2(μ)>ν2(n)=ν2(λ)\nu_{2}(\mu)>\nu_{2}(n)=\nu_{2}(\lambda) for all for some 0λσuv+(L(X)){m}0\neq\lambda\in\sigma_{uv}^{+}(L(X))\cup\{m\} and θσuv(L(X))\theta\in\sigma_{uv}^{-}(L(X)). This proves (3aii).

  • Let ν2(μ)=ν2(λ)\nu_{2}(\mu)=\nu_{2}(\lambda) for some 0λσuv+(L(X)){m}0\neq\lambda\in\sigma_{uv}^{+}(L(X))\cup\{m\} and μσuv(L(X))\mu\in\sigma_{uv}^{-}(L(X)). For the strict inequality in (14) to hold, it is required that ν2(μ)=ν2(λ)=ν2(n)\nu_{2}(\mu)=\nu_{2}(\lambda)=\nu_{2}(n) for all 0λσuv+(L(X)){m}0\neq\lambda\in\sigma_{uv}^{+}(L(X))\cup\{m\} and μσuv(L(X))\mu\in\sigma_{uv}^{-}(L(X)). From this, the conditions in (3aiii) are straightforward.

Combining these cases proves (3a). Now, if XX is disconnected, then nσuv+(L(X))n\in\sigma_{uv}^{+}(L(X)) by Theorem 19(1b), and so Theorem 22(3) holds if and only if (14) holds and ν2(n)>ν2(μ+n)\nu_{2}(n)>\nu_{2}(\mu+n) for all μσuv(L(X))\mu\in\sigma_{uv}^{-}(L(X)). Equivalently, (3ai) holds. This proves (3b). If X=O2X=O_{2}, then Theorem 22(3) holds if and only if ν2(n+2)>ν2(n)\nu_{2}(n+2)>\nu_{2}(n), i.e., n2n\equiv 2 (mod 4). This proves (3c). The minimum PST time follows from Theorem 22. ∎

Remark 24.

If XK2X\neq K_{2} is unweighted, then we may drop the condition mσuv(L(X))m\in\sigma_{uv}^{-}(L(X)) in Theorem 23(1i).

The following result is immediate from Theorem 23(3).

Corollary 25.

Let m2m\geq 2 and ϕ(L(X),t)[t]\phi(L(X),t)\in\mathbb{Z}[t]. If m+nm+n is odd, then XYX\vee Y has no Laplacian perfect state transfer. Moreover, if mm or nn is odd and XX is either disconnected or admits Laplacian perfect state transfer, then XX does not admit Laplacian perfect state transfer in XYX\vee Y.

7.2 Adjacency case

Theorem 26.

Let m2m\geq 2, k,k,\ell\in\mathbb{Z} and ϕ(A(X),t)[t]\phi(A(X),t)\in\mathbb{Z}[t]. Vertices uu and vv in XX admit adjacency perfect state transfer in XYX\vee Y if and only if all of the following conditions hold.

  1. 1.

    Either (i) uu and vv are adjacency strongly cospectral in XX and λσuv(L(X))\lambda^{-}\notin\sigma_{uv}^{-}(L(X)) or (ii) X=O2(k)X=O_{2}(k).

  2. 2.

    One of the following conditions hold.

    1. (a)

      The eigenvalues in σu(A(X))\sigma_{u}(A(X)) are all integers and DD is a perfect square.

    2. (b)

      XX is connected, each λjσu(A(X))\{k}\lambda_{j}\in\sigma_{u}(A(X))\backslash\{k\} is of the form 12(k++bjΔ)\frac{1}{2}(k+\ell+b_{j}\sqrt{\Delta}) and λ±=12(k+±bΔ)\lambda^{\pm}=\frac{1}{2}(k+\ell\pm b\sqrt{\Delta}), where bj,b,Δb_{j},b,\Delta are integers with b>bjb>b_{j} for each jj and Δ>1\Delta>1 is square-free.

  3. 3.

    For all λ,ησuv+(A(X))\{k}\lambda,\eta\in\sigma_{uv}^{+}(A(X))\backslash\{k\}\cup\mathcal{R} and μ,θσuv(A(X))\mu,\theta\in\sigma_{uv}^{-}(A(X)), ν2(ληΔ)>ν2(λμΔ)=ν2(λθΔ)\nu_{2}\left(\frac{\lambda-\eta}{\sqrt{\Delta}}\right)>\nu_{2}\left(\frac{\lambda-\mu}{\sqrt{\Delta}}\right)=\nu_{2}\left(\frac{\lambda-\theta}{\sqrt{\Delta}}\right), where Δ=1\Delta=1 whenever (2a) holds, \mathcal{R} is given in Theorem 19(2).

Further, τXY=πgΔ\tau_{X\vee Y}=\frac{\pi}{g\sqrt{\Delta}}, where g=gcd(𝒯)g=\operatorname{gcd}\left(\mathcal{T}\right), 𝒯={λ0λΔ:λσu(A)\{k}}\mathcal{T}=\left\{\frac{\lambda_{0}-\lambda}{\sqrt{\Delta}}:\lambda\in\sigma_{u}(A)\backslash\{k\}\cup\mathcal{R}\right\} and λ0σuv+(A)\{k}\lambda_{0}\in\sigma_{uv}^{+}(A)\backslash\{k\}\cup\mathcal{R} is fixed.

Proof.

This is a direct consequence of Theorems 8(2), 19(2) and 22. ∎

The next result can be viewed as an analogue of Corollary 25.

Corollary 27.

If k+k+\ell is odd in Theorem 26, then XYX\vee Y has no adjacency perfect state transfer.

Proof.

If k+=λ++λ+k+\ell=\lambda^{+}+\lambda^{+} is odd, then so is λ+λ\lambda^{+}-\lambda^{-}, a contradiction to Theorem 26(3). ∎

8 PST in XX and XYX\vee Y

We now utilize our results from the previous section to determine when PST is preserved in the join.

8.1 Laplacian case

Corollary 28.

Let m2m\geq 2 and ϕ(L(X),t)[t]\phi(L(X),t)\in\mathbb{Z}[t]. Suppose Laplacian perfect state transfer occurs between vertices uu and vv in XX and τX=πh\tau_{X}=\frac{\pi}{h}. Then it occurs between uu and vv in XYX\vee Y if and only if mσuv(L(X))m\notin\sigma_{uv}^{-}(L(X)) and m,n0m,n\equiv 0 (mod 2α2^{\alpha}), where α>ν2(h)\alpha>\nu_{2}(h). In this case, τXY=πg\tau_{X\vee Y}=\frac{\pi}{g}, where g=gcd(𝒯)g=\operatorname{gcd}(\mathcal{T}), 𝒯\mathcal{T} is given in Theorem 23 and ν2(g)=ν2(h)\nu_{2}(g)=\nu_{2}(h). Moreover, if XX is connected, then hg\frac{h}{g} is rational. Otherwise, hg=lcm(h/h1,h/h2)\frac{h}{g}=\operatorname{lcm}(h/h_{1},h/h_{2}) is an odd integer where h1=gcd(m,h)h_{1}=\operatorname{gcd}(m,h) and h2=gcd(n,h)h_{2}=\operatorname{gcd}(n,h).

Proof.

If PST occurs in XX, then uu and vv are strongly cospectral in XX and Theorem 23(2) holds. Hence, PST occurs in XYX\vee Y if and only if (1i) and (3ai) of Theorem 23 holds. Equivalently, mσuv(L(X))m\notin\sigma_{uv}^{-}(L(X)), ν2(m)>ν2(μ)\nu_{2}(m)>\nu_{2}(\mu) and ν2(n)>ν2(μ)\nu_{2}(n)>\nu_{2}(\mu) for all μσuv(L(X))\mu\in\sigma_{uv}^{-}(L(X)). As ν2(h)=ν2(μ)=ν2(μ+n)\nu_{2}(h)=\nu_{2}(\mu)=\nu_{2}(\mu+n) for all μσuv(L(X))\mu\in\sigma_{uv}^{-}(L(X)), these conditions are equivalent to m,n0m,n\equiv 0 (mod 2α2^{\alpha}), where α>ν2(h)\alpha>\nu_{2}(h). The minimum PST time follows from Theorem 23. ∎

Corollary 29.

Let XX be an unweighted graph on m=2pm=2^{p} vertices. Then conditions (3aii-3aiii) of Theorem 23 do not hold. If we add that Laplacian perfect state transfer occurs between vertices uu and vv in XX, then:

  1. 1.

    If p=1p=1, then Laplacian perfect state transfer does not occur between uu and vv in K2YK_{2}\vee Y for any YY.

  2. 2.

    If p2p\geq 2, then Laplacian perfect state transfer occurs between uu and vv in XYX\vee Y if and only if n0n\equiv 0 (mod 2α2^{\alpha}), where α>ν2(g)\alpha>\nu_{2}(g) and πh\frac{\pi}{h} is the minimum PST time between uu and vv in XX.

Proof.

The assumption implies that each eigenvalue λ\lambda of L(X)L(X) is at most mm. Thus, ν2(λ)<ν2(m)\nu_{2}(\lambda)<\nu_{2}(m), and so conditions (3aii-iii) of Theorem 22 do not hold. Now, from Corollary 28, PST occurs between uu and vv in XYX\vee Y if and only if mσuv(L(X))m\notin\sigma_{uv}^{-}(L(X)) and n0n\equiv 0 (mod 2α2^{\alpha}), where α>ν2(h)\alpha>\nu_{2}(h). Thus, if p=1p=1, then Example 21(1) implies that uu and vv are not strongly cospectral in K2YK_{2}\vee Y, and so (1) holds. However, if p2p\geq 2, then mσuv(L(X))m\notin\sigma_{uv}^{-}(L(X)), and so (2) holds. ∎

In Corollaries 28 and 29(2), the YY with the least number of vertices and edges that works is Y=OnY=O_{n}, where n=2ν2(g)+1n=2^{\nu_{2}(g)+1}. Next, combining Corollary 28 and Theorem 23(3ai) also yields the following result.

Corollary 30.

Let uu and vv be strongly cospectral vertices in XYX\vee Y and σu(L(X))\sigma_{u}(L(X))\subseteq\mathbb{Z}. If condition (3ai) of Theorem 23 holds, then Laplacian perfect state transfer occurs between uu and vv in XX and XYX\vee Y.

From Theorem 23(3a), it is evident that PST in XYX\vee Y does not necessarily yield PST in XX. Thus, the above result can be viewed as a characterization of the equivalence of PST in a graph and its join.

Using Corollary 28, we provide an infinite family of graphs such that XX and XYX\vee Y both have PST. Recall that a graph XX is Hadamard diagonalizable if L(X)L(X) is diagonalizable by a Hadamard matrix [Barik2011].

Example 31.

Let XX be a Hadamard diagonalizable graph on m4m\geq 4 vertices with PST between uu and vv. From [Johnston2017], σuv+(L(X))\sigma_{uv}^{+}(L(X)) and σuv(L(X))\sigma_{uv}^{-}(L(X)) respectively consist of integers λ0\lambda\equiv 0 and μ2\mu\equiv 2 (mod 4). Since m0m\equiv 0 (mod 4), Corollary 28 yields PST between uu and vv in XYX\vee Y if and only if n0n\equiv 0 (mod 4).

Our next result shows that for a graph XX with PST between uu and vv, it is possible for PST to fail between uu and vv in XYX\vee Y for any graph YY, yet occur in (XZ)Y(X\cup Z)\vee Y for some graph ZZ.

Corollary 32.

Let m2m\geq 2 and ϕ(L(X),t)[t]\phi(L(X),t)\in\mathbb{Z}[t]. If Laplacian perfect state transfer occurs between vertices uu and vv in XX with mσuv(L(X))m\in\sigma_{uv}^{-}(L(X)), then the following hold.

  1. 1.

    For any graph YY, Laplacian perfect state transfer does not occur between uu and vv in XYX\vee Y.

  2. 2.

    Let ZZ be a graph on rr vertices such that m+rσuv(L(X))m+r\notin\sigma_{uv}^{-}(L(X)). Then Laplacian perfect state transfer occurs between uu and vv in (XZ)Y(X\cup Z)\vee Y if and only if ν2(n)>ν2(m)=ν2(r)\nu_{2}(n)>\nu_{2}(m)=\nu_{2}(r).

Proof.

Since mσuv(L(X))m\in\sigma_{uv}^{-}(L(X)), uu and vv are not strongly cospectral in XYX\vee Y by Theorem 19(1). We now prove (2). Since m+rσuv(L(X))m+r\notin\sigma_{uv}^{-}(L(X)), Corollary 28 implies that PST occurs between uu and vv in XYX\vee Y if and only if ν2(m+r)>ν2(m)\nu_{2}(m+r)>\nu_{2}(m) and ν2(n)>ν2(m)\nu_{2}(n)>\nu_{2}(m) for all μσuv(L(X))\mu\in\sigma_{uv}^{-}(L(X)). Equivalently, ν2(n)>ν2(m)=ν2(r)\nu_{2}(n)>\nu_{2}(m)=\nu_{2}(r). ∎

In Corollary 32(2), the ZZ and YY with the least number of vertices and edges that work are Z=OrZ=O_{r} and Y=OnY=O_{n} respectively, where r=2ν2(m)r=2^{\nu_{2}(m)} and r=2ν2(m)+1r=2^{\nu_{2}(m)+1}.

Example 33.

Let X=K2X=K_{2} with vertices uu and vv. From Corollary 29(1), PST between uu and vv fails in K2YK_{2}\vee Y for any graph YY. Since r+2σuv(L(K2))r+2\notin\sigma_{uv}^{-}(L(K_{2})) for all r1r\geq 1, Corollary 32 yields PST between uu and vv in (K2Z)Y(K_{2}\cup Z)\vee Y if and only if ν2(n)>ν2(r)=1\nu_{2}(n)>\nu_{2}(r)=1. In particular, we may take Z=O2Z=O_{2} and Y=O4Y=O_{4}.

The following remark can be used to generate infinite families of weighted graphs XX whereby PST occurs between uu and vv in XX and (XZ)Y(X\cup Z)\vee Y for some graph ZZ, but not in XYX\vee Y for any graph YY.

Remark 34.

Let m2m\geq 2 and ϕ(L(X),t)[t]\phi(L(X),t)\in\mathbb{Z}[t]. Suppose vertices uu and vv admit Laplacian PST in XX, and suppose μ\mu divides mm for some μσuv(L(X))\mu\in\sigma_{uv}^{-}(L(X)). Let XX^{\prime} be the graph obtained by scaling all edge weights of XX by a factor of mμ\frac{m}{\mu}. Then mσuv(L(X))m\in\sigma_{uv}^{-}(L(X^{\prime})), and so PST does not occur between uu and vv in XYX^{\prime}\vee Y for any graph YY by Corollary 32(1). But since mμ\frac{m}{\mu} is an integer, we get ϕ(L(X),t)[t]\phi(L(X^{\prime}),t)\in\mathbb{Z}[t], and so we may apply Corollary 32(2) to obtain a graph ZZ such that PST occurs between uu and vv in (XZ)Y(X^{\prime}\cup Z)\vee Y.

8.2 Adjacency case

In contrast with the Laplacian case in Corollary 32(2), it turns out that if XX is regular and has adjacency PST between two vertices, then we can always choose a regular graph YY such that adjacency PST is preserved in XYX\vee Y.

Corollary 35.

Let m2m\geq 2, k,k,\ell\in\mathbb{Z} and ϕ(A(X),t)[t]\phi(A(X),t)\in\mathbb{Z}[t]. Suppose adjacency perfect state transfer occurs between vertices uu and vv in XX and τX=πh\tau_{X}=\frac{\pi}{h}. Then it occurs between them in XYX\vee Y if and only if

  1. 1.

    n=s(k+s)mn=\frac{s(k-\ell+s)}{m} for some integer ss such that mm divides s(k+s)s(k-\ell+s) and sσuv(A(X))\ell-s\notin\sigma_{uv}^{-}(A(X)), and

  2. 2.

    for all μσuv(A(X))\mu\in\sigma_{uv}^{-}(A(X)), ν2(s)>ν2(kμ)\nu_{2}(s)>\nu_{2}(k-\mu) and ν2(k)>ν2(kμ)\nu_{2}(\ell-k)>\nu_{2}(k-\mu).

In this case, τXY=πg\tau_{X\vee Y}=\frac{\pi}{g}, where g=gcd(𝒯)g=\operatorname{gcd}(\mathcal{T}), 𝒯\mathcal{T} is given in Theorem 26, λ0σuv+(A)\{k}\lambda_{0}\in\sigma_{uv}^{+}(A)\backslash\{k\}\cup\mathcal{R} is fixed and ν2(g)=ν2(h)\nu_{2}(g)=\nu_{2}(h). Moreover, if XX is connected, hg\frac{h}{g} is rational. Otherwise, hg=lcm(h/h1,h/h2)\frac{h}{g}=\operatorname{lcm}(h/h_{1},h/h_{2}) is an odd integer and h1=gcd(λ0λ+,h)h_{1}=\operatorname{gcd}(\lambda_{0}-\lambda^{+},h) and h2=gcd(λ0λ,h)h_{2}=\operatorname{gcd}(\lambda_{0}-\lambda^{-},h).

Proof.

From Theorem 26, it follows that PST occurs between uu and vv in XYX\vee Y if and only if

  • (a)

    DD is a perfect square and λσuv(A(X))\lambda^{-}\notin\sigma_{uv}^{-}(A(X)), and

  • (b)

    ν2(λ±η)>ν2(λ±μ)=ν2(λ±θ)\nu_{2}(\lambda^{\pm}-\eta)>\nu_{2}(\lambda^{\pm}-\mu)=\nu_{2}(\lambda^{\pm}-\theta) for all ησuv+(A(X))\{k}𝒮\eta\in\sigma_{uv}^{+}(A(X))\backslash\{k\}\cup\mathcal{S} and μ,θσuv(A(X))\mu,\theta\in\sigma_{uv}^{-}(A(X)).

By the choice of ss in (1), D=(k)2+4mn=(k+2s)2D=(k-\ell)^{2}+4mn=(k-\ell+2s)^{2}. Since λ±=12(k+±D)\lambda^{\pm}=\frac{1}{2}(k+\ell\pm\sqrt{D}), we get λ+=k+s\lambda^{+}=k+s and λ=s\lambda^{-}=\ell-s. Thus, (1) ensures that condition (a) above holds. Using the fact that XX has PST between uu and vv, Theorem 22(3) yields

ν2(kη)>ν2(kμ)=ν2(kθ)\nu_{2}(k-\eta)>\nu_{2}(k-\mu)=\nu_{2}(k-\theta) (15)

for all ησuv+(A(X))\{k}\eta\in\sigma_{uv}^{+}(A(X))\backslash\{k\} and μ,θσuv(A(X))\mu,\theta\in\sigma_{uv}^{-}(A(X)). Observe that for each λσu(A(X))\{k}\lambda\in\sigma_{u}(A(X))\backslash\{k\}, we can write λ+λ=(λ+k)+(kλ)=s+(kλ)\lambda^{+}-\lambda=(\lambda^{+}-k)+(k-\lambda)=s+(k-\lambda). Similarly, λλ=(sk)+(kλ)\lambda^{-}-\lambda=(\ell-s-k)+(k-\lambda). Thus,

ν2(λ+λ)min{ν2(s),ν2(kλ)}andν2(λλ)min{ν2(sk),ν2(kλ)}.\nu_{2}(\lambda^{+}-\lambda)\geq\min\{\nu_{2}(s),\nu_{2}(k-\lambda)\}\quad\text{and}\quad\nu_{2}(\lambda^{-}-\lambda)\geq\min\{\nu_{2}(\ell-s-k),\nu_{2}(k-\lambda)\}. (16)

We have the following cases.

  • Suppose ν2(s)>ν2(kμ)\nu_{2}(s)>\nu_{2}(k-\mu) for some μσuv(A(X))\mu\in\sigma_{uv}^{-}(A(X)). Then (15) implies that ν2(s)>ν2(kμ)\nu_{2}(s)>\nu_{2}(k-\mu) for all μσuv(A(X))\mu\in\sigma_{uv}^{-}(A(X)), and combining this with (16) yields ν2(λ+μ)=ν2(λ+θ)=ν2(kμ)\nu_{2}(\lambda^{+}-\mu)=\nu_{2}(\lambda^{+}-\theta)=\nu_{2}(k-\mu). Moreover, (15) and (16) give us ν2(λ+η)>ν2(kμ)\nu_{2}(\lambda^{+}-\eta)>\nu_{2}(k-\mu) for all μσuv(A(X))\mu\in\sigma_{uv}^{-}(A(X)), from which we get ν2(λ+η)>ν2(λ+μ)=ν2(λ+θ)\nu_{2}(\lambda^{+}-\eta)>\nu_{2}(\lambda^{+}-\mu)=\nu_{2}(\lambda^{+}-\theta) for all ησuv+(A(X))\{k}\eta\in\sigma_{uv}^{+}(A(X))\backslash\{k\} and μ,θσuv(A(X))\mu,\theta\in\sigma_{uv}^{-}(A(X)).

  • Suppose ν2(s)ν2(kμ)\nu_{2}(s)\leq\nu_{2}(k-\mu) for some μσuv(A(X))\mu\in\sigma_{uv}^{-}(A(X)). Then one can check using (15) and (16) that ν2(λ+η)=ν2(s)ν2(λ+μ)\nu_{2}(\lambda^{+}-\eta)=\nu_{2}(s)\leq\nu_{2}(\lambda^{+}-\mu), a violation of condition (b).

Consequently, ν2(λ+η)>ν2(λ+μ)=ν2(λ+θ)\nu_{2}(\lambda^{+}-\eta)>\nu_{2}(\lambda^{+}-\mu)=\nu_{2}(\lambda^{+}-\theta) for all ησuv+(A(X))\{k}\eta\in\sigma_{uv}^{+}(A(X))\backslash\{k\} and μ,θσuv(A(X))\mu,\theta\in\sigma_{uv}^{-}(A(X)) if and only if ν2(s)>ν2(kμ)\nu_{2}(s)>\nu_{2}(k-\mu) for all μσuv(A(X))\mu\in\sigma_{uv}^{-}(A(X)). Using this fact and again arguing similarly as in the above two cases using (15) and (16) yields ν2(λ+η)>ν2(λ+μ)=ν2(λ+θ)\nu_{2}(\lambda^{+}-\eta)>\nu_{2}(\lambda^{+}-\mu)=\nu_{2}(\lambda^{+}-\theta) for all ησuv+(A(X))\{k}\eta\in\sigma_{uv}^{+}(A(X))\backslash\{k\} and μ,θσuv(A(X))\mu,\theta\in\sigma_{uv}^{-}(A(X)) if and only if ν2(k)>ν2(kμ)\nu_{2}(\ell-k)>\nu_{2}(k-\mu) for all μσuv(A(X))\mu\in\sigma_{uv}^{-}(A(X)). Thus, the assumption in (2) is equivalent to condition (b) above, which yields the desired conclusion. ∎

The following example illustrates Corollary 35.

Example 36.

Let XX be a simple integer-weighted Hadamard diagonalizable graph on m4m\geq 4 vertices with PST between vertices uu and vv. Then m0m\equiv 0 (mod 4) and XX is pp-regular for some integer pp. From Example 31, σuv+(A(X))\sigma_{uv}^{+}(A(X)) consists of integers pλp-\lambda, where λ0\lambda\equiv 0 (mod 4), while σuv(A(X))\sigma_{uv}^{-}(A(X)) consists of integers pμp-\mu such that μ2\mu\equiv 2 (mod 4). Invoking Corollary 35(2), we get PST between uu and vv in XYX\vee Y if and only if (i) n=s(p+s)mn=\frac{s(p-\ell+s)}{m}, where ss is an integer such that mm divides s(p+s)s(p-\ell+s) and sσuv(A(X))\ell-s\notin\sigma_{uv}^{-}(A(X)) and (ii) ν2(s)2\nu_{2}(s)\geq 2 and ν2(p)2\nu_{2}(\ell-p)\geq 2. In particular, if X=QpX=Q_{p}, then we may take =p\ell=p and s=2(p+1)/2as=2^{(p+1)/2}a (i.e., n=2an=2a) if pp is odd and s=2p/2as=2^{p/2}a (i.e., n=an=a) otherwise, to get PST in both QpQ_{p} and QpYQ_{p}\vee Y.

Remark 37.

If X=K2X=K_{2}, then k=μ=1k=-\mu=1, and so the condition sσuv(A(X))\ell-s\notin\sigma_{uv}^{-}(A(X)) in Corollary 35(1) is equivalent to n1\ell\neq n-1, while the conditions ν2(s)>1\nu_{2}(s)>1 and ν2(1)>1\nu_{2}(\ell-1)>1 in Corollary 35(2) are equivalent to ν2(s2)=1\nu_{2}(s-2)=1 and ν2(+3)>1\nu_{2}(\ell+3)>1. Thus, taking X=K2X=K_{2} in Corollary 35 recovers a known characterization of PST in connected double cones [Kirkland2023, Theorem 12(1)].

Next, we have an analogue of Corollary 30, which determines when PST in XYX\vee Y is inherited by XX.

Corollary 38.

Let m2m\geq 2, k,k,\ell\in\mathbb{Z} and ϕ(A(X),t)[t]\phi(A(X),t)\in\mathbb{Z}[t].

  1. 1.

    Let XX be disconnected. If perfect state transfer occurs between vertices uu and vv in XYX\vee Y, then it also occurs between them in XX if and only if XO2(k)X\neq O_{2}(k) for any kk\in\mathbb{Z}.

  2. 2.

    Suppose XX is connected. If perfect state transfer occurs between vertices uu and vv in XYX\vee Y, then perfect state transfer occurs between uu and vv in XX if and only if both conditions below hold.

    1. (a)

      The eigenvalues in σu(A(X))\sigma_{u}(A(X)) are all integers and DD is a perfect square.

    2. (b)

      ν2(kη)>ν2(kμ)=ν2(kθ)\nu_{2}(k-\eta)>\nu_{2}(k-\mu)=\nu_{2}(k-\theta) for any ησuv+(A(X))\{k}\eta\in\sigma_{uv}^{+}(A(X))\backslash\{k\} and μ,θσuv(A(X))\mu,\theta\in\sigma_{uv}^{-}(A(X)).

9 PST in XYX\vee Y but not in XX

In Corollary 28, we characterized graphs with PST that also admit PST in the join. In this section, our goal is to show that PST can be induced in the join despite the lack of PST in the underlying graph.

9.1 Laplacian case

The following is immediate from Theorem 23.

Corollary 39.

Let m2m\geq 2 and ϕ(L(X),t)[t]\phi(L(X),t)\in\mathbb{Z}[t]. Suppose uu and vv are strongly cospectral vertices in XX with σu(L(X))\sigma_{u}(L(X))\subseteq\mathbb{Z}. Then Laplacian perfect state transfer occurs between uu and vv in XYX\vee Y but not in XX if and only if mσuv(L(X))m\notin\sigma_{uv}^{-}(L(X)), ν2(m)=ν2(n)\nu_{2}(m)=\nu_{2}(n) and either (i) Theorem 23(3aii-3aiii) holds or (ii) X=O2X=O_{2}.

For strongly cospectral vertices in XX whose elements in σuv(L(X))\sigma_{uv}^{-}(L(X)) consists of integers whose largest powers of two in their factorizations are larger than those in σuv+(L(X))\sigma_{uv}^{+}(L(X)), Theorem 22 yields no PST between them. Nonetheless, we show that Laplacian PST can be induced in XYX\vee Y by appropriate choice of nn. The following result follows directly from Corollary 39 and Theorem 23(3aii).

Theorem 40.

Let m2m\geq 2 and ϕ(L(X),t)[t]\phi(L(X),t)\in\mathbb{Z}[t]. Suppose uu and vv are strongly cospectral vertices in XX with σu(L)\sigma_{u}(L)\subseteq\mathbb{Z} such that the ν2(μ)>ν2(λ)\nu_{2}(\mu)>\nu_{2}(\lambda) for all 0λσuv+(L(X))0\neq\lambda\in\sigma_{uv}^{+}(L(X)) and μσuv(L(X))\mu\in\sigma_{uv}^{-}(L(X)). Then Laplacian perfect state transfer occurs between uu and vv in XYX\vee Y if and only if mσuv(L(X))m\notin\sigma_{uv}^{-}(L(X)), XX is connected and ν2(λ)=ν2(m)=ν2(n)\nu_{2}(\lambda)=\nu_{2}(m)=\nu_{2}(n) for all 0λσuv+(L(X))0\neq\lambda\in\sigma_{uv}^{+}(L(X)).

Example 41.

Let X=CP(m)X=CP(m), where m2m\equiv 2 (mod 4). Then XX is connected, (m2)(m-2)-regular and any two non-adjacent vertices uu and vv in XX are strongly cospectral with σuv+(L(X))={0,m}\sigma_{uv}^{+}(L(X))=\{0,m\} and σuv(L(X))={m2}\sigma_{uv}^{-}(L(X))=\{m-2\}. As ν2(m2)>ν2(m)\nu_{2}(m-2)>\nu_{2}(m), Laplacian PST does not occur between uu and vv in XX by Theorem 22(3). Applying Theorem 40, Laplacian PST occurs between uu and vv in XYX\vee Y if and only if n2n\equiv 2 (mod 4). In particular, if we take n=2n=2, then XY=CP(m+2)X\vee Y=CP(m+2) admits PST between any two non-adjacent vertices.

For strongly cospectral vertices in XX whose eigenvalue supports consist of integers with equal largest powers of two in their factorizations, Theorem 22 again yields no PST between them. Despite this fact, we show that under certain assumptions, Laplacian PST can be induced in XYX\vee Y by appropriate choice of nn.

Theorem 42.

Let m2m\geq 2 and ϕ(L(X),t)[t]\phi(L(X),t)\in\mathbb{Z}[t]. Suppose uu and vv are strongly cospectral vertices in XX with σu(L)\sigma_{u}(L)\subseteq\mathbb{Z} such that the ν2(λ)\nu_{2}(\lambda)’s are equal for all λσu(L)\lambda\in\sigma_{u}(L), say to α\alpha. Write each 0λrσuv+(L(X))0\neq\lambda_{r}\in\sigma_{uv}^{+}(L(X)) as λr=2α(2pr1)\lambda_{r}=2^{\alpha}(2p_{r}-1) and each μsσuv(L(X))\mu_{s}\in\sigma_{uv}^{-}(L(X)) as μs=2α(2qs1)\mu_{s}=2^{\alpha}(2q_{s}-1) for some pr,qsp_{r},q_{s}\in\mathbb{Z}. Then Laplacian perfect state transfer occurs between uu and vv in XYX\vee Y if and only if XX is connected, mσuv(L(X))m\notin\sigma_{uv}^{-}(L(X)), ν2(m)=ν2(n)=α\nu_{2}(m)=\nu_{2}(n)=\alpha (so we may write m=2α(2y1)m=2^{\alpha}(2y-1) and n=2α(2z+1)n=2^{\alpha}(2z+1) for some y,zy,z\in\mathbb{Z}), and one of the following conditions below hold.

  1. 1.

    All ν2(qs)\nu_{2}(q_{s})’s are equal, ν2(pr)>ν2(qs)\nu_{2}(p_{r})>\nu_{2}(q_{s}), and ν2(y)ν2(z)ν2(qs)+1\nu_{2}(y)\geq\nu_{2}(z)\geq\nu_{2}(q_{s})+1 for all r,sr,s.

  2. 2.

    All ν2(pr)\nu_{2}(p_{r})’s are equal and ν2(pr)=ν2(y)=ν2(z)<ν2(qs)\nu_{2}(p_{r})=\nu_{2}(y)=\nu_{2}(z)<\nu_{2}(q_{s}) for all ss.

  3. 3.

    For any r,s,tr,s,t, we have ν2(pr)=ν2(qs)=ν2(y)=ν2(z)\nu_{2}(p_{r})=\nu_{2}(q_{s})=\nu_{2}(y)=\nu_{2}(z), ν2(pr+z2ν2(z))>ν2(qs+z2ν2(z))=ν2(qt+z2ν2(z))\nu_{2}\left(\frac{p_{r}+z}{2^{\nu_{2}(z)}}\right)>\nu_{2}\left(\frac{q_{s}+z}{2^{\nu_{2}(z)}}\right)=\nu_{2}\left(\frac{q_{t}+z}{2^{\nu_{2}(z)}}\right) and ν2(y+z2ν2(z))>ν2(qs+z2ν2(z))\nu_{2}\left(\frac{y+z}{2^{\nu_{2}(z)}}\right)>\nu_{2}\left(\frac{q_{s}+z}{2^{\nu_{2}(z)}}\right).

Proof.

Corollary 39 implies that PST occurs between uu and vv in XYX\vee Y if and only if ν2(m)=ν2(n)=α\nu_{2}(m)=\nu_{2}(n)=\alpha and Theorem 23(3aii) holds. Let m=2α(2y1)m=2^{\alpha}(2y-1) and n=2α(2z+1)n=2^{\alpha}(2z+1) for some integers y,zy,z. Note that m+n=2α+1(y+z)m+n=2^{\alpha+1}(y+z), λr+n=2α+1(pr+z)\lambda_{r}+n=2^{\alpha+1}(p_{r}+z) and μs+n=2α+1(qs+z)\mu_{s}+n=2^{\alpha+1}(q_{s}+z). If XX is disconnected, then nσuv+(L(X))n\in\sigma_{uv}^{+}(L(X)). Now, Theorem 23(3aii) holds if and only if XX is connected and

ν2(pr+z)>ν2(qs+z)=ν2(qt+z)andν2(y+z)>ν2(qs+z)\nu_{2}(p_{r}+z)>\nu_{2}(q_{s}+z)=\nu_{2}(q_{t}+z)\quad\text{and}\quad\nu_{2}(y+z)>\nu_{2}(q_{s}+z) (17)

for any r,s,tr,s,t. One can then proceed by cases similar to the proof of Theorem 23(2a). ∎

Remark 43.

Let XX be a simple unweighted graph with strongly cospectral vertices uu and vv such that σu(L)\sigma_{u}(L)\subseteq\mathbb{Z}. If XX has an odd number of vertices and an odd number of spanning trees, then the matrix-tree theorem implies that all nonzero eigenvalues in σu(L)\sigma_{u}(L) are odd. Thus, Theorem 42 applies with α=0\alpha=0.

We illustrate Theorem 42 and Remark 43 using a simple example.

Example 44.

Let X=P3X=P_{3} with end vertices uu and vv. Then σuv+(L(X))={0,3}\sigma_{uv}^{+}(L(X))=\{0,3\}, σuv(L(X))={1}\sigma_{uv}^{-}(L(X))=\{1\}, and so by Theorem 22(2), uu and vv does not admit Laplacian PST in XX. Applying Theorem 42(1), we get Laplacian PST between uu and vv in XYX\vee Y if and only if n1n\equiv 1 (mod 4). In particular, taking Y=OnY=O_{n} yields XY=Kd\eX\vee Y=K_{d}\backslash e with d0d\equiv 0 (mod 4), a graph known to have PST between non-adjacent vertices.

We close this subsection with a result that determines all graphs with isolated vertices that exhibit Laplacian PST in the join. It is immediate from Theorems 19(1a) and 23(3c).

Theorem 45.

If uu and vv are two isolated vertices in XX, then Laplacian perfect state transfer occurs between uu and vv in XYX\vee Y if and only if X=O2X=O_{2} and n2n\equiv 2 (mod 4).

From Theorem 45, O2O_{2} is the only graph with isolated vertices that exhibits PST in the join. This coincides with a known characterization of Laplacian PST in double cones [Alvir2016, Corollary 4].

9.2 Adjacency case

For the adjacency case, we give one scenario where PST can be induced in XYX\vee Y. The following is immediate from Theorems 22(3) and 26(3).

Corollary 46.

Let m2m\geq 2, k,k,\ell\in\mathbb{Z} and ϕ(A(X),t)[t]\phi(A(X),t)\in\mathbb{Z}[t]. Suppose XX is connected and vertices uu and vv are strongly cospectral in XX such that σu(A(X))\sigma_{u}(A(X))\subseteq\mathbb{Z} and

ν2(λη)>ν2(λμ)=ν2(λθ),\nu_{2}\left(\lambda-\eta\right)>\nu_{2}\left(\lambda-\mu\right)=\nu_{2}\left(\lambda-\theta\right), (18)

for all λ,ησuv+(A(X))\{k}\lambda,\eta\in\sigma_{uv}^{+}(A(X))\backslash\{k\} and μ,θσuv(A(X))\mu,\theta\in\sigma_{uv}^{-}(A(X)). If (18) does not hold whenever λ=k\lambda=k, then

  1. 1.

    adjacency perfect state transfer does not occur between uu and vv in XX; and

  2. 2.

    adjacency perfect state transfer occurs between uu and vv in XYX\vee Y if and only if λσuv(A(X))\lambda^{-}\notin\sigma_{uv}^{-}(A(X)), DD is a perfect square and (18) also holds whenever λ{λ±}\lambda\in\{\lambda^{\pm}\}.

Using Corollary 39, we generate an infinite family of graphs with no PST but which admit PST in the join.

Example 47.

Let m6m\geq 6 and X=CP(m)X=CP(m), where m2m\equiv 2 (mod 4). From Example 41, XX is connected, (m2)(m-2)-regular and PST does not occur between uu and vv in XX by Corollary 46(1). Applying Corollary 46(2), we get adjacency PST between uu and vv in XYX\vee Y if and only if n=z(m2+z)mn=\frac{z(m-\ell-2+z)}{m} for some integer zz such that mm divides z(m2+z)z(m-\ell-2+z), ν2()>ν2(z)=1\nu_{2}(\ell)>\nu_{2}(z)=1 and z0\ell-z\neq 0. The latter three conditions are equivalent to DD being a perfect square, ν2(λ+λ)=ν2(λ0)=2\nu_{2}(\lambda^{+}-\lambda^{-})=\nu_{2}(\lambda^{-}-0)=2 and λ0\lambda^{-}\neq 0, respectively. In particular, we may take =0\ell=0 and z=2z=2 to get n=2n=2. In this case, XY=CP(m+2)X\vee Y=CP(m+2) which admits PST between any two non-adjacent vertices.

We finish this section with an analogue of Theorem 45. It is immediate from Theorem 26.

Theorem 48.

If uu and vv are two isolated vertices in XX, then adjacency perfect state transfer occurs between uu and vv in XYX\vee Y if and only if X=O2(k)X=O_{2}(k), DD is a perfect square and ν2(λ+k)=ν2(λk)\nu_{2}(\lambda^{+}-k)=\nu_{2}(\lambda^{-}-k).

10 Families of joins

We now determine quantum state transfer in self-joins and iterated join graphs.

10.1 Self-joins

Denote the rr-fold join of XX with itself by Xr:=j=1rXX^{r}:=\bigvee_{j=1}^{r}X. The following characterizes periodicity in XrX^{r}. It follows from Corollaries 7 and 9, and Theorem 11.

Theorem 49.

Let M{A,L}M\in\{A,L\} and ϕ(M(X),t)[t]\phi(M(X),t)\in\mathbb{Z}[t]. Vertex uu is periodic in XX if and only if it is periodic in XrX^{r} for all rr, if and only if σu(M(X))\sigma_{u}(M(X))\subseteq\mathbb{Z}. Moreover, XX is periodic if and only if XrX^{r} is periodic for all rr, if and only if XX is (Laplacian) integral.

Next, we characterize strong cospectrality in XrX^{r}.

Theorem 50.

Suppose vertices uu and vv are strongly cospectral vertices in XX with respect to MM.

  1. 1.

    If mσuv(L(X))m\notin\sigma_{uv}^{-}(L(X)), then uu and vv are Laplacian strongly cospectral in XrX^{r} for all r1r\geq 1 with

    σuv+(L)={λ+(r1)m:λσuv+(L(X))\{0}}andσuv(L)={μ+(r1)m:μσuv(L(X))},\sigma_{uv}^{+}(L)=\{\lambda+(r-1)m:\lambda\in\sigma_{uv}^{+}(L(X))\backslash\{0\}\}\cup\mathcal{R}\quad\text{and}\quad\sigma_{uv}^{-}(L)=\{\mu+(r-1)m:\mu\in\sigma_{uv}^{-}(L(X))\},

    where ={0,rm}\mathcal{R}=\{0,rm\} if XX is connected and ={0,rm,(r1)m}\mathcal{R}=\{0,rm,(r-1)m\} otherwise. If we add that XK2X\neq K_{2} is unweighted, then uu and vv are strongly cospectral in XrX^{r} for all r1r\geq 1.

  2. 2.

    If kmσuv(A(X))k-m\notin\sigma_{uv}^{-}(A(X)), then uu and vv are adjacency strongly cospectral in XrX^{r} for all r1r\geq 1 with

    σuv+(A)=σuv+(A(X))\{k}andσuv(A)=σuv(A(X)),\sigma_{uv}^{+}(A)=\sigma_{uv}^{+}(A(X))\backslash\{k\}\cup\mathcal{R}\quad\text{and}\quad\sigma_{uv}^{-}(A)=\sigma_{uv}^{-}(A(X)),

    where ={km,k+(r1)m}\mathcal{R}=\{k-m,k+(r-1)m\} if XX is connected and ={km,k+(r1)m,k}\mathcal{R}=\{k-m,k+(r-1)m,k\} otherwise.

Proof.

Note that (1) follows from Theorem 19(1) and Remark 20. To prove (2), note that XrX^{r} is (k+(r1)m)(k+(r-1)m)-regular with rmrm vertices. Taking Y=Xr1Y=X^{r-1}, we get Xr=XYX^{r}=X\vee Y and D=((k+(r2)m)k)2+4(r1)m2=r2m2D=((k+(r-2)m)-k)^{2}+4(r-1)m^{2}=r^{2}m^{2}, and so λ=km\lambda^{-}=k-m and λ+=k+(r1)m\lambda^{+}=k+(r-1)m. Invoking Theorem 19(2) yields (2). ∎

Taking Y=Xr1Y=X^{r-1} in Theorem 23 yields a characterization of Laplacian PST in rr-fold joins.

Theorem 51.

Let r,m2r,m\geq 2 and ϕ(L(X),t)[t]\phi(L(X),t)\in\mathbb{Z}[t]. Vertices uu and vv in XX admit Laplacian perfect state transfer in XrX^{r} if and only if either:

  1. 1.

    X=O2X=O_{2}, in which case Xr=CP(2r)X^{r}=CP(2r), where rr is even.

  2. 2.

    Vertices uu and vv belong to the same connected component of XX and all of the conditions below hold.

    1. (a)

      Vertices uu and vv are Laplacian strongly cospectral in XX and mσuv(L(X))m\notin\sigma_{uv}^{-}(L(X)).

    2. (b)

      The eigenvalues in σu(L(X))\sigma_{u}(L(X)) are all integers.

    3. (c)

      One of the following conditions hold.

      1. i.

        XX is connected and one of the following conditions hold for all 0λσuv+(L(X)){m}0\neq\lambda\in\sigma_{uv}^{+}(L(X))\cup\{m\} and μ,θσuv(L(X))\mu,\theta\in\sigma_{uv}^{-}(L(X)).

        1. A.

          ν2(λ)>ν2(μ)=ν2(θ)\nu_{2}(\lambda)>\nu_{2}(\mu)=\nu_{2}(\theta) and rr is any integer.

        2. B.

          ν2(μ)>ν2(λ)\nu_{2}(\mu)>\nu_{2}(\lambda) and rr is even.

        3. C.

          ν2(λ)=ν2(μ)\nu_{2}(\lambda)=\nu_{2}(\mu), ν2(λ+(r1)m2ν2(m))>ν2(μ+(r1)m2ν2(m))=ν2(θ+(r1)m2ν2(m))\nu_{2}\left(\frac{\lambda+(r-1)m}{2^{\nu_{2}(m)}}\right)>\nu_{2}\left(\frac{\mu+(r-1)m}{2^{\nu_{2}(m)}}\right)=\nu_{2}\left(\frac{\theta+(r-1)m}{2^{\nu_{2}(m)}}\right) and rr is even.

      2. ii.

        XO2X\neq O_{2} is disconnected and (iA) above holds.

If these conditions hold, then τXr=πg\tau_{X^{r}}=\frac{\pi}{g}, where g=gcd(𝒯)g=\operatorname{gcd}\left(\mathcal{T}\right), 𝒯={rm}{λm:λσu(L(X))\{0}}\mathcal{T}=\{rm\}\cup\{\lambda-m:\lambda\in\sigma_{u}(L(X))\backslash\{0\}\} if XX is connected, and 𝒯={m}σu(L(X))\{0}\mathcal{T}=\{m\}\cup\sigma_{u}(L(X))\backslash\{0\} otherwise.

Example 52.

Let X=P3X=P_{3} with end vertices uu and vv, and consider XrX^{r} for any integer rr. From Example 44, σuv+(L(X))={0,3}\sigma_{uv}^{+}(L(X))=\{0,3\}, σuv(L(X))={1}\sigma_{uv}^{-}(L(X))=\{1\} and Laplacian PST does not occur in XX. Since λ=m=3\lambda=m=3 and μ=1\mu=1 are odd, we have ν2(λ+(r1)m)=ν2(r)\nu_{2}(\lambda+(r-1)m)=\nu_{2}(r) and ν2(μ+(r1)m)=1\nu_{2}(\mu+(r-1)m)=1. By Theorem 51(2ci), PST occurs between uu and vv in XrX^{r} if and only if r0r\equiv 0 (mod 4). In this case, P3r=(O2O1)r=CP(2r)KrP_{3}^{r}=(O_{2}\vee O_{1})^{r}=CP(2r)\vee K_{r}. Indeed, Theorem 51(1) yields PST in CP(2r)CP(2r) between non-adjacent vertices with minimum time π2\frac{\pi}{2}, and as r0r\equiv 0 (mod 4), Corollary 28 implies that this PST is preserved in CP(2r)KrCP(2r)\vee K_{r} at the same time.

Using Corollary 28, we characterize when Laplacian PST in preserved in self-joins.

Corollary 53.

Let m2m\geq 2 and ϕ(L(X),t)[t]\phi(L(X),t)\in\mathbb{Z}[t]. Suppose Laplacian perfect state transfer occurs between vertices uu and vv in XX and τX=πh\tau_{X}=\frac{\pi}{h}. Then it occurs between them in XrX^{r} if and only if mσuv(L(X))m\notin\sigma_{uv}^{-}(L(X)) and m0m\equiv 0 (mod 2α2^{\alpha}), where α>ν2(h)\alpha>\nu_{2}(h). In this case, τXr=πg\tau_{X^{r}}=\frac{\pi}{g}, where g=gcd(𝒯)g=\operatorname{gcd}(\mathcal{T}), 𝒯\mathcal{T} is given in Theorem 51 and ν2(g)=ν2(h)\nu_{2}(g)=\nu_{2}(h). Moreover, if XX is connected, then hg\frac{h}{g} is rational. Otherwise, hg=hgcd(h,m)\frac{h}{g}=\frac{h}{\operatorname{gcd}(h,m)} is odd.

In Theorem 51(2a) and Corollary 53, we may drop the condition mσuv(L(X))m\notin\sigma_{uv}^{-}(L(X)) if XK2X\neq K_{2} is unweighted. Taking Y=Xr1Y=X^{r-1} in Theorem 26 yields a characterization of adjacency PST in rr-fold joins.

Theorem 54.

Let r,m2r,m\geq 2, kk\in\mathbb{Z} and ϕ(A(X),t)[t]\phi(A(X),t)\in\mathbb{Z}[t]. If XX is simple, then vertices uu and vv in XX admit adjacency perfect state transfer in XrX^{r} if and only if either:

  1. 1.

    X=O2X=O_{2}, in which case Xr=CP(2r)X^{r}=CP(2r), where rr is even.

  2. 2.

    Vertices uu and vv belong to the same connected component of XX and all of the conditions below hold.

    1. (a)

      Vertices uu and vv are adjacency strongly cospectral in XX and kmσuv(L(X))k-m\notin\sigma_{uv}^{-}(L(X)).

    2. (b)

      The eigenvalues in σu(A(X))\sigma_{u}(A(X)) are all integers.

    3. (c)

      One of the following conditions hold.

      1. i.

        XX is connected and one of the following conditions hold for all 0λσuv+(A(X))\{k}0\neq\lambda\in\sigma_{uv}^{+}(A(X))\backslash\{k\} and μ,θσuv(A(X))\mu,\theta\in\sigma_{uv}^{-}(A(X)).

        1. A.

          ν2(m)>ν2(kμ)\nu_{2}(m)>\nu_{2}(k-\mu), ν2(kλ)>ν2(kμ)\nu_{2}(k-\lambda)>\nu_{2}(k-\mu) and rr is any integer.

        2. B.

          ν2(kμ)>ν2(m)=ν2(kλ)\nu_{2}(k-\mu)>\nu_{2}(m)=\nu_{2}(k-\lambda) and rr is even

        3. C.

          ν2(m)=ν2(kη)=ν2(kμ)\nu_{2}(m)=\nu_{2}(k-\eta)=\nu_{2}(k-\mu), ν2(kmλ2ν2(m))>ν2(kmμ2ν2(m))=ν2(kmθ2ν2(m))\nu_{2}\left(\frac{k-m-\lambda}{2^{\nu_{2}(m)}}\right)>\nu_{2}\left(\frac{k-m-\mu}{2^{\nu_{2}(m)}}\right)=\nu_{2}\left(\frac{k-m-\theta}{2^{\nu_{2}(m)}}\right) and rr is an integer such that ν2(r)>ν2(kmμ2ν2(m))\nu_{2}(r)>\nu_{2}\left(\frac{k-m-\mu}{2^{\nu_{2}(m)}}\right).

      2. ii.

        XO2X\neq O_{2} is disconnected and (iA) above holds.

If these conditions hold, then τXr=πg\tau_{X^{r}}=\frac{\pi}{g}, where g=gcd(𝒯)g=\operatorname{gcd}\left(\mathcal{T}\right), 𝒯={rm}{kmλ:λσu(A(X))\{k}}\mathcal{T}=\{rm\}\cup\{k-m-\lambda:\lambda\in\sigma_{u}(A(X))\backslash\{k\}\} if XX is connected and 𝒯={m}{kλ:λσu(A(X))\{k}}\mathcal{T}=\{m\}\cup\{k-\lambda:\lambda\in\sigma_{u}(A(X))\backslash\{k\}\} otherwise.

Example 55.

Let X=CP(m)X=CP(m), where m2m\equiv 2 (mod 4). From Example 41, σuv+(A(X))={m2,2}\sigma_{uv}^{+}(A(X))=\{m-2,-2\}, σuv(A(X))={0}\sigma_{uv}^{-}(A(X))=\{0\} and adjacency PST does not occur between non-adjacent vertices uu and vv. If we let k=m2k=m-2, λ=2\lambda=-2 and μ=0\mu=0, then ν2(kμ)>ν2(m)=ν2(kλ)\nu_{2}(k-\mu)>\nu_{2}(m)=\nu_{2}(k-\lambda). By Theorem 54(2ciB), adjacency PST occurs between uu and vv in XrX^{r} if and only if rr is even. In this case, Xr=CP(m)r=CP(rm)X^{r}=CP(m)^{r}=CP(rm). Since rm=2r(m/2)rm=2r(m/2), we may also use Theorem 54(1) to conclude that XrX^{r} has PST for all even rr.

We also characterize when adjacency PST in preserved in self-joins using Corollary 35.

Corollary 56.

Let r,m2r,m\geq 2, kk\in\mathbb{Z} and ϕ(A(X),t)[t]\phi(A(X),t)\in\mathbb{Z}[t]. Suppose XX is simple and adjacency perfect state transfer occurs between vertices uu and vv in XX and τX=πh\tau_{X}=\frac{\pi}{h}. Then it occurs between them in XrX^{r} if and only if kmσuv(A(X))k-m\notin\sigma_{uv}^{-}(A(X)) and ν2(m)>ν2(kμ)\nu_{2}(m)>\nu_{2}(k-\mu) for all μσuv(A(X))\mu\in\sigma_{uv}^{-}(A(X)). Here, τXr=πg\tau_{X^{r}}=\frac{\pi}{g}, where g=gcd(𝒯)g=\operatorname{gcd}(\mathcal{T}), 𝒯\mathcal{T} is given in Theorem 54 and ν2(g)=ν2(h)\nu_{2}(g)=\nu_{2}(h). Further, if XX is connected, hg\frac{h}{g} is rational. Otherwise, hg=hgcd(h,m)\frac{h}{g}=\frac{h}{\operatorname{gcd}(h,m)} is odd.

10.2 Iterated join graphs

In this subsection, we characterize Laplacian strong cospectrality and PST in iterated join graphs.

An iterated join graph is a weighted graph that is either of the form:

  • (i)

    ((((Xm1Xm2)Xm3)Xm4))Xm2k:=Γ(Xm1,,Xm2k)((((X_{m_{1}}\vee X_{m_{2}})\cup X_{m_{3}})\vee X_{m_{4}})\ldots)\vee X_{m_{2k}}:=\Gamma(X_{m_{1}},\ldots,X_{m_{2k}})

  • (ii)

    ((((Xm1Xm2)Xm3)Xm4))Xm2k+1:=Γ(Xm1,,Xm2k+1)((((X_{m_{1}}\cup X_{m_{2}})\vee X_{m_{3}})\cup X_{m_{4}})\ldots)\vee X_{m_{2k+1}}:=\Gamma(X_{m_{1}},\ldots,X_{m_{2k+1}}),

where Xm1,,Xm2rX_{m_{1}},\ldots,X_{m_{2r}} are weighted graphs on m1,,m2r1m_{1},\ldots,m_{2r}\geq 1 vertices, respectively.

If we take Xj=OmjX_{j}=O_{m_{j}} for odd jj and Xj=KmjX_{j}=K_{m_{j}} otherwise, then Γ(Om1,Km2,Km2k)\Gamma(O_{m_{1}},K_{m_{2}}\ldots,K_{m_{2k}}) is a connected threshold graph. Γ(Km1,Om2,,Km2k+1)\Gamma(K_{m_{1}},O_{m_{2}},\ldots,K_{m_{2k+1}}) is also a threshold graph, where Xj=OmjX_{j}=O_{m_{j}} for even jj and Xj=KmjX_{j}=K_{m_{j}} otherwise. In fact, graphs of either form completely describe the class of connected threshold graphs [Severini]. Thus, iterated join graphs are generalizations of connected theshold graphs.

Lemma 3 and the fact that the union preserves eigenvalue supports yield the following result which determines the eigenvalue support of each vertex in an iterated join graph.

Proposition 57.

The following hold.

  1. 1.

    Let X=Γ(Xm1,,Xm2k)X=\Gamma(X_{m_{1}},\ldots,X_{m_{2k}}) and suppose uV(Xmj)u\in V(X_{m_{j}}) for some j{1,,2k}j\in\{1,\ldots,2k\}. Define αh==1hm\alpha_{h}=\sum_{\ell=1}^{h}m_{\ell}, βh==1(2kh)/2mh+2\beta_{h}=\sum_{\ell=1}^{(2k-h)/2}m_{h+2\ell} when h2kh\neq 2k and β2k=0\beta_{2k}=0. The following hold.

    1. (a)

      If jj is odd, then σu(L(X))={λ+βj1:λσu(L(Xmj))\{0}}{αh+βh:jh2kis even}\sigma_{u}(L(X))=\{\lambda+\beta_{j-1}:\lambda\in\sigma_{u}(L(X_{m_{j}}))\backslash\{0\}\}\cup\{\alpha_{h}+\beta_{h}:j\leq h\leq 2k\ \text{is even}\}\cup\mathcal{R}, where ={0}\mathcal{R}=\{0\} whenever j=1j=1 and Xm1X_{m_{1}} is connected, and ={0,βj1}\mathcal{R}=\{0,\beta_{j-1}\} otherwise.

    2. (b)

      If jj is even, then σu(L(X))={λ+αj1+βj:λσu(L(Xmj))\{0}}{αh+βh:jh2kis even}\sigma_{u}(L(X))=\{\lambda+\alpha_{j-1}+\beta_{j}:\lambda\in\sigma_{u}(L(X_{m_{j}}))\backslash\{0\}\}\cup\{\alpha_{h}+\beta_{h}:j\leq h\leq 2k\ \text{is even}\}\cup\mathcal{R}, where ={0}\mathcal{R}=\{0\} whenever XjX_{j} is connected and ={0,αj1+βj}\mathcal{R}=\{0,\alpha_{j-1}+\beta_{j}\} otherwise.

  2. 2.

    Let X=Γ(Xm1,,Xm2k+1)X=\Gamma(X_{m_{1}},\ldots,X_{m_{2k+1}}) and suppose uV(Xmj)u\in V(X_{m_{j}}) for some j{1,,2k+1}j\in\{1,\ldots,2k+1\}. Define αh==1hm\alpha_{h}=\sum_{\ell=1}^{h}m_{\ell}, βh==1(2k+1h)/2mh+2\beta_{h}=\sum_{\ell=1}^{(2k+1-h)/2}m_{h+2\ell} when h2k+1h\neq 2k+1 and β2k+1=0\beta_{2k+1}=0. The following hold.

    1. (a)

      If j=1j=1, then σu(L(X))={λ+β1:λσu(L(Xmj))\{0}}{αh+1+βh:1h2k+1is odd}\sigma_{u}(L(X))=\{\lambda+\beta_{1}:\lambda\in\sigma_{u}(L(X_{m_{j}}))\backslash\{0\}\}\cup\{\alpha_{h+1}+\beta_{h}:1\leq h\leq 2k+1\ \text{is odd}\}\cup\mathcal{R}, where ={0,β1}\mathcal{R}=\{0,\beta_{1}\}.

    2. (b)

      If j3j\geq 3 is odd, then σu(L(X))={λ+αj1+βj:λσu(L(Xmj))\{0}}{αh+βh:jh2k+1is odd}\sigma_{u}(L(X))=\{\lambda+\alpha_{j-1}+\beta_{j}:\lambda\in\sigma_{u}(L(X_{m_{j}}))\backslash\{0\}\}\cup\{\alpha_{h}+\beta_{h}:j\leq h\leq 2k+1\ \text{is odd}\}\cup\mathcal{R}, where ={0}\mathcal{R}=\{0\} if XX is connected and ={0,αj1+βj}\mathcal{R}=\{0,\alpha_{j-1}+\beta_{j}\} otherwise.

    3. (c)

      If jj is even, then σu(L(X))={λ+βj1:λσu(L(Xmj))\{0}}{αh+βh:j+1h2k+1is odd}\sigma_{u}(L(X))=\{\lambda+\beta_{j-1}:\lambda\in\sigma_{u}(L(X_{m_{j}}))\backslash\{0\}\}\cup\{\alpha_{h}+\beta_{h}:j+1\leq h\leq 2k+1\ \text{is odd}\}\cup\mathcal{R}, where ={0,βj1}\mathcal{R}=\{0,\beta_{j-1}\}.

Using Proposition 57 and the same argument in the proof of Theorem 19 yield the following characterization of Laplacing strong cospectrality in iterated join graphs.

Theorem 58.

Let X=Γ(Xm1,,Xm2k)X=\Gamma(X_{m_{1}},\ldots,X_{m_{2k}}) and consider αh\alpha_{h} and βh\beta_{h} in Proposition 57(1). Two vertices uu and vv of XmjX_{m_{j}} are Laplacian strongly cospectral in XX if and only if one of the following conditions hold.

  1. 1.

    jj is odd and either:

    1. (a)

      Vertices uu and vv are Laplacian strongly cospectral in XmjX_{m_{j}} and αjσuv(L(Xmj))\alpha_{j}\notin\sigma_{uv}^{-}(L(X_{m_{j}})). In this case, σuv+(L)={λ+βj1:λσuv+(L(Xmj))\{0}}{αh+βh:jh2kis even}\sigma_{uv}^{+}(L)=\{\lambda+\beta_{j-1}:\lambda\in\sigma_{uv}^{+}(L(X_{m_{j}}))\backslash\{0\}\}\cup\{\alpha_{h}+\beta_{h}:j\leq h\leq 2k\ \text{is even}\}\cup\mathcal{R} and σuv(L)={μ+βj1:μσuv(L(Xmj))}\sigma_{uv}^{-}(L)=\{\mu+\beta_{j-1}:\mu\in\sigma_{uv}^{-}(L(X_{m_{j}}))\}, where \mathcal{R} is given in Proposition 57(1a).

    2. (b)

      j=1j=1 and Xm1=O2X_{m_{1}}=O_{2} with σuv+(L)={αh+βh:1h2kis even}{0}\sigma_{uv}^{+}(L)=\{\alpha_{h}+\beta_{h}:1\leq h\leq 2k\ \text{is even}\}\cup\{0\} and σuv(L)={β0}\sigma_{uv}^{-}(L)=\{\beta_{0}\}.

  2. 2.

    jj is even and either:

    1. (a)

      Vertices uu and vv are Laplacian strongly cospectral in XmjX_{m_{j}} and mjσuv(L(Xmj))m_{j}\notin\sigma_{uv}^{-}(L(X_{m_{j}})). In this case, σuv+(L)={λ+αj1+βj:λσuv+(L(Xmj))\{0}}{αh+βh:jh2kis even}\sigma_{uv}^{+}(L)=\{\lambda+\alpha_{j-1}+\beta_{j}:\lambda\in\sigma_{uv}^{+}(L(X_{m_{j}}))\backslash\{0\}\}\cup\{\alpha_{h}+\beta_{h}:j\leq h\leq 2k\ \text{is even}\}\cup\mathcal{R} and σuv(L)={μ+αj1+βj:μσuv(L(Xmj))}\sigma_{uv}^{-}(L)=\{\mu+\alpha_{j-1}+\beta_{j}:\mu\in\sigma_{uv}^{-}(L(X_{m_{j}}))\}, where \mathcal{R} is given in Proposition 57(1b).

    2. (b)

      Xmj=O2X_{m_{j}}=O_{2} with σuv+(L)={αh+βh:jh2kis even}{0}\sigma_{uv}^{+}(L)=\{\alpha_{h}+\beta_{h}:j\leq h\leq 2k\ \text{is even}\}\cup\{0\} and σuv(L)={αj1+βj}\sigma_{uv}^{-}(L)=\{\alpha_{j-1}+\beta_{j}\}.

Theorem 59.

Let X=Γ(Xm1,,Xm2k+1)X=\Gamma(X_{m_{1}},\ldots,X_{m_{2k+1}}) and consider αh\alpha_{h}, βh\beta_{h} and \mathcal{R} in Proposition 57(2). Two vertices uu and vv of XmjX_{m_{j}} are Laplacian strongly cospectral in XX if and only if one of the following conditions hold.

  1. 1.

    j=1j=1, vertices uu and vv are Laplacian strongly cospectral in Xm1X_{m_{1}} and m1+m2σuv(L(Xm1))m_{1}+m_{2}\notin\sigma_{uv}^{-}(L(X_{m_{1}})). In this case, σuv+(L)={λ+β1:λσuv+(L(Xm1))\{0}}{αh+1+βh:1h2k+1is odd}\sigma_{uv}^{+}(L)=\{\lambda+\beta_{1}:\lambda\in\sigma_{uv}^{+}(L(X_{m_{1}}))\backslash\{0\}\}\cup\{\alpha_{h+1}+\beta_{h}:1\leq h\leq 2k+1\ \text{is odd}\}\cup\mathcal{R} and σuv(L)={μ+β1:μσuv(L(Xm1))}\sigma_{uv}^{-}(L)=\{\mu+\beta_{1}:\mu\in\sigma_{uv}^{-}(L(X_{m_{1}}))\}, where \mathcal{R} is given in Proposition 57(2a).

  2. 2.

    j3j\geq 3 is odd and either:

    1. (a)

      Vertices uu and vv are Laplacian strongly cospectral in XmjX_{m_{j}} and mjσuv(L(Xmj))m_{j}\notin\sigma_{uv}^{-}(L(X_{m_{j}})). In this case, σuv+(L)={λ+αj1+βj:λσuv+(L(Xmj))\{0}}{αh+βh:jh2k+1is odd}\sigma_{uv}^{+}(L)=\{\lambda+\alpha_{j-1}+\beta_{j}:\lambda\in\sigma_{uv}^{+}(L(X_{m_{j}}))\backslash\{0\}\}\cup\{\alpha_{h}+\beta_{h}:j\leq h\leq 2k+1\ \text{is odd}\}\cup\mathcal{R} and σuv(L)={μ+αj1+βj:μσuv(L(Xmj))}\sigma_{uv}^{-}(L)=\{\mu+\alpha_{j-1}+\beta_{j}:\mu\in\sigma_{uv}^{-}(L(X_{m_{j}}))\}, where \mathcal{R} is given in Proposition 57(2b).

    2. (b)

      Xmj=O2X_{m_{j}}=O_{2} with σuv+(L)={αh+βh:jh2kis odd}{0}\sigma_{uv}^{+}(L)=\{\alpha_{h}+\beta_{h}:j\leq h\leq 2k\ \text{is odd}\}\cup\{0\} and σuv(L)={αj1+βj}\sigma_{uv}^{-}(L)=\{\alpha_{j-1}+\beta_{j}\}.

  3. 3.

    jj is even, vertices uu and vv are Laplacian strongly cospectral in XmjX_{m_{j}}, and αjσuv(L(Xmj))\alpha_{j}\notin\sigma_{uv}^{-}(L(X_{m_{j}})). In this case, σuv+(L)={λ+βj1:λσuv+(L(Xmj))\{0}}{αh+βh:jh2k+1is odd}\sigma_{uv}^{+}(L)=\{\lambda+\beta_{j-1}:\lambda\in\sigma_{uv}^{+}(L(X_{m_{j}}))\backslash\{0\}\}\cup\{\alpha_{h}+\beta_{h}:j\leq h\leq 2k+1\ \text{is odd}\}\cup\mathcal{R} and σuv(L)={μ+βj1:μσuv(L(Xmj))}\sigma_{uv}^{-}(L)=\{\mu+\beta_{j-1}:\mu\in\sigma_{uv}^{-}(L(X_{m_{j}}))\}, where \mathcal{R} is given in Proposition 57(2c).

Remark 60.

From Theorems 58 and 59, a pair of strongly cospectral vertices in an iterated join graph belong to the same XmjX_{m_{j}}. Consequently, if uu and vv belong to XmjX_{m_{j}} and XmX_{m_{\ell}} respectively, then PST cannot occur between them.

Our next endeavour is to characterize PST in iterated join graphs, starting with those of the form Γ(Xm1,,Xm2k)\Gamma(X_{m_{1}},\ldots,X_{m_{2k}}). We separate the cases when jj is odd and jj is even. The following result follows from combining Corollary 7, Theorem 58, the same argument in the proof of Theorem 23, and the fact that

αh+βh=αj+γh+βj1\alpha_{h}+\beta_{h}=\alpha_{j}+\gamma_{h}+\beta_{j-1} (19)

for a fixed odd jj and for all even jh2kj\leq h\leq 2k in Theorem 58(1) with γh=0\gamma_{h}=0 whenever h=j+1h=j+1 and γh=mj+2+mj+4++mj+6+mh1\gamma_{h}=m_{j+2}+m_{j+4}++m_{j+6}\ldots+m_{h-1} otherwise, and

αh+βh=αj1+δh+βj\alpha_{h}+\beta_{h}=\alpha_{j-1}+\delta_{h}+\beta_{j} (20)

for a fixed even jj and for all even jh2kj\leq h\leq 2k in Theorem 58(2) with δh=mj+mj+1+mj+3++mh1\delta_{h}=m_{j}+m_{j+1}+m_{j+3}+\ldots+m_{h-1}.

Theorem 61.

Let X=Γ(Xm1,,Xm2k)X=\Gamma(X_{m_{1}},\ldots,X_{m_{2k}}), ϕ(L(X),t)[t]\phi(L(X),t)\in\mathbb{Z}[t] and mj2m_{j}\geq 2 for some odd jj. Consider αh\alpha_{h} and βh\beta_{h} in Proposition 57(1) and γh\gamma_{h} in (19). Two vertices uu and vv in XmjX_{m_{j}} admit Laplacian perfect state transfer in XX if and only if all of the following conditions hold.

  1. 1.

    Either (i) uu and vv are Laplacian strongly cospectral in XmjX_{m_{j}} and αjσuv(L(Xmj))\alpha_{j}\notin\sigma_{uv}^{-}(L(X_{m_{j}})) or (ii) j=1j=1 and Xm1=O2X_{m_{1}}=O_{2}.

  2. 2.

    The eigenvalues in σu(L(Xmj))\sigma_{u}(L(X_{m_{j}})) are all integers.

  3. 3.

    One of the following conditions hold.

    1. (a)

      XmjX_{m_{j}} is connected and one of the following conditions hold for all λσuv+(L(Xmj))\{0}\lambda\in\sigma_{uv}^{+}(L(X_{m_{j}}))\backslash\{0\}, μ,θσuv(L(Xmj))\mu,\theta\in\sigma_{uv}^{-}(L(X_{m_{j}})) and even jh2kj\leq h\leq 2k.

      1. i.

        ν2(βj1)>ν2(μ)=ν2(θ)\nu_{2}(\beta_{j-1})>\nu_{2}(\mu)=\nu_{2}(\theta), ν2(λ)>ν2(μ)\nu_{2}(\lambda)>\nu_{2}(\mu) and ν2(αj+γh)>ν2(μ)\nu_{2}(\alpha_{j}+\gamma_{h})>\nu_{2}(\mu).

      2. ii.

        ν2(μ)>ν2(βj1)=ν2(λ)=ν2(αj+γh)\nu_{2}(\mu)>\nu_{2}(\beta_{j-1})=\nu_{2}(\lambda)=\nu_{2}(\alpha_{j}+\gamma_{h}).

      3. iii.

        ν2(βj1)=ν2(μ)=ν2(λ)=ν2(αj+γh)\nu_{2}(\beta_{j-1})=\nu_{2}(\mu)=\nu_{2}(\lambda)=\nu_{2}(\alpha_{j}+\gamma_{h}), ν2(λ+βj12ν2(βj1))>ν2(μ+βj12ν2(βj1))=ν2(θ+βj12ν2(βj1))\nu_{2}\left(\frac{\lambda+\beta_{j-1}}{2^{\nu_{2}(\beta_{j-1})}}\right)>\nu_{2}\left(\frac{\mu+\beta_{j-1}}{2^{\nu_{2}(\beta_{j-1})}}\right)=\nu_{2}\left(\frac{\theta+\beta_{j-1}}{2^{\nu_{2}(\beta_{j-1})}}\right) and ν2(αh+βh2ν2(βj1))>ν2(μ+βj12ν2(βj1)).\nu_{2}\left(\frac{\alpha_{h}+\beta_{h}}{2^{\nu_{2}(\beta_{j-1})}}\right)>\nu_{2}\left(\frac{\mu+\beta_{j-1}}{2^{\nu_{2}(\beta_{j-1})}}\right).

    2. (b)

      XmjX_{m_{j}} is disconnected and (ai) above holds.

    3. (c)

      j=1j=1, Xm1=O2X_{m_{1}}=O_{2} and ν2(2+γh+β0)>ν2(β0)\nu_{2}(2+\gamma_{h}+\beta_{0})>\nu_{2}(\beta_{0}) for each even 1h2k1\leq h\leq 2k.

Further, τXY=πg\tau_{X\vee Y}=\frac{\pi}{g}, where g=gcd(𝒯)g=\operatorname{gcd}\left(\mathcal{T}\right), where 𝒯=σu(L(X))\mathcal{T}=\sigma_{u}(L(X)) is given in Proposition 57(1).

Theorem 62.

Let X=Γ(Xm1,,Xm2k)X=\Gamma(X_{m_{1}},\ldots,X_{m_{2k}}), ϕ(L(X),t)[t]\phi(L(X),t)\in\mathbb{Z}[t] and mj2m_{j}\geq 2 for some even jj. Consider αh\alpha_{h} and βh\beta_{h} in Proposition 57(1) and δh\delta_{h} in (20) Two vertices uu and vv in XmjX_{m_{j}} admit Laplacian perfect state transfer in XX if and only if all of the following conditions hold.

  1. 1.

    Either (i) uu and vv are Laplacian strongly cospectral in XmjX_{m_{j}} and mjσuv(L(Xmj))m_{j}\notin\sigma_{uv}^{-}(L(X_{m_{j}})) or (ii) Xmj=O2X_{m_{j}}=O_{2}.

  2. 2.

    The eigenvalues in σu(L(Xmj))\sigma_{u}(L(X_{m_{j}})) are all integers.

  3. 3.

    One of the following conditions hold.

    1. (a)

      XmjX_{m_{j}} is connected and one of the following conditions hold for all λσuv+(L(Xmj))\{0}\lambda\in\sigma_{uv}^{+}(L(X_{m_{j}}))\backslash\{0\}, μ,θσuv(L(Xmj))\mu,\theta\in\sigma_{uv}^{-}(L(X_{m_{j}})) and even jh2kj\leq h\leq 2k.

      1. i.

        ν2(αj1+βj)>ν2(μ)=ν2(θ)\nu_{2}(\alpha_{j-1}+\beta_{j})>\nu_{2}(\mu)=\nu_{2}(\theta), ν2(λ)>ν2(μ)\nu_{2}(\lambda)>\nu_{2}(\mu) and ν2(δh)>ν2(μ)\nu_{2}(\delta_{h})>\nu_{2}(\mu).

      2. ii.

        ν2(μ)>ν2(αj1+βj)=ν2(λ)=ν2(δh)\nu_{2}(\mu)>\nu_{2}(\alpha_{j-1}+\beta_{j})=\nu_{2}(\lambda)=\nu_{2}(\delta_{h}).

      3. iii.

        ν2(αj1+βj)=ν2(μ)=ν2(λ)=ν2(δh)\nu_{2}(\alpha_{j-1}+\beta_{j})=\nu_{2}(\mu)=\nu_{2}(\lambda)=\nu_{2}(\delta_{h}), ν2(λ+αj1+βj2ν2(αj1+βj))>ν2(μ+αj1+βj2ν2(αj1+βj))=ν2(θ+αj1+βj2ν2(αj1+βj))\nu_{2}\left(\frac{\lambda+\alpha_{j-1}+\beta_{j}}{2^{\nu_{2}(\alpha_{j-1}+\beta_{j})}}\right)>\nu_{2}\left(\frac{\mu+\alpha_{j-1}+\beta_{j}}{2^{\nu_{2}(\alpha_{j-1}+\beta_{j})}}\right)=\nu_{2}\left(\frac{\theta+\alpha_{j-1}+\beta_{j}}{2^{\nu_{2}(\alpha_{j-1}+\beta_{j})}}\right) and ν2(αh+βh2ν2(αj1+βj))>ν2(μ+αj1+βj2ν2(αj1+βj)).\nu_{2}\left(\frac{\alpha_{h}+\beta_{h}}{2^{\nu_{2}(\alpha_{j-1}+\beta_{j})}}\right)>\nu_{2}\left(\frac{\mu+\alpha_{j-1}+\beta_{j}}{2^{\nu_{2}(\alpha_{j-1}+\beta_{j})}}\right).

    2. (b)

      XmjX_{m_{j}} is disconnected and (ai) above holds.

    3. (c)

      Xmj=O2X_{m_{j}}=O_{2} and ν2(αj1+δh+βj)>ν2(αj1+βj)\nu_{2}(\alpha_{j-1}+\delta_{h}+\beta_{j})>\nu_{2}(\alpha_{j-1}+\beta_{j}) for each even jh2kj\leq h\leq 2k.

Further, τXY=πg\tau_{X\vee Y}=\frac{\pi}{g}, where g=gcd(𝒯)g=\operatorname{gcd}\left(\mathcal{T}\right), where 𝒯=σu(L(X))\mathcal{T}=\sigma_{u}(L(X)) is given in Proposition 57(1).

Remark 63.

We have the following.

  1. 1.

    We obtain a characterization for PST between vertices in XmjX_{m_{j}} in the graph X=Γ(Xm1,,Xm2k+1)X=\Gamma(X_{m_{1}},\ldots,X_{m_{2k+1}}) whenever mj2m_{j}\geq 2 for some even j2j\geq 2 by replacing ‘even jh2kj\leq h\leq 2k’ with ‘odd jh2k+1j\leq h\leq 2k+1’ in Theorem 61.

  2. 2.

    We obtain a characterization for PST between vertices in XmjX_{m_{j}} in the graph X=Γ(Xm1,,Xm2k+1)X=\Gamma(X_{m_{1}},\ldots,X_{m_{2k+1}}) whenever mj2m_{j}\geq 2 for some odd j3j\geq 3 by replacing ‘even jh2kj\leq h\leq 2k’ with ‘odd jh2k+1j\leq h\leq 2k+1’ in Theorem 62.

In order to complete our characterization of PST in iterated join graphs, it suffices to determine PST between vertices in Xm1X_{m_{1}} in the graph X=Γ(Xm1,,Xm2k+1)X=\Gamma(X_{m_{1}},\ldots,X_{m_{2k+1}}) by virtue of Remark 63. To do this, we combine Corollary 7, Theorem 58, the same argument in the proof of Theorem 23, and the fact that

αh+1+βh=ϕh+β1\alpha_{h+1}+\beta_{h}=\phi_{h}+\beta_{1} (21)

for all odd 1h2k+11\leq h\leq 2k+1 in Theorem 59(1) for a fixed odd jj with ϕh=m1+m2+m4+m6++mh1\phi_{h}=m_{1}+m_{2}+m_{4}+m_{6}+\ldots+m_{h-1}.

Theorem 64.

Let X=Γ(Xm1,,Xm2k+1)X=\Gamma(X_{m_{1}},\ldots,X_{m_{2k+1}}), m12m_{1}\geq 2 and ϕ(L(X),t)[t]\phi(L(X),t)\in\mathbb{Z}[t]. Consider αh\alpha_{h} and βh\beta_{h} in Proposition 57(2) and ϕh\phi_{h} in (21). Two vertices uu and vv in Xm1X_{m_{1}} admit Laplacian perfect state transfer in XX if and only if all of the following conditions hold.

  1. 1.

    Vertices uu and vv are Laplacian strongly cospectral in Xm1X_{m_{1}} and m1+m2σuv(L(Xm1))m_{1}+m_{2}\notin\sigma_{uv}^{-}(L(X_{m_{1}})).

  2. 2.

    The eigenvalues in σu(L(Xm1))\sigma_{u}(L(X_{m_{1}})) are all integers.

  3. 3.

    One of the following conditions hold.

    1. (a)

      Xm1X_{m_{1}} is connected and one of the following conditions hold for all λσuv+(L(Xm1))\{0}\lambda\in\sigma_{uv}^{+}(L(X_{m_{1}}))\backslash\{0\}, μ,θσuv(L(Xm1))\mu,\theta\in\sigma_{uv}^{-}(L(X_{m_{1}})) and odd 1h2k+11\leq h\leq 2k+1.

      1. i.

        ν2(β1)>ν2(μ)=ν2(θ)\nu_{2}(\beta_{1})>\nu_{2}(\mu)=\nu_{2}(\theta), ν2(λ)>ν2(μ)\nu_{2}(\lambda)>\nu_{2}(\mu) and ν2(ϕh)>ν2(μ)\nu_{2}(\phi_{h})>\nu_{2}(\mu).

      2. ii.

        ν2(μ)>ν2(β1)=ν2(λ)=ν2(ϕh)\nu_{2}(\mu)>\nu_{2}(\beta_{1})=\nu_{2}(\lambda)=\nu_{2}(\phi_{h}).

      3. iii.

        ν2(β1)=ν2(μ)=ν2(λ)=ν2(ϕh)\nu_{2}(\beta_{1})=\nu_{2}(\mu)=\nu_{2}(\lambda)=\nu_{2}(\phi_{h}), ν2(λ+β12ν2(β1))>ν2(μ+β12ν2(β1))=ν2(θ+β12ν2(β1))\nu_{2}\left(\frac{\lambda+\beta_{1}}{2^{\nu_{2}(\beta_{1})}}\right)>\nu_{2}\left(\frac{\mu+\beta_{1}}{2^{\nu_{2}(\beta_{1})}}\right)=\nu_{2}\left(\frac{\theta+\beta_{1}}{2^{\nu_{2}(\beta_{1})}}\right) and ν2(αh+βh2ν2(β1))>ν2(μ+β12ν2(β1)).\nu_{2}\left(\frac{\alpha_{h}+\beta_{h}}{2^{\nu_{2}(\beta_{1})}}\right)>\nu_{2}\left(\frac{\mu+\beta_{1}}{2^{\nu_{2}(\beta_{1})}}\right).

    2. (b)

      Xm1X_{m_{1}} is disconnected and (ai) above holds.

    In both cases, uu and vv belong to the same component in Xm1X_{m_{1}}.

Further, τXY=πg\tau_{X\vee Y}=\frac{\pi}{g}, where g=gcd(𝒯)g=\operatorname{gcd}\left(\mathcal{T}\right), where 𝒯=σu(L(X))\mathcal{T}=\sigma_{u}(L(X)) is given in Proposition 57(2).

Remark 65.

If XmjK2X_{m_{j}}\neq K_{2} is unweighted, then we may drop the condition m1+m2σuv(L(Xm1))m_{1}+m_{2}\in\sigma_{uv}^{-}(L(X_{m_{1}})) in Theorem 59(1), the condition αjσuv(L(Xmj))\alpha_{j}\notin\sigma_{uv}^{-}(L(X_{m_{j}})) in Theorems 58(1a) and 59(3), and the condition mjσuv(L(Xmj))m_{j}\notin\sigma_{uv}^{-}(L(X_{m_{j}})) in Theorems 58(2a) and 59(2a). The same applies to Theorems 61, 62 and 64.

Remark 66.

The conditions in Theorem 61(3c) are equivalent to m1=2m_{1}=2 and

ν2(2+β0)>ν2(β0),ν2(2+m3+β0)>ν2(β0),ν2(2+m3+m5++m2k1+β0)>ν2(β0).\nu_{2}(2+\beta_{0})>\nu_{2}(\beta_{0}),\quad\nu_{2}(2+m_{3}+\beta_{0})>\nu_{2}(\beta_{0}),\quad\ldots\quad\nu_{2}(2+m_{3}+m_{5}+\ldots+m_{2k-1}+\beta_{0})>\nu_{2}(\beta_{0}).

The first inequality above is equivalent to β02\beta_{0}\equiv 2 (mod 4), and the latter ones yield mj0m_{j}\equiv 0 (mod 4) for all odd j3j\geq 3. Since X=Z1Xm2kX=Z_{1}\vee X_{m_{2k}} for some disconnected graph Z1Z_{1}, applying Theorem 23(3ab) yields ν2(m2k)>ν2(β0)\nu_{2}(m_{2k})>\nu_{2}(\beta_{0}) and ν2(m1+m2+m4++m2k2)>ν2(β0)\nu_{2}(m_{1}+m_{2}+m_{4}+\ldots+m_{2k-2})>\nu_{2}(\beta_{0}). Now, Corollary 30 implies that PST also occurs between uu and vv in Z2Z_{2}, where Z2Z_{2} is the graph such that Z1=Z2Xm2k1Z_{1}=Z_{2}\cup X_{m_{2k-1}}. The same argument applied k2k-2 times yields ν2(mj)>ν2(β0)\nu_{2}(m_{j})>\nu_{2}(\beta_{0}) for all even j3j\geq 3 and PST between uu and vv in Zmk2=O2Xm2Z_{m_{k-2}}=O_{2}\vee X_{m_{2}}. As β02\beta_{0}\equiv 2 (mod 4), the former condition is equivalent to mj0m_{j}\equiv 0 (mod 4) for all even j3j\geq 3, while the latter is equivalent to m22m_{2}\equiv 2 (mod 4) by virtue of Theorem 23(3c). These considerations imply that Theorem 61(3c) holds if and only if m1=2m_{1}=2, m22m_{2}\equiv 2 (mod 4) and mj0m_{j}\equiv 0 (mod 4) for all j3j\geq 3. One can also check that Theorem 64 holds for Xm1=K2X_{m_{1}}=K_{2} if and only if m1=2m_{1}=2, m22m_{2}\equiv 2 (mod 4) and mj0m_{j}\equiv 0 (mod 4) for all j3j\geq 3. In both cases, the minimum PST time is π2\frac{\pi}{2}.

Corollary 67.

Let XX be a connected threshold graph. Then Laplacian perfect state transfer occurs in XX if and only if m1=2m_{1}=2, m22m_{2}\equiv 2 (mod 4) and mj2m_{j}\equiv 2 (mod 4) for all j3j\geq 3. In this case, Laplacian perfect state transfer occurs between vertices uu and vv in Xm1{O2,K2}X_{m_{1}}\in\{O_{2},K_{2}\} with minimum PST time is π2\frac{\pi}{2}.

Proof.

By assumption, X{Γ(Om1,Km1,,Km2k),Γ(Km1,Om1,,Km2k+1)}X\in\{\Gamma(O_{m_{1}},K_{m_{1}},\ldots,K_{m_{2k}}),\Gamma(K_{m_{1}},O_{m_{1}},\ldots,K_{m_{2k+1}})\}. First, suppose X=Γ(Om1,Km1,,Km2k)X=\Gamma(O_{m_{1}},K_{m_{1}},\ldots,K_{m_{2k}}). For two vertices uu and vv to be strongly cospectral in XX, Theorem 19(3) implies that they both belong to OmjO_{m_{j}} if jj is odd or KmjK_{m_{j}} if jj is even. Hence, we have three cases.

  • Let j=1j=1. Then vertices uu and vv of Om1O_{m_{1}} are strongly cospectral in XX if and only if m1=2m_{1}=2.

  • Suppose j3j\geq 3 is odd so that vertices uu and vv both belong to OmjO_{m_{j}}. Then we can write X=(((YOmj)Kmj+1))Km2kX=(((Y\cup O_{m_{j}})\vee K_{m_{j+1}})\cup\ldots)\vee K_{m_{2k}} for some graph YY. Now, Theorem 19(1) and an inductive argument implies that either (i) uu and vv are strongly cospectral in (YXmj)Xmj+1(Y\cup X_{m_{j}})\vee X_{m_{j+1}} and mjσuv(L(Xj))m_{j}\notin\sigma_{uv}^{-}(L(X_{j})) or (ii) YXmj=O2Y\cup X_{m_{j}}=O_{2}. Since YY has at least one vertex, the latter case does not hold. Hence, it must be that uu and vv are strongly cospectral in (YOmj)Kmj+1(Y\cup O_{m_{j}})\vee K_{m_{j+1}}. But this implies that uu and vv are strongly cospectral in OmjO_{m_{j}}, which cannot happen since by Theorem 19, the only graph with isolated vertices that admit strong cospectrality in the join is O2O_{2}, which is not equal to YOmjY\cup O_{m_{j}}.

  • Suppose j3j\geq 3 is even so that uu and vv belong to KmjK_{m_{j}}. Then the same argument as the previous case implies that uu and vv are strongly cospectral in YKmjY\vee K_{m_{j}} and mjσuv(L(Xj))m_{j}\notin\sigma_{uv}^{-}(L(X_{j})). But this implies that uu and vv are strongly cospectral in KmjK_{m_{j}}, i.e., mj=2σuv(L(Xj))m_{j}=2\in\sigma_{uv}^{-}(L(X_{j})), which is a contradiction.

Thus, in X=Γ(Om1,Km1,,Km2k)X=\Gamma(O_{m_{1}},K_{m_{1}},\ldots,K_{m_{2k}}), only the vertices of Om1O_{m_{1}} admit strong cospectrality in XX, in which case m1=2m_{1}=2. The same holds for X=Γ(Om1,Km1,,Km2k)X=\Gamma(O_{m_{1}},K_{m_{1}},\ldots,K_{m_{2k}}). Cmbining this with Remark 66 completes the proof. ∎

11 Bounds

From the previous sections, PST between uu and ww in XX need not guarantee PST between uu and vv in XYX\vee Y. The same can be said about periodicity of a vertex. Thus, we ask, if u,vV(X)u,v\in V(X) then how far can |UM(XY,t)u,v||U_{M}(X\vee Y,t)_{u,v}| be from |UM(X,t)u,v||U_{M}(X,t)_{u,v}| as tt ranges over \mathbb{R}? We answer this question by providing an upper bound for the absolute value of |UM(XY,t)u,v||UM(X,t)u,v||U_{M}(X\vee Y,t)_{u,v}|-|U_{M}(X,t)_{u,v}|. To do this, we define

αM(t)={meitn+neitm(m+n)m(m+n),if M=Leitλ+(kλ)mDeitλ(kλ+)mDeitkm,if M=A.\alpha_{M}(t)=\begin{cases}\dfrac{me^{-itn}+ne^{itm}-(m+n)}{m(m+n)},&\text{if }M=L\\ \dfrac{e^{it\lambda^{+}}(k-\lambda^{-})}{m\sqrt{D}}-\dfrac{e^{it\lambda^{-}}(k-\lambda^{+})}{m\sqrt{D}}-\dfrac{e^{itk}}{m},&\text{if }M=A.\end{cases} (22)

and

TM={{2jπ/g:j},if M=L{t:eitk=eitλ+=eitλ},if M=A,T_{M}=\begin{cases}\{2j\pi/g:j\in\mathbb{Z}\},&\text{if }M=L\\ \{t\in\mathbb{R}:e^{itk}=e^{it\lambda^{+}}=e^{it\lambda^{-}}\},&\text{if }M=A,\end{cases} (23)

where g=gcd(m,n)g=\operatorname{gcd}(m,n) and λ±\lambda^{\pm} are given in Lemma 2. If the context is clear, then we simply write as αM(t)\alpha_{M}(t) and TMT_{M} as α(t)\alpha(t) and TT, respectively. If we add that k,k,\ell\in\mathbb{Z} and Δ\Delta in Lemma 2 is a perfect square, then we may write TA={2jπ/h:j}T_{A}=\{2j\pi/h:j\in\mathbb{Z}\}, where h=gcd(λ+k,λk)h=\operatorname{gcd}(\lambda^{+}-k,\lambda^{-}-k).

Combining (4) and (6) with (22) yields the following lemma.

Lemma 68.

We have αM(t)=0\alpha_{M}(t)=0 if and only if tTMt\in T_{M}. Moreover, for any u,vV(X)u,v\in V(X) and for all tt,

UL(XY,t)u,veitnUL(X,t)u,v=eitnαL(t)andUA(XY,t)u,vUA(X,t)u,v=αA(t).U_{L}(X\vee Y,t)_{u,v}-e^{itn}U_{L}(X,t)_{u,v}=e^{itn}\alpha_{L}(t)\quad\text{and}\quad U_{A}(X\vee Y,t)_{u,v}-U_{A}(X,t)_{u,v}=\alpha_{A}(t).

11.1 Laplacian case

Theorem 69.

For all u,vV(X)u,v\in V(X) and for all tt

|UL(XY,t)u,veitnUL(X,t)u,v|2/m\big{|}\ U_{L}(X\vee Y,t)_{u,v}-e^{itn}U_{L}(X,t)_{u,v}\ \big{|}\leq 2/m (24)

with equality if and only if ν2(m)=ν2(n)\nu_{2}(m)=\nu_{2}(n), in which case equality holds in (24) at time τ=jπ/g\tau=j\pi/g, where jj is any odd integer and g=gcd(m,n)g=\operatorname{gcd}(m,n). Moreover, if equality holds in (24), then

|UL(XY,τ)u,veiτnUL(X,τ)u,v|=UL(XY,τ)u,v+UL(X,τ)u,v=2/m.|\ U_{L}(X\vee Y,\tau)_{u,v}-e^{i\tau n}U_{L}(X,\tau)_{u,v}\ |=U_{L}(X\vee Y,\tau)_{u,v}+U_{L}(X,\tau)_{u,v}=2/m. (25)
Proof.

Lemma 68 yields (24) with equality if and only if eiτm=eiτn=1e^{i\tau m}=e^{i\tau n}=-1 for some τ>0\tau>0. Equivalently, ν2(m)=ν2(n)\nu_{2}(m)=\nu_{2}(n), in which case τ=jπ/g\tau=j\pi/g for some odd jj. From this, (25) is immediate. ∎

Corollary 70.

For all u,vV(X)u,v\in V(X) and for all tt,

||UL(XY,t)u,v||UL(X,t)u,v||2/m\bigg{|}\ |U_{L}(X\vee Y,t)_{u,v}|-|U_{L}(X,t)_{u,v}|\ \bigg{|}\leq 2/m (26)

with equality if and only if ν2(m)=ν2(n)\nu_{2}(m)=\nu_{2}(n), UL(X,τ)u,vU_{L}(X,\tau)_{u,v}\in\mathbb{R} and either UL(X,τ)u,v0U_{L}(X,\tau)_{u,v}\leq 0 or UL(X,τ)u,v2/mU_{L}(X,\tau)_{u,v}\geq 2/m, in which case equality holds in (26) at time τ\tau in Theorem 69.

Proof.

Applying the triangle inequality to (24) yields (26). Using (25), we get equality in (26) if and only if ||UL(XY,τ)u,v||UL(X,τ)u,v||=||2mUL(X,τ)u,v||UL(X,τ)u,v||=2m\big{|}\ |U_{L}(X\vee Y,\tau)_{u,v}|-|U_{L}(X,\tau)_{u,v}|\ \big{|}=\big{|}\ \big{|}\frac{2}{m}-U_{L}(X,\tau)_{u,v}\big{|}-|U_{L}(X,\tau)_{u,v}|\ \big{|}=\frac{2}{m}. Result follows. ∎

11.2 Adjacency case

The following is an analogue of Theorem 69 for the adjacency case.

Theorem 71.

For all u,vV(X)u,v\in V(X) and for all tt,

|UA(XY,t)u,vUA(X,t)u,v|2/m\big{|}\ U_{A}(X\vee Y,t)_{u,v}-U_{A}(X,t)_{u,v}\ \big{|}\leq 2/m (27)

with equality if and only if there is a time τ>0\tau>0 such that eiτλ+=eiτλ=eiτke^{i\tau\lambda^{+}}=e^{i\tau\lambda^{-}}=-e^{i\tau k}, in which case αA(t)=2eiτk/m\alpha_{A}(t)=-2e^{i\tau k}/m. If we add that k,k,\ell\in\mathbb{Z} and Δ\Delta in Lemma 2 is a perfect square, then the latter condition yields {k,λ±}\{k,\lambda^{\pm}\}\subseteq\mathbb{Z} and ν2(λ+k)=ν2(λk)\nu_{2}(\lambda^{+}-k)=\nu_{2}(\lambda^{-}-k), in which case τ=jπ/h\tau=j\pi/h, where h=gcd(λ+k,λk)h=\operatorname{gcd}(\lambda^{+}-k,\lambda^{-}-k) and jj is any odd.

Proof.

Since λ±=12(k+±D)\lambda^{\pm}=\frac{1}{2}(k+\ell\pm\sqrt{D}), one checks that

eitλ+(kλ)mDeitλ(kλ+)mD=eitλ+(k+D)2mDeitλ(kD)2mD=1mD[Dcos(tD/2)+i(k)sin(tD/2)].\begin{split}\frac{e^{it\lambda^{+}}(k-\lambda^{-})}{m\sqrt{D}}-\frac{e^{it\lambda^{-}}(k-\lambda^{+})}{m\sqrt{D}}&=\frac{e^{it\lambda^{+}}(k-\ell+\sqrt{D})}{2m\sqrt{D}}-\frac{e^{it\lambda^{-}}(k-\ell-\sqrt{D})}{2m\sqrt{D}}\\ &=\frac{1}{m\sqrt{D}}\left[\sqrt{D}\cos(t\sqrt{D}/2)+i(k-\ell)\sin(t\sqrt{D}/2)\right].\end{split} (28)

Consequently, the following equation holds for all tt

|eitλ+(kλ)mDeitλ(kλ+)mD|=1mDD+((k)2D)sin2(tD/2).\begin{split}\left|\frac{e^{it\lambda^{+}}(k-\lambda^{-})}{m\sqrt{D}}-\frac{e^{it\lambda^{-}}(k-\lambda^{+})}{m\sqrt{D}}\right|&=\frac{1}{m\sqrt{D}}\sqrt{D+((k-\ell)^{2}-D)\sin^{2}(t\sqrt{D}/2)}.\end{split} (29)

Since mn>0mn>0 and D=(k)2+4mnD=(k-\ell)^{2}+4mn, we get that (k)2D=4mn<0(k-\ell)^{2}-D=-4mn<0. Hence, (29) gives us |eitλ+(kλ)mDeitλ(kλ+)mD|1m\left|\frac{e^{it\lambda^{+}}(k-\lambda^{-})}{m\sqrt{D}}-\frac{e^{it\lambda^{-}}(k-\lambda^{+})}{m\sqrt{D}}\right|\leq\frac{1}{m}. Combining this with (6) yields

|UA(XY,t)u,vUA(X,t)u,v||eitλ+(kλ)mDeitλ(kλ+)mD|+1m2m\big{|}U_{A}(X\vee Y,t)_{u,v}-U_{A}(X,t)_{u,v}\big{|}\leq\left|\frac{e^{it\lambda^{+}}(k-\lambda^{-})}{m\sqrt{D}}-\frac{e^{it\lambda^{-}}(k-\lambda^{+})}{m\sqrt{D}}\right|+\frac{1}{m}\leq\frac{2}{m}

which establishes (27). Using the first equality in (28), we can write (6) as

UA(XY,t)u,vUA(X,t)u,v=eitλ+(k+D)2mDeitλ(kD)2mDeitkmU_{A}(X\vee Y,t)_{u,v}-U_{A}(X,t)_{u,v}=\frac{e^{it\lambda^{+}}(k-\ell+\sqrt{D})}{2m\sqrt{D}}-\frac{e^{it\lambda^{-}}(k-\ell-\sqrt{D})}{2m\sqrt{D}}-\frac{e^{itk}}{m}.

Thus, equality holds in (27) if and only if there is a time τ\tau such that

eiτλ+=eiτλ=eiτk,e^{i\tau\lambda^{+}}=e^{i\tau\lambda^{-}}=-e^{i\tau k}, (30)

in which case, UA(XY,t)u,vUA(X,t)u,v=2eitkmU_{A}(X\vee Y,t)_{u,v}-U_{A}(X,t)_{u,v}=-\frac{2e^{itk}}{m}. This implies that {k,λ±}\{k,\lambda^{\pm}\} satisfies the ratio condition. Thus, if k,k,\ell\in\mathbb{Z} and DD is a perfect square, then Lemma 6 yields {k,λ±}\{k,\lambda^{\pm}\}\subseteq\mathbb{Z}. Hence, (30) holds if and only if ν2(λ+k)=ν2(λk)\nu_{2}(\lambda^{+}-k)=\nu_{2}(\lambda^{-}-k), in which case τ=jπ/g\tau=j\pi/g for any odd integer jj. ∎

Corollary 72.

For all u,vV(X)u,v\in V(X) and for all tt,

||UA(XY,t)u,v||UA(X,t)u,v||2/m\begin{split}\bigg{|}\ |U_{A}(X\vee Y,t)_{u,v}|-|U_{A}(X,t)_{u,v}|\ \bigg{|}&\leq 2/m\end{split} (31)

with equality if and only if the bound in (27) holds at time τ\tau in Theorem 71, UA(X,τ)u,v=|UA(X,τ)u,v|eiτkU_{A}(X,\tau)_{u,v}=|U_{A}(X,\tau)_{u,v}|e^{i\tau k} and |UA(X,τ)u,v|2/m|U_{A}(X,\tau)_{u,v}|\geq 2/m.

Proof.

Again applying triangle inequality to (27) yields (31). Making use of Theorem 71, we get equality in (31) if and only if ||UA(X,τ)u,v2eiτk/m||UA(X,τ)u,v||=2/m\big{|}\ |U_{A}(X,\tau)_{u,v}-2e^{i\tau k}/m\ |-|U_{A}(X,\tau)_{u,v}|\ \big{|}=2/m. Thus, our result is immediate. ∎

Combining Corollary 70 and 72 yields the following result.

Corollary 73.

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

||UM(XY,t)u,v||UM(X,t)u,v||0as m.\bigg{|}\ |U_{M}(X\vee Y,t)_{u,v}|-|U_{M}(X,t)_{u,v}|\ \bigg{|}\rightarrow 0\quad\text{as $m\rightarrow\infty$}.

From Corollary 73, we find that for vertices in XX, the state transfer properties of the quantum walk on XYX\vee Y tend to mimic those of the quantum walk on X,X, as we increase the number of vertices of XX.

11.3 Tightness

Define F(t)u,v:=|UM(XY,t)u,v||UM(X,t)u,v|F(t)_{u,v}:=|U_{M}(X\vee Y,t)_{u,v}|-|U_{M}(X,t)_{u,v}|. From Lemma 68, F(t)u,v=|UM(X,t)u,v+αM(t)||UM(X,t)u,v|F(t)_{u,v}=\big{|}U_{M}(X,t)_{u,v}+\alpha_{M}(t)\big{|}-|U_{M}(X,t)_{u,v}|. for any u,vV(X)u,v\in V(X). We now illustrate that the values of F(t)u,vF(t)_{u,v} can be zero, positive or negative.

Lemma 74.

The following are immediate from (23).

  1. 1.

    If τTM\tau\in T_{M} then αM(τ)=0\alpha_{M}(\tau)=0, and so F(τ)u,v=0F(\tau)_{u,v}=0.

  2. 2.

    If UM(X,τ)u,v=0U_{M}(X,\tau)_{u,v}=0 for some τ\TM\tau\in\mathbb{R}\backslash T_{M}, then F(τ)u,v=|αM(τ)|>0F(\tau)_{u,v}=|\alpha_{M}(\tau)|>0. If XX is disconnected and uu and vv are in different connected components, then UM(X,t)u,v=0U_{M}(X,t)_{u,v}=0 for all tt\in\mathbb{R}, and so F(t)u,v>0F(t)_{u,v}>0 for all t\TMt\in\mathbb{R}\backslash T_{M}.

  3. 3.

    Let |UM(X,τ)u,v|=1|U_{M}(X,\tau)_{u,v}|=1 for some τ\TM\tau\in\mathbb{R}\backslash T_{M}. Since αM(t)0\alpha_{M}(t)\neq 0 and |UM(XY,t)u,v|1|U_{M}(X\vee Y,t)_{u,v}|\leq 1, we get F(τ)u,v<0F(\tau)_{u,v}<0. If uu is isolated in XX, then UM(X,t)u,u=1U_{M}(X,t)_{u,u}=1 for all tt, and so F(t)u,u<0F(t)_{u,u}<0 for all t\TMt\in\mathbb{R}\backslash T_{M}.

We now use Lemma 74 to show that under mild conditions on mm and nn, if XX has PST between uu and vv, then there are choices of tt for which |F(t)u,v|=2/m|F(t)_{u,v}|=2/m. Thus, the bounds in Corollaries 70 and 72 are tight.

Example 75.

Suppose XX exhibits Laplacian PST between uu and vv at π/2\pi/2. Let τ=rπ\tau=r\pi for any odd rr so that UL(X,τ/2)u,v=1U_{L}(X,\tau/2)_{u,v}=1 and UL(X,τ/2)u,u=0U_{L}(X,\tau/2)_{u,u}=0. If ν2(m)=ν2(n)=1\nu_{2}(m)=\nu_{2}(n)=1, then we can write TL={jπ/g:j}T_{L}=\{j\pi/g^{\prime}:j\in\mathbb{Z}\}, where g=g/2g^{\prime}=g/2 is odd. Hence, τ/2\TL\tau/2\in\mathbb{R}\backslash T_{L}. In this case, αL(τ/2)=2m\alpha_{L}(\tau/2)=-\frac{2}{m}, and so F(τ/2)u,u=F(τ/2)u,v=2mF(\tau/2)_{u,u}=-F(\tau/2)_{u,v}=\frac{2}{m}.

Example 76.

Suppose XX admits adjacency PST between uu and vv at π/2\pi/2. Let k,k,\ell\in\mathbb{Z} and Δ\Delta be a perfect square so that, k,λ+k,\lambda^{+} and λ\lambda^{-} are integers. Let τ=rπ\tau=r\pi for any odd rr so that UA(X,τ/2)u,v=eiτk/2U_{A}(X,\tau/2)_{u,v}=e^{i\tau k/2} and UA(X,τ/2)u,u=0U_{A}(X,\tau/2)_{u,u}=0. If ν2(λ+k)=1\nu_{2}(\lambda^{+}-k)=1 and ν2(λk)=1\nu_{2}(\lambda^{-}-k)=1, then we can write TA={jπ/h:j}T_{A}=\{j\pi/h^{\prime}:j\in\mathbb{Z}\} where h=h/2h^{\prime}=h/2 is odd. Thus, τ/2\TA\tau/2\in\mathbb{R}\backslash T_{A}, and so αA(τ/2)=2eiτk/2m\alpha_{A}(\tau/2)=-\frac{2e^{i\tau k/2}}{m}, which yields F(τ/2)u,u=F(τ/2)u,v=2mF(\tau/2)_{u,u}=-F(\tau/2)_{u,v}=\frac{2}{m}.

One checks that disconnected double cones with PST at π2\frac{\pi}{2} satisfy the conditions in Examples 75 and 76, and so they provide infinite families of graphs whereby the upper bound in Corollaries 70 and 72 is met.

12 Future work

In this paper, we investigated strong cospectrality, periodicity and perfect state transfer in join graphs, and characterized the conditions in which these properties are preserved or induced in a join. In order to inspire further systematic study of quantum walks on join graphs, we present some open questions.

If XX is a weighted join graph on an odd number of vertices and ϕ(L(X),t)[t]\phi(L(X),t)\in\mathbb{Z}[t], then Corollary 25 implies that Laplacian PST does not occur in XX. This is also known to be true for unweighted graphs on n{3,5}n\in\{3,5\} vertices, which are not necessarily joins. This leads us to conjecture the following.

Conjecture 77.

If XX is a simple connected unweighted graph on an odd number of vertices, then Laplacian perfect state transfer does not occur in XX.

We note however that if XX is a weighted join graph with an odd number of vertices, then it is possible to get PST in XX if we drop the condition that ϕ(L(X),t)[t]\phi(L(X),t)\in\mathbb{Z}[t]. Indeed, if XX is a weighted K2K_{2} with vertices uu and vv joined by an edge of weight ab>0\frac{a}{b}>0 where a>0a>0 is odd and b0b\equiv 0 (mod 4), then ϕ(L(X),t)[t]\phi(L(X),t)\notin\mathbb{Z}[t], but one checks that XO1X\vee O_{1} is a weighted K3K_{3} with PST between uu and vv at time 2π3\frac{2\pi}{3}.

Next, observe that XX is an induced subgraph of XYX\vee Y. In Corollaries 70 and 72, we showed for M{A,L}M\in\{A,L\} that ||UM(XY,t)u,v||UM(X,t)u,v||f(|V(X)|)\big{|}|U_{M}(X\vee Y,t)_{u,v}|-|U_{M}(X,t)_{u,v}|\big{|}\leq f(|V(X)|) for all tt, where f(|V(X)|)=2/|V(X)|f(|V(X)|)=2/|V(X)|, and this inequality is tight for infinite families. In general, if XX is a subgraph of ZZ induced by WV(Z)W\subseteq V(Z) with |W|2|W|\geq 2, then for u,vV(X)u,v\in V(X), we ask: what is an upper bound for ||UM(Z,t)u,v||UM(X,t)u,v||\big{|}|U_{M}(Z,t)_{u,v}|-|U_{M}(X,t)_{u,v}|\big{|} in terms of |W||W|? In particular, if XX is either complete or empty, what is supt>0||UM(Z,t)u,v||UM(X,t)u,v||\sup_{t>0}\big{|}|U_{M}(Z,t)_{u,v}|-|U_{M}(X,t)_{u,v}|\big{|}? In line with Corollary 73, we further ask: for which induced subgraphs XX of ZZ does ||UM(Z,t)u,v||UM(X,t)u,v||0\big{|}|U_{M}(Z,t)_{u,v}|-|U_{M}(X,t)_{u,v}|\big{|}\rightarrow 0 for all tt as |V(X)||V(X)|\rightarrow\infty?

If X=KmX=K_{m} and YY is a graph on mm vertices, then we have proof for M{A,L}M\in\{A,L\} that |UM(XY,t)u,v||UM(X,t)u,v|=0|U_{M}(X\vee Y,t)_{u,v}|-|U_{M}(X,t)_{u,v}|=0 if and only if tt belongs to some countable set. Hence, we pose the question, are there families of graphs such that for some set TT with positive measure, |UM(XY,t)u,v||UM(X,t)u,v|=0|U_{M}(X\vee Y,t)_{u,v}|-|U_{M}(X,t)_{u,v}|=0 for all tTt\in T?

Finally, one of the surprising results in this paper is that the join operation can induce PST in the resulting graph under mild conditions. Owing to the fact that adjacency PST is rare in unweighted graphs [Godsil2012a, Corollary 10.2], it would be interesting to determine other graph operations that induce PST in the resulting graph. In [bhattacharjya2023quantum, Theorem 12], the authors show that the blow-up operation exhibits this property for some families of graphs (such as the blow-up of two copies of K1,nK_{1,n}). However, under Cartesian products, PST between vertices (u,v)(u,v) and (w,x)(w,x) with (u,v)(w,x)(u,v)\neq(w,x) requires that either (i) uu and ww, and vv and xx admit PST in the underlying graphs (ii) uu and ww admit PST and v=xv=x is periodic in the underlying graphs or (iii) u=wu=w is periodic and vv and ww admit PST in the underlying graphs. Hence, the Cartesian product cannot induce PST in the resulting graph.

Acknowledgements

H. Monterde is supported by the University of Manitoba Faculty of Science and Faculty of Graduate Studies. S. Kirkland is supported by NSERC Discovery Grant RGPIN-2019-05408.

References