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

Spectral properties of the ladder-like Josephson junction array

Daryna Bukatova [email protected] Ivan O. Starodub [email protected] Yaroslav Zolotaryuk [email protected] Bogolyubov Institute for Theoretical Physics, National Academy of Sciences of Ukraine, Kyiv 03143, Ukraine
Abstract

In this paper theoretical analysis of the ladder-like multirow array of inductively coupled Josephson junctions is presented. An external dc current is applied at the top to each of the columns of the array and is extracted at the bottom of that column. The density of states of the Josephson plasma waves has a δ\delta-function term due to the flat band and 3N23N-2 singularities where NN is the number of rows. The spatial distribution of the amplitudes of the plasmon wave is computed analytically for any given value of the wavenumber qq. It is expressed through the orthogonal polynomials that are similar but not identical to the Chebyshev polynomials.

keywords:
Josephson junction arrays , plasma waves , flat bands , density of states

1 Introduction

Systems with dispersionless or flat bands have been an important research topic in recent years [1]. After the initial theoretical predictions for the fermionic systems [2, 3, 4], research in this area has spred further into Josephson junctions arrays [5, 6], Dirac materials [7, 8, 9], photonic lattices [10], magnets [11] and other materials. In many cases the flat bands have been observed experimentally.

Ladders of inductively coupled Josephson junctions also support flat bands. They have been studied in connection with various nonlinear wave phenomena such as fluxons [12, 13], discrete breathers [14, 15, 16, 17] and meandering [18]. It has been shown first in [19, 16, 17] that a completely flat band can appear in the simplest JJ ladder with two rows. In [20] the dispersion law for the Josephson plasma waves in the general case of NN rows that has a NN-fold degenerate flat band was obtained. We remark that besides the weak superconductivity, the ladder systems have been researched actively in various areas of modern physics ranging from spin ladders [21, 22] to integrable systems [23, 24].

This work is focused on more deep investigation of the linear wave properties in the multi-row ladder-like Josephson junction arrays. We derive the plasmon density of states and obtain a general analytical formula for the plasma wave amplitude. We also compute the complete set of the plasmon amplitudes in the horizontal and vertical subsystems.

The paper is organized as follows. The model of the ladder-like Josephson junction array is presented in the next section. Section 3 is devoted to the presentation of the main results. Discussions and conclusions are presented in the last section.

2 Model and equations of motion

A ladder-like quasi-one-dimensional Josephson junction array (JJA) or a Josephson transmission line is considered. The array forms a rectangular-shaped network as shown in Fig. 1, where the junctions (denoted by ×\times) lie the links that connect the vertices of the lattice. It is assumed that the array has a finite number of rows in the YY direction. We denote this number by NN. The number of columns in the XX direction is assumed to be very large, LxNL_{x}\gg N. For the sake of simplicity we will consider the LxL_{x}\to\infty limit. Each junction that belongs to the kkth row and mmth column is described by the Josephson phase ϕm,k(v,h)\phi^{(v,h)}_{m,k}, which is the difference of the wavefunction phases of the superconductors that form the particular junction. The letters in the superscript correspond to the vertical (v) or horizontal (h) junctions. Each column is biased by the current IBI_{B}. The array is anisotropic in the sense that the junctions that belong to the rows (horizontal junctions) and to the columns (vertical junctions) have different properties. This anisotropy is characterized by the parameter η\eta:

ηIc(h)Ic(v)=ChCv,\eta\equiv\frac{I^{(h)}_{c}}{I^{(v)}_{c}}=\frac{C_{h}}{C_{v}}, (1)

where Ic(h,v)I^{(h,v)}_{c} are respective critical currents and Ch,vC_{h,v} are respective capacitances of the horizontal and vertical junctions.

×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\times×\timesIBI_{B}IBI_{B}IBI_{B}IBI_{B}IBI_{B}IBI_{B}IBI_{B}IBI_{B}IBI_{B}IBI_{B}IBI_{B}IBI_{B}xxyy
Figure 1: Schematic representation of the biased ladder-like Josephson junction array with N=4N=4 rows. The junctions are marked by ×\times.

The two important limiting cases of the anisotropy parameter are: (i) η0\eta\to 0 - Josephson phases oscillate mainly in the horizontal subsystem and (ii) η\eta\to\infty - oscillations of the vertical junctions dominate over the horizontal ones.

The temporal dynamics of each Josephson junctions in the arrays can be described by the resistively and capacitatively shunted junction model [25]. Within this model this dynamics is governed by the set of differential equations

Cv,h2eIc(v,h)ϕ¨m,k(v,h)+sinϕm,k(v,h)=Im,k(v,h)Ic(v,h),\frac{C_{v,h}\hbar}{2eI_{c}^{(v,h)}}\ddot{\phi}^{(v,h)}_{m,k}+\sin\phi^{(v,h)}_{m,k}=\frac{I^{(v,h)}_{m,k}}{I^{(v,h)}_{c}}, (2)

where Im,k(v,h)I^{(v,h)}_{m,k} is the current passing through the ϕm,k(v,h)\phi^{(v,h)}_{m,k} junction. The dissipative term ϕ˙m,k(v,h)/(2eRv,h)\hbar\dot{\phi}^{(v,h)}_{m,k}/(2eR_{v,h}) is omitted because it will only add the imaginary decay factor to the dispesion laws that we are goint to study. Together with the flux quantization rule inside the array loop and the Kirchoff laws, a set of coupled discrete sine-Gordon equation is derived. The derivation procedure is based on the paper [26]. The set of equations is written explicitly in [20], it is quite cumbersome, so we do not repeat it here. As shown in Fig. 1, the constant bias current IBI_{B} is injected on the top of each column and, consequently, is extracted at the bottom of that column. Similarly to [20] we work under the assumption of the uniform stationary current flow along each column. An elementary cell of the JJA consists of 2N12N-1 junctions: NN horizontal and N1N-1 vertical.

It is convenient to introduce the dimensionless time τ=tωp\tau=t\omega_{p}, where ωp\omega_{p} is the Josephson plasma frequency. We also introduce the dimensionless inductance parameter that measures the array discreteness in XX direction. Both these parameters are written as follows:

ωp=2πIcvΦ0Cv,βL=2πLIc(v)Φ0,Φ0=πe,\omega_{p}=\sqrt{\frac{2\pi I_{c}^{v}}{\Phi_{0}C^{v}}},\;\;\;\beta_{L}=\frac{2\pi LI^{(v)}_{c}}{\Phi_{0}},\;\;\Phi_{0}=\frac{\pi\hbar}{e}, (3)

where Φ0\Phi_{0} is the magnetic flux quantum. The dispersion law is obtained after linearization of the equations of motion around the ground state

ϕm,k(v)=arcsin(IBIc(v))arcsinγ,k1,N1¯;ϕn,k(h)=0,k1,N¯;m.\phi^{(v)}_{m,k}=\arcsin\left(\frac{I_{B}}{I^{(v)}_{c}}\right)\equiv\arcsin\gamma,k\in\overline{1,N-1};\;\;\;\phi^{(h)}_{n,k}=0,\,k\in\overline{1,N};\;m\in{\mathbb{Z}}. (4)

This ground state describes the situation when the external bias flows uniformy along each of the columns.

3 Properties of the Josephson plasma waves

In this Section we discuss various properties of Josephson plasma waves (plasmons) in the quasi-one-dimensional ladder-like array.

3.1 Plasmon spectrum and its main properties

Josephson plasmons are elecromagnetic waves that propagate along the array in the XX direction. At this point we remind the Josephson plasmon dispersion law for the NN-row JJA obtained in [20]. This dispesion law consists is derived from the discrete system of the sine-Gordon equations that are linearized around the groundstate (4). It consists of 2N12N-1 branches:

ω0=1,\displaystyle\omega_{0}=1, (5)
ω±n2(q)=Wn+(q)±Wn(q)2+2(1+αn+1)1cosqηβL2,\displaystyle\omega^{2}_{\pm n}(q)=W^{+}_{n}(q)\pm\sqrt{{W^{-}_{n}(q)}^{2}+2(1+\alpha_{n+1})\frac{1-\cos q}{\eta\beta_{L}^{2}}}, (6)
Wn±(q)=12[1±(ωi2(q)+1+αn+1ηβL)],\displaystyle W^{\pm}_{n}(q)=\frac{1}{2}\left[1\pm\left(\omega^{2}_{i}(q)+\frac{1+\alpha_{n+1}}{\eta\beta_{L}}\right)\right], (7)
αn=12cosπ(n1)N,n=1,N1¯.\displaystyle\alpha_{n}=1-2\cos\frac{\pi(n-1)}{N},\;n=\overline{1,N-1}. (8)

In Eq. (6) the sign of the superscript in ω±\omega_{\pm} is defined only by the sign near the radical and not by the superscripts in W±W^{\pm}. The function ωi(q)\omega_{i}(q) is the plasmon dispersion law of the simple dc biased parallel JJA with only one row of vertical junctions [27]:

ωi(q)=1γ2+2βL(1cosq).\omega_{i}(q)=\sqrt{\sqrt{1-\gamma^{2}}+\frac{2}{\beta_{L}}(1-\cos q)}. (9)

The coefficients αn\alpha_{n} always lie in the interval ]1,3[]-1,3[. For N=2N=2 rows there is only one coefficient α2=1\alpha_{2}=1, for N=3N=3 there are two coefficients, α2=0\alpha_{2}=0, α3=2\alpha_{3}=2, for N=4N=4 αn=12cosπ(n1)4\alpha_{n}=1-2\cos\frac{\pi(n-1)}{4}\Rightarrow α2=12\alpha_{2}=1-\sqrt{2}, α3=1\alpha_{3}=1, α4=1+2\alpha_{4}=1+\sqrt{2} and so on.

The Josephson plasmon spectrum consists of 2N12N-1 branches. Within the enumeration used in Eqs. (5)-(18) these branches are positioned as follows: (i) highly dispersive branches 1<ω1<ω2<ωN11<\omega_{1}<\omega_{2}<\cdots\omega_{N-1}; (ii) flat band with the Jopsehson plasma frequency ω0=1\omega_{0}=1; (iii) almost flat branches ωN+1<<ω2<ω1<1\omega_{-N+1}<\cdots<\omega_{-2}<\omega_{-1}<1. Thus, the highly dispersive branches are numbered by the positive superscripts, lie above the flat band and look similar to the dispersion law of the simple parallel array ωi(q)\omega_{i}(q) but are shifted from each other. In the unbiased case γ=0\gamma=0 all the ω|n|\omega_{-|n|} branches become flat: ωN+1(q)=ωN+2(q)==ω2(q)=ω1(q)=ω0=1\omega_{-N+1}(q)=\omega_{-N+2}(q)=\cdots=\omega_{-2}(q)=\omega_{-1}(q)=\omega_{0}=1. Otherwise they form a very narrow subband that is confined in the following frequency intherval: (1γ2)1/4ωn<ωn(π)<1(1-\gamma^{2})^{1/4}\leq\omega_{-n}<\omega_{-n}(\pi)<1. Even for rather high values of γ\gamma this interval will be much smaller than 11.

Group velocity characterizes the speed of the plasmon wavepackage propagation. By looking at it one can have an idea how different Josephson plasmon modes transport energy. The group velocity for each of the ωn(q)\omega_{n}(q) modes of (5)-(6) has the following form:

vgr,0(q)=0,\displaystyle v_{gr,0}(q)=0, (10)
vgr,±n(q)=dω±n(q)dq=sinq2βLω±n(q)[1±\displaystyle v_{gr,\pm n}(q)=\frac{d\omega_{\pm n}(q)}{dq}=\frac{\sin q}{2\beta_{L}\omega_{\pm n}(q)}\left[1\pm\right.
±ωi2(q)+αn+1ηβL12Wn(q)2+2(1+αn+1)ηβL2(1cosq)],n=1,N1¯.\displaystyle\left.\pm\frac{\omega^{2}_{i}(q)+\frac{\alpha_{n+1}}{\eta\beta_{L}}-1}{2\sqrt{{W^{-}_{n}(q)}^{2}+\frac{2(1+\alpha_{n+1})}{\eta\beta_{L}^{2}}(1-\cos q)}}\right],\;n=\overline{1,N-1}. (11)

It is obviously zero for the flat branch ω0=1\omega_{0}=1, similarly vgr,n=0v_{gr,-n}=0 for the almost flat branches if the array is not biased, γ=0\gamma=0. If γ0\gamma\neq 0, the respective velocities for the n0n\neq 0 modes are shown in Fig. 2. There are three curves for each of the subbands, for the strongly dispersive branches ω1,2,3\omega_{1,2,3} and for the almost flat branches ω1,2,3\omega_{-1,-2,-3}.

Refer to caption
Refer to caption
Refer to caption
Figure 2: The group velocities for the strongly dispersive branches, n=1,2,3n=1,2,3 (black, left scale) and almost flat branches n=3,2,1n=-3,-2,-1 (red,right scale) for γ=0.3\gamma=0.3, βL=1.5\beta_{L}=1.5, N=4N=4 and η=0.5\eta=0.5 (a), η=1.5\eta=1.5 (b) and η=3\eta=3 (c).

Since ω1<ω2<ω3\omega_{1}<\omega_{2}<\omega_{3}, the same is true for the group velocities: vgr,1<vgr,2<vgr,3v_{gr,1}<v_{gr,2}<v_{gr,3}. The group velocity for the strongly dispersive modes, vgr,|n|v_{gr,|n|}, does not deviate strongly from the sinq\sin q dependence. The maximal values of vgr,|n|(q)v_{gr,|n|}(q) are positioned near q=π/2q=\pi/2. The situation is more complex for the almost flat modes. The dispersion law for these modes is parabolic near the origin q=0q=0 but then reaches a plateau. The inflexion point of the ω|n|(q)\omega_{-|n|}(q) dispersion laws which is the maximum of the group velocity shifts closer to the origin as η\eta increases. It should be noted that the following is true for the almost flat modes only close to the origin: vgr,1>vgr,2>vgr,3v_{gr,-1}>v_{gr,-2}>v_{gr,-3}. Away from the maximum these velocities begin to cross. The group velocity for the highest mode ω1\omega_{-1} is the sharpest and its maximum is the closest to the origin. Thus, we conclude that the group velocity for the almost flat modes is at least one order of magnitude smaller than for the strongly dispersive modes. However, the almost flat modes, especially the mode ω1\omega_{-1}, exhibit a sharp peak not far from the q=0q=0 point.

3.2 Plasmon density of states

The density of states (DoS) in various condensed matter systems contains useful information that can be accessed experimentally. We compute the Josephson plasmon density of states (DoS) per unit cell as a function of the frequency from the dispersion laws (5)-(6). The general expression for the plasmon DoS reads

ρ(ω)=1π{n=N+1;n0N1θ[ωωn(0)]θ[ωωn(π)]|dωn(q)dq|+δ(ω1)}.\rho(\omega)=\frac{1}{\pi}\left\{\sum_{n=-N+1;n\neq 0}^{N-1}\frac{\theta[\omega-\omega_{n}(0)]-\theta[\omega-\omega_{n}(\pi)]}{\left|\frac{d\omega_{n}(q)}{dq}\right|}+\delta(\omega-1)\right\}\,. (12)

The sum runs over the all non-flat bands n0n\neq 0, and the flat branch contribution is represented by the Diracs δ\delta-function term δ(ω1)\delta(\omega-1). The Heaviside θ\theta-function is used to account for the overlaping branches explicitly. This point will be discussed below.

Before discussing the plasmon DoS for the general array with an arbitrary number of rows let us mention the simplest case of the JJ array of parallelly shunted junctions. The dispersion law for the plasmons, ωi(q)\omega_{i}(q) is given by equation (9). The respective density of states reads

ρ(ω)=βLωπ1[1βL2(ω21γ2)]2.\rho(\omega)=\frac{\beta_{L}\omega}{\pi\sqrt{1-\left[1-\frac{\beta_{L}}{2}(\omega^{2}-\sqrt{1-\gamma^{2}})\right]^{2}}}. (13)

It it not hard to notice that this function has two singularities at the edges of the plasmonic band: ωi(0)=(1γ)1/4\omega_{i}(0)=(1-\gamma)^{1/4} and ωi(π)=1γ2+4/βL\omega_{i}(\pi)=\sqrt{\sqrt{1-\gamma^{2}}+4/\beta_{L}}. There is a minimum somewhere between these two singularities.

By extracting from the dispersion law (6) the dependence of the wavenumber qq on the frequency ω\omega it is possible to write down an exact expression for the plasmon density of states:

ρ(ω)=1π{2βLn=N+1;n0N1{θ[ωωn(0)]θ[ωωn(π)]}[1+sign(n)Zn(ω)]Sn(ω)ω+δ(ω1)},\rho(\omega)=\frac{1}{\pi}\left\{2\beta_{L}\sum_{n=-N+1;n\neq 0}^{N-1}\frac{\{\theta[\omega-\omega_{n}(0)]-\theta[\omega-\omega_{n}(\pi)]\}}{[1+\mbox{sign}(n)~{}Z_{n}(\omega)]S_{n}(\omega)}\omega+\delta(\omega-1)\right\}, (14)

where the auxilliary functions are given by

Zn(ω)=1γ2+ω|n|2(0)2+C¯n(ω)[1γ2ω|n|2(0)+C¯n(ω)]2+4[ω|n|2(0)1]C¯n(ω),\displaystyle Z_{n}(\omega)=\frac{\sqrt{1-\gamma^{2}}+\omega_{|n|}^{2}(0)-2+\bar{C}_{n}(\omega)}{\sqrt{\left[\sqrt{1-\gamma^{2}}-\omega_{|n|}^{2}(0)+\bar{C}_{n}(\omega)\right]^{2}+4\left[\omega_{|n|}^{2}(0)-1\right]\bar{C}_{n}(\omega)}}, (15)
Sn(ω)=1[1βL2C¯n(ω)]2,ω|n|2(0)=1+1+α|n|+1ηβL,\displaystyle S_{n}(\omega)=\sqrt{1-\left[1-\frac{\beta_{L}}{2}\bar{C}_{n}(\omega)\right]^{2}},\;\;\omega_{|n|}^{2}(0)=1+\frac{1+\alpha_{|n|+1}}{\eta\beta_{L}}, (16)
C¯n(ω)=[ω2ω|n|2(0)](ω21γ2)ω21,n=±1,,±(N1).\displaystyle\bar{C}_{n}(\omega)=\frac{\left[\omega^{2}-\omega_{|n|}^{2}(0)\right]\left(\omega^{2}-\sqrt{1-\gamma^{2}}\right)}{\omega^{2}-1},\;n=\pm 1,\ldots,\pm(N-1). (17)

and the sign(n)\mbox{sign}(n) is used in order to take into account the n>0n>0 or n<0n<0 branches. The quantity ω|n|2(0)=[1+(1+α|n|+1)/(ηβL)]\omega_{|n|}^{2}(0)=[1+(1+\alpha_{|n|+1})/(\eta\beta_{L})] is the minimum frequecy for the nnth strongly dispersive branch.

Before analyzing the plasmon DoS given by Eqs. (14) it is useful to discuss the behaviour of the edge frequency values ωn(0)\omega_{n}(0) and ωn(π)\omega_{n}(\pi) of the strongly dispersive modes (n>0n>0). They are shown in Fig. 3 as a function of the inverse anisotropy parameter 1/η1/\eta.

Refer to caption
Figure 3: Upper (red) and lower (blue) bonds of the strongly dispersive plasmon branches ω|n|\omega_{|n|}, n=1,2,3n=1,2,3, for the case of N=4N=4 rows with the dimensionless bias γ=0.1\gamma=0.1 and different values of βL\beta_{L}. Different type of lines (solid, -- or -\cdot-) correspond to a particular branch ωn(q)\omega_{n}(q).

In the limit 1/η1/\eta\to\infty (η0\eta\to 0) the branches are well separated and there are gaps between them. We remind that this limit corresponds to the situation when most of the plasma oscillations takes place in the horizontal subsystem. As η\eta increases the gaps start to close and the dispersion curve begin to overlap.

Being equipped with this knowledge we can analyze the plasmon DoS plots presented in Figs. 4-5. In those figures the DoS is plotted as a function of the plasmon frequency for different values of η\eta. The situation if Fig. 4(a) corresponds to the non-overlaping branches. For ω>1\omega>1 one can observe three well separated smooth curves with two van Hove singularities each that correspond to the edge frequency values. As the anisotropy is increased from η=0.25\eta=0.25 to η=0.5\eta=0.5 the separate branches begin to overlap [see Fig. 4(b)]. The gaps that were visible in the previous figure now are closed. The singularities remain but now ω2(0)<ω1(π)\omega_{2}(0)<\omega_{1}(\pi) and ω3(0)<ω2(π)\omega_{3}(0)<\omega_{2}(\pi).

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 4: Plasmon (log scale) DoS for the array with N=4N=4 rows and γ=0.3\gamma=0.3, βL=1.5\beta_{L}=1.5 and η=0.25\eta=0.25 (a), η=0.5\eta=0.5 (b), η=1\eta=1 (c) and η=5\eta=5 (d). Red vertical lines correspond to edge values ωn(0)\omega_{n}(0), ωn(π)\omega_{n}(\pi) for each of the plasmon branches.

Further increasing of η\eta causes the singularities at ω=ω1,2,3(0)\omega=\omega_{1,2,3}(0) to move further to the left, while the singularities at ω=ω1,2,3(π)\omega=\omega_{1,2,3}(\pi) move to the right. The limit η1\eta\gg 1 corresponds to the situation when the amplitude of the oscillations in the vertical subsystem is much larger then in the horizontal. In this limit the dispersion curves are nested inside each other rather deeply and the DoS singularities concentrate at the respective edges of the band [compare Figs. 4(c) and (d)].

We have not discussed yet the properties of the plasmon DoS in the frequency region ω<1\omega<1. At ω=1\omega=1 we have a perfect flat band and there is a solid vertical line in Fig. 4 to demonstrate it. The almost flat branches ω|n|(q)\omega_{-|n|}(q) are densely squeezed in the interval ω|n|(0)(1γ2)1/4ω|n|(q)ω|n|(π)<1\omega_{-|n|}(0)\equiv(1-\gamma^{2})^{1/4}\leq\omega_{-|n|}(q)\leq\omega_{-|n|}(\pi)<1. Even for rather large bias values this interval is very narrow. For example, (1γ2)1/40.931(1-\gamma^{2})^{1/4}\approx 0.931 for γ=0.5\gamma=0.5. The point q=0q=0 is (N1)(N-1)-fold degenerate. For the Brillouin zone edge we have ω3<ω2<ω1\omega_{-3}<\omega_{-2}<\omega_{-1}. The full expressions for ω|n|(π)\omega_{-|n|}(\pi) is rather complicated and, therefore, we omit it. So, the dispersion curves are completely nested inside each other. As a result, we always have NN van Hove singularities, one at ω|n|(0)\omega_{-|n|}(0) and N1N-1 singularities at ω|n|(π)\omega_{-|n|}(\pi). When η\eta is increased the singularities at ω|n|(π)\omega_{-|n|}(\pi) concentrate on the right side of the DoS dependence.

In figure 5 the dependence of the plasmon DoS on the discreteness parameter βL\beta_{L} is demonstrated.

Refer to caption
Refer to caption
Figure 5: Plasmon DoS (log scale) for the array with N=6N=6 rows, γ=0.2\gamma=0.2, η=5\eta=5 and βL=0.5\beta_{L}=0.5 (a) and βL=2.5\beta_{L}=2.5 (b). Red vertical lines have the same meaning as in the previous figure. The insets show the detailed picture around the ω|n|(π)\omega_{-|n|}(\pi) singularities.

There are N=6N=6 rows, therefore we observe 3N2=163N-2=16 van Hove singularities. We observe that the change of βL\beta_{L} does not bring any qualitative difference. The discreteness parameter influences the total width of the upper (ω>1\omega>1) plasmon band. If βL\beta_{L} increases, the discreteness of the array increases and band narrows; it widens otherwise. Since the width of the lower (ω<1\omega<1, almost flat) subband depends mainly on the bias current, no visible changes are seen there when βL\beta_{L} is varied. As a final remark we note that the minimal value of DoS is several orders of magnitude larger for the almost flat branches as compared to the strongly dispersive branches.

3.3 Classification of the plasmon modes

The spatial structure of the plasmon modes that correspond to each of the branches ωn(q)\omega_{n}(q) is given by the matrix equation

([ωi2(q)ω2]𝕀^N11eiqβLS^1eiqβLηSTD^hω2𝕀^N)(𝐀(v)𝐀(h))=0,\left(\begin{array}[]{cc}[\omega_{i}^{2}(q)-\omega^{2}]\hat{\mathbb{I}}_{N-1}&\frac{1-e^{-iq}}{\beta_{L}}\hat{S}\\ \frac{1-e^{iq}}{\beta_{L}\eta}S^{T}&\hat{D}_{h}-\omega^{2}\hat{\mathbb{I}}_{N}\end{array}\right)\left(\begin{array}[]{c}\mathbf{A}^{(v)}\\ \mathbf{A}^{(h)}\end{array}\right)=0, (18)

that appears directly from the equations of motion [20]. Here the vectors 𝐀(v,h)\mathbf{A}^{(v,h)} are amplitudes of the plasmon waves, 𝕀N\mathbb{I}_{N} is N×NN\times N unit matrix, D^h\hat{D}_{h} is a tridiagonal N×NN\times N matrix

D^h=1ηβL[1+ηβL1000012+ηβL100000012+ηβL1000011+ηβL],\displaystyle\hat{D}_{h}=\frac{1}{\eta\beta_{L}}\left[\begin{array}[]{ccccccc}\vspace{2mm}1+\eta\beta_{L}&-1&0&\cdots&0&0&0\\ \vspace{2mm}-1&2+\eta\beta_{L}&-1&\cdots&0&0&0\\ \vspace{2mm}\vdots&\vdots&\vdots&\ddots&\vdots&\vdots&\vdots\\ \vspace{2mm}0&0&0&\cdots&-1&2+\eta\beta_{L}&-1\\ \vspace{2mm}0&0&0&\cdots&0&-1&1+\eta\beta_{L}\\ \end{array}\right], (24)

and S^\hat{S} is a (N1)×N(N-1)\times N bidiagonal matrix with S11=S22==SN1,N1=1S_{11}=S_{22}=\cdots=S_{N-1,N-1}=1, S12=S23==SN1,N=1S_{12}=S_{23}=\cdots=S_{N-1,N}=-1 and Sij=0S_{ij}=0 elsewhere. We can introduce a normalized real (2N1)(2N-1)-component eigenvector

𝐞n(q)=(𝐞n(v)𝐞n(h)),\mathbf{e}_{n}(q)=\left(\begin{array}[]{c}\mathbf{e}_{n}^{(v)}\\ \mathbf{e}_{n}^{(h)}\end{array}\right), (25)

that consists of N1N-1 components for the vertical junctions and NN components for the horizontal junctions.

In the long-wave limit (q=0q=0) the horizontal and vertical subsystems decouple and the respective eigenvectors are easy to find. Previously only qualitative description of the eigenvectors [20] has been given. In this paper the complete set of solutions will be presented.

The flat branch ω0=1\omega_{0}=1

For all values of the wavenumber qq only horizontal subsystem is excited while all vertical junctions do not oscillate. It can be easily seen from the separate equations for each of the subsystems: [ωi2(q)1]𝕀N1𝐀(v)=0[\omega^{2}_{i}(q)-1]\mathbb{I}_{N-1}\mathbf{A}^{(v)}=0 for the vertical one and D^h𝐀(v)=0\hat{D}_{h}\mathbf{A}^{(v)}=0 for the horizontal one. Since usually flat bands appear due to destructive inteference, it easy to understand how such a dispersionless mode appears in a ladder: when wave propagates in the XX direction, all excitations in the YY direction cancel each other. The respective normalized eigenvector 𝐞0\mathbf{e}_{0} equals

𝐞0=1N(0,0,,0=𝐞0(v)1,1,,1=𝐞0(h))T.\mathbf{e}_{0}=\frac{1}{\sqrt{N}}\left(\underbrace{0,0,\cdots,0}_{=\mathbf{e}_{0}^{(v)}}\underbrace{1,1,\cdots,1}_{=\mathbf{e}_{0}^{(h)}}\right)^{T}. (26)

It can easily be checked that the 𝐀(v)=0\mathbf{A}^{(v)}=0, A1(h)==AN(h)A^{(h)}_{1}=\cdots=A^{(h)}_{N} state solves Eq. (18) for any qq and γ\gamma if ω=1\omega=1.

The long-wave limit q=0q=0 for the ωn=1\omega_{n}=1 modes at γ=0\gamma=0

Note that for γ=0\gamma=0 the ω=1\omega=1 branch is NN-fold degenerate. In the vertical subsystem we have 0𝕀N1𝐀(v)=00\cdot\mathbb{I}_{N-1}\mathbf{A}^{(v)}=0. The horizontal subsystem the equality A1(h)=A1(h)==AN(h)A^{(h)}_{1}=A^{(h)}_{1}=\cdots=A^{(h)}_{N} still holds. Thus, we can create the normalized eigenvectors that solve Eq. (18) and are orthogonal to 𝐞0\mathbf{e}_{0} and each other:

𝐞1(v)=(100),𝐞2(v)=(010),,𝐞N+1(v)=(001);\displaystyle\mathbf{e}_{-1}^{(v)}=\left(\begin{array}[]{c}1\\ 0\\ \cdots\\ 0\end{array}\right),\mathbf{e}_{-2}^{(v)}=\left(\begin{array}[]{c}0\\ 1\\ \cdots\\ 0\end{array}\right),\ldots,\mathbf{e}_{-N+1}^{(v)}=\left(\begin{array}[]{c}0\\ 0\\ \cdots\\ 1\end{array}\right); (39)
𝐞n(h)=(000),n=1,N1¯.\displaystyle\mathbf{e}_{-n}^{(h)}=\left(\begin{array}[]{c}0\\ 0\\ \cdots\\ 0\end{array}\right),\;\;n=\overline{1,N-1}. (44)

Almost flat bands ω|n|\omega_{-|n|}, n=1,N1¯n=\overline{1,N-1} for q=0q=0 and γ0\gamma\neq 0

In this case ω|n|(0)=(1γ2)1/41\omega_{-|n|}(0)=(1-\gamma^{2})^{1/4}\neq 1. This point is (N1)(N-1)-fold degenerate. The vertical subsystem is governed by the 0𝕀N1𝐀(v)=00\cdot\mathbb{I}_{N-1}\mathbf{A}^{(v)}=0 equation. Thus, the 𝐞|n|(v)\mathbf{e}_{-|n|}^{(v)} components of the respective eigenvector can be chosen the same as in Eq. (LABEL:16). One of the options is to keep components of the horizontal subsystem 𝐀h=0\mathbf{A}^{h}=0 since it is a solution. There is no other nontrivial solution (see A for details). Thus, we need to keep the same solutions as in the previous subparagraph.

Strongly dispersive branches ωn>0\omega_{n>0}

The susbsystems stay decoupled, the vertical subsystem is not exited while in the horizontal system the plasmon amplitudes are non-zero. The components of the eigenvector are expressed through the polynomials 𝒯k(ξ){\cal T}_{k}(\xi) with the recurrence relation 𝒯k+1(ξ)=(ξ+1)𝒯k+1(ξ)𝒯k1(ξ){\cal T}_{k+1}(\xi)=(\xi+1){\cal T}_{k+1}(\xi)-{\cal T}_{k-1}(\xi), which are defined in A:

𝐞n(h)=1𝒩n[1𝒯2(αn+1)𝒯3(αn+1)𝒯N1(αn+1)𝒯N(αn+1)],𝒩n=k=1N𝒯k2(αn+1).\displaystyle\mathbf{e}_{n}^{(h)}=\frac{1}{\sqrt{\cal N}_{n}}\left[\begin{array}[]{c}1\\ {\cal T}_{2}(-\alpha_{n+1})\\ {\cal T}_{3}(-\alpha_{n+1})\\ \cdots\\ {\cal T}_{N-1}(-\alpha_{n+1})\\ {\cal T}_{N}(-\alpha_{n+1})\end{array}\right],\;\;{\cal N}_{n}=\sum_{k=1}^{N}{\cal T}^{2}_{k}(-\alpha_{n+1}). (51)

There is also the boundary condition 𝒯N1(αn+1)=αn+1𝒯N1(αn+1){\cal T}_{N-1}(-\alpha_{n+1})=-\alpha_{n+1}{\cal T}_{N1}(-\alpha_{n+1}) which should be satisfied. In fact, it satisfies automatically. Note that sometimes αn+1=0\alpha_{n+1}=0. This happens, for example, for the N=3N=3 row array, for the mode ω1(0)\omega_{1}(0). In that case α2=0\alpha_{2}=0. In general, this will happen for arrays with N=3mN=3m, m=1,2,3,m=1,2,3,\dots, because there will always exist some mode nn with cosπ/3=1/2\cos\pi/3=1/2 in the coefficient αn+1\alpha_{n+1}. In those cases the recurrence relation together with the boundary conditions will produce the following normalized eigenvector:

𝐞n(h)=12N/3[1,0,1,1,0,1,1,0,1,1,0,1,,0,(1)N/3]T.\displaystyle\mathbf{e}_{n}^{(h)}=\frac{1}{\sqrt{2N/3}}\left[1,0,-1,-1,0,1,1,0,-1,-1,0,1,\ldots,0,(-1)^{N/3}\right]^{T}. (52)

The power (1)N/3(-1)^{N/3} in the last component can be easily explained by the fact that these eigenvectors consist of alternating triples (1,0,1)(1,0,-1) and depending on whether N/3N/3 is odd or even the eigenvector will end with 1-1 or 11.

Let us produce some examples. For the array with N=3N=3 rows there are two strongly dispersive modes ω1\omega_{1} and ω2\omega_{2}. There are two respective normalized eigenvectors:

𝐞n(v)=(00),𝐞1(h)=12(101),𝐞2(h)=16(121).\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\mathbf{e}_{n}^{(v)}=\left(\begin{array}[]{c}0\\ 0\end{array}\right),\mathbf{e}_{1}^{(h)}=\frac{1}{\sqrt{2}}\left(\begin{array}[]{c}1\\ 0\\ -1\end{array}\right),\mathbf{e}_{2}^{(h)}=\frac{1}{\sqrt{6}}\left(\begin{array}[]{c}1\\ -2\\ 1\end{array}\right). (53)

The eigenvector for ω1\omega_{1} has α2=0\alpha_{2}=0, thus its structure is defined by (52). Similarly, for the N=4N=4 rows there are three highly dispersive modes ω1,2,3\omega_{1,2,3} and the respective the eigenvectors read

𝐞n(v)=(000),𝐞1(h)=1222(121121),𝐞2(h)=12(1111),\displaystyle\mathbf{e}_{n}^{(v)}=\left(\begin{array}[]{c}0\\ 0\\ 0\end{array}\right),\mathbf{e}_{1}^{(h)}=\frac{1}{2\sqrt{2-\sqrt{2}}}\left(\begin{array}[]{c}1\\ \sqrt{2}-1\\ 1-\sqrt{2}\\ -1\end{array}\right),\mathbf{e}_{2}^{(h)}=\frac{1}{\sqrt{2}}\left(\begin{array}[]{c}1\\ -1\\ -1\\ 1\end{array}\right), (65)
𝐞3(h)=17+22(1121+21).\displaystyle\mathbf{e}_{3}^{(h)}=\frac{1}{\sqrt{7+2\sqrt{2}}}\left(\begin{array}[]{c}1\\ -1-\sqrt{2}\\ 1+\sqrt{2}\\ -1\end{array}\right). (70)

When these component are calculated the boundary condition for the last two components is satisfied automatically. The reader is welcome to check it.

For the N=6N=6 rows we have 55 strongly dispersive branches, the coefficients αn+1\alpha_{n+1} are: α2=13\alpha_{2}=1-\sqrt{3}, α3=0\alpha_{3}=0, α4=1\alpha_{4}=1, α5=2\alpha_{5}=2 and α6=1+3\alpha_{6}=1+\sqrt{3}. The eigenvector for the n=2n=2 branch (α3=0\alpha_{3}=0) can be easily computed from the recurrence relations: A1=1A_{1}=1, A2=0A_{2}=0, A3=A2A1=1A_{3}=A_{2}-A_{1}=-1, A4=A3A2=1A_{4}=A_{3}-A_{2}=-1, A5=0A_{5}=0, A6=A5A4=1A_{6}=A_{5}-A_{4}=1. Thus, we have 𝐞2(h)=(1/2)(1,0,1,1,0,1)T\mathbf{e}^{(h)}_{2}=(1/2)(1,0,-1,-1,0,1)^{T} which satisfies the general Eq. (52).

Non-flat bands for q0q\neq 0

Finally we discuss all branches ω±n(q)\omega_{\pm n}(q), n=1,2,,N1n=1,2,\ldots,N-1. In this case the phases from both the horizontal subsystem are involved in the dynamics. The amplitudes for the horizontal subsystems satisfy exactly the same relations (51) (see A for details). In fact, both the strongly dispersive (ω|n|\omega_{|n|}) and almost flat (ω|n|\omega_{-|n|}) modes have the same distribution in of the horizontal oscillatons. The amplitudes of the vertical subsystem can be expessed through the same polynomials 𝒯k{\cal T}_{k} (see A for details):

𝐞±n(v)(q)=1𝒩n(q)1cosqβL[ω±n2(q)ωi2(q)][1+αn+1𝒯2(αn+1)𝒯3(αn+1)𝒯3(αn+1)𝒯4(αn+1)𝒯N2(αn+1)𝒯N1(αn+1)𝒯N1(αn+1)𝒯N(αn+1)],\displaystyle\mathbf{e}^{(v)}_{\pm n}(q)=\frac{1}{\sqrt{{\cal N}_{n}(q)}}\frac{1-\cos{q}}{\beta_{L}[\omega^{2}_{\pm n}(q)-\omega^{2}_{i}(q)]}\left[\begin{array}[]{c}1+\alpha_{n+1}\\ {\cal T}_{2}(-\alpha_{n+1})-{\cal T}_{3}(-\alpha_{n+1})\\ {\cal T}_{3}(-\alpha_{n+1})-{\cal T}_{4}(-\alpha_{n+1})\\ \cdots\\ {\cal T}_{N-2}(-\alpha_{n+1})-{\cal T}_{N-1}(-\alpha_{n+1})\\ {\cal T}_{N-1}(-\alpha_{n+1})-{\cal T}_{N}(-\alpha_{n+1})\end{array}\right], (77)
(78)

with the normalization coefficient

𝒩n(q)\displaystyle{\cal N}_{n}(q) =\displaystyle= k=1N𝒯k2(αn+1)+(1cosq)2βL2[ω±n2(q)ωi2(q)]2×\displaystyle\sum_{k=1}^{N}{\cal T}_{k}^{2}(-\alpha_{n+1})+\frac{(1-\cos{q})^{2}}{\beta^{2}_{L}[\omega^{2}_{\pm n}(q)-\omega^{2}_{i}(q)]^{2}}\times
×\displaystyle\times {k=1N1[𝒯k(αn+1)𝒯k+1(αn+1)]2}.\displaystyle\left\{\sum_{k=1}^{N-1}\left[{\cal T}_{k}(-\alpha_{n+1})-{\cal T}_{k+1}(-\alpha_{n+1})\right]^{2}\right\}.

We see that different modes, ωn\omega_{n} and ωn\omega_{-n} with demonstrate the same dynamics in the horizontal subsystem. Thus, the difference between these two types of modes lie in the dynamics of the vertical phases. If one focuses on the coefficient in front of the vector in (78) in the limit of small qq’s, one can see that for the strongly dispersive modes it equals

12βL(11γ2+1+αn+1ηβL)q2+𝒪(q4).\frac{1}{2\beta_{L}\left(1-\sqrt{1-\gamma^{2}}+\frac{1+\alpha_{n+1}}{\eta\beta_{L}}\right)}q^{2}+{\cal O}(q^{4}). (80)

For the almost flat bands (ωn\omega_{-n}) the denominator of the same coefficient is q2\propto q^{2}. Its expansion has the following form:

11γ22βL1γ2(11γ2+1+αn+1ηβL)+𝒪(q2).\frac{1-\sqrt{1-\gamma^{2}}}{\sqrt{2}\beta_{L}\sqrt{1-\gamma^{2}}\left(1-\sqrt{1-\gamma^{2}}+\frac{1+\alpha_{n+1}}{\eta\beta_{L}}\right)}+{\cal O}(q^{2}). (81)

Thus, the vertical subsystem behaves differently for the two types of modes. The strongly dispersive modes do not have a constant term, while the almost flat ones do.

From the properties of the 𝒯k{\cal T}_{k} polynomials (see Fig. 6) we can see how the structure of the amplitudes of the vertical and horizontal junctions. They can be in phase, in anti-phase or some more complicated picture can appear. The special cases of particular vaues of αn+1=0,2-\alpha_{n+1}=0,-2 can easily be spotted. Here are some examples of the amplitude distribution in the vertical subsystem. For the N=3N=3 case the 𝐞±1(v)\mathbf{e}_{\pm 1}^{(v)} vector can be derived from the 𝐞±1(h)\mathbf{e}_{\pm 1}^{(h)} vector (53): 𝐞±1(v)(1,1)T\mathbf{e}_{\pm 1}^{(v)}\propto(1,1)^{T}. In the same way we have another vector 𝐞±2(v)(1,1)T\mathbf{e}_{\pm 2}^{(v)}\propto(1,-1)^{T}. So, the vertical subsystem have both the in-phase and anti-phase modes. As another example, let us consider the ω±2\omega_{\pm 2} mode of the N=6N=6 array. The horizontal part of the respective eigenvector was obtained previously: 𝐞±2(h)=(1/2)(1,0,1,1,0,1)T\mathbf{e}^{(h)}_{\pm 2}=(1/2)(1,0,-1,-1,0,1)^{T}. The respective vertical part is 𝐞±2(v)(1,1,0,1,1)T\mathbf{e}^{(v)}_{\pm 2}\propto(1,1,0,-1,-1)^{T}.

4 Conclusions

This work is devoted to the spectral properties of the Josephson plasma waves (Josephson plasmons) in the inductively coupled quasi-one-dimensional ladder-like array of Josephson junctions. An intriguing feature of this array is a NN-fold degenerate flat band in the unbiased case. If an external bias is applied the degeneracy is lifted for all wavenumbers except the q=0q=0 point and the set of very weakly dispersive or almost flat bands appears.

The plasmon density of states (DoS) has 3N23N-2 van Hove singularities that appear at the edge values of the individual branches and one δ\delta-function-like peak due to the flat band. There are NN singularities for the almost flat subband that lies under the ω=1\omega=1 (in the units of the Josephson plasma frequency) branch and 2(N1)2(N-1) for the strongly dispersive subband that lies above the ω=1\omega=1 branch.

The eigenvectors that describe the spatial distribution of the Josephson phase vibrations within the elementary cell have been obtained. They are expressed through the set of orthogonal polynomials that have a recurrence relation similar as in the Chebyshev polynomials but still different. It should be noted that Chebyshev-like polynomials of more general form have appeared before for an unrelated system that also involves nearest-neighbour lattice interations [30].

We believe that these results will be important for the studies of the nonlinear excitations in Josephson ladders. For example, fluxon interaction with the plasma waves in the simple parallel array gives a rise to the novel resonant effects [28]. Since the NN-row ladder has such a rich linear spectrum, it would be intriguing to study the fluxon-plasmon interaction in such a system.

Acknowledgements

We would like to thank the Armed Forces of Ukraine for providing security to perform this work. I.O.S. and Y.Z. acknowledge support from the National Research Foundation of Ukraine, grant (2023.03.0097) ”Electronic and transport properties of Dirac materials and Josephson junctions”.

Appendix A Details of the eigenvector calculation

In this Appendix the procedure of the plasmon amplitude calculation is presented. The plasmon amplitudes (𝐀(v),𝐀(h))(\mathbf{A}^{(v)},\mathbf{A}^{(h)}) do not depend on the row number mm, but depend on the column number. They are governed by the set of matrix equations (18) which can be rewritten as

[ωi2(q)ω2]𝕀^N1𝐀(v)+1eiqβLS^𝐀(h)=0\displaystyle[\omega_{i}^{2}(q)-\omega^{2}]{\hat{\mathbb{I}}}_{N-1}\mathbf{A}^{(v)}+\frac{1-e^{-iq}}{\beta_{L}}\hat{S}\mathbf{A}^{(h)}=0 (82)
1eiqβLηS^T𝐀(v)+(D^hω2𝕀^N)𝐀(h)=0.\displaystyle\frac{1-e^{iq}}{\beta_{L}\eta}\hat{S}^{T}\mathbf{A}^{(v)}+(\hat{D}_{h}-\omega^{2}{\hat{\mathbb{I}}}_{N})\mathbf{A}^{(h)}=0. (83)

Here ω\omega is the plasmon frequency that belongs to any of the branches (5)-(6). After substituting the explicit form of the matrices D^h\hat{D}_{h}, S^\hat{S} and S^T\hat{S}^{T} we rewrite the equations (82)-(83). The first equation allows us to express the vertical amplitudes through the horizontal ones:

[ωi2(q)ωn2]Ak(v)+1eiqβL(Ak(h)Ak+1(h))=0,k=1,N1¯.\displaystyle[\omega_{i}^{2}(q)-\omega_{n}^{2}]A_{k}^{(v)}+\frac{1-e^{-iq}}{\beta_{L}}\left(A_{k}^{(h)}-A_{k+1}^{(h)}\right)=0,\;k=\overline{1,N-1}. (84)

The second equation takes the following form

(1eiq)(Ak(v)Ak1(v))Ak1(h)Ak+1(h)+\displaystyle({1-e^{iq}})\left(A_{k}^{(v)}-A_{k-1}^{(v)}\right)-A_{k-1}^{(h)}-A^{(h)}_{k+1}+ (85)
+[βLη(1ω2)+2]Ak(h)=0,k=2,N1¯,\displaystyle+[\beta_{L}\eta(1-\omega^{2})+2]A_{k}^{(h)}=0,\;k=\overline{2,N-1},
(1eiq)A1(v)+[βLη(1ω2)+1]A1(h)A2(h)=0,\displaystyle({1-e^{iq}})A_{1}^{(v)}+[\beta_{L}\eta(1-\omega^{2})+1]A_{1}^{(h)}-A_{2}^{(h)}=0, (86)
(1eiq)AN(v)+AN1(h)[βLη(1ω2)+1]AN(h)=0.\displaystyle({1-e^{iq}})A_{N}^{(v)}+A_{N-1}^{(h)}-[\beta_{L}\eta(1-\omega^{2})+1]A_{N}^{(h)}=0~{}. (87)

In the long wave limit q=0q=0 the prefactors 1e±iq1-e^{\pm iq} in Eqs. (82)-(83) vanish and, consequently, the vertical and horizontal subsystems are decoupled:

[1γ2ω2]𝕀^N1𝐀(v)=0,[D^hω2𝕀^N]𝐀(h)=0.[\sqrt{1-\gamma^{2}}-\omega^{2}]\hat{\mathbb{I}}_{N-1}\mathbf{A}^{(v)}=0,\;\;[\hat{D}_{h}-\omega^{2}\hat{\mathbb{I}}_{N}]\mathbf{A}^{(h)}=0. (88)

The strongly degenerate case of q=0q=0, γ=0\gamma=0 has been discussed in the main text.

For the dispersive modes ωn(0)1\omega_{n}(0)\neq 1, n0n\neq 0, the amplitudes in the horizontal subsystem generally satisfy the system

A2(h)=[βLη(1ω2)+1]A1(h),\displaystyle A_{2}^{(h)}=[\beta_{L}\eta(1-\omega^{2})+1]A_{1}^{(h)}, (89)
AN(h)=[βLη(1ω2)+1]1AN1(h),\displaystyle A_{N}^{(h)}=[\beta_{L}\eta(1-\omega^{2})+1]^{-1}A_{N-1}^{(h)}, (90)
Ak+1(h)[βLη(1ω2)+2]Ak(h)+Ak1(h)=0,k=2,N1¯.\displaystyle A_{k+1}^{(h)}-[\beta_{L}\eta(1-\omega^{2})+2]A_{k}^{(h)}+A_{k-1}^{(h)}=0,\;k=\overline{2,N-1}. (91)

The solutions of this system are real. It appears that these horizontal amplitudes satisfy the recurrence relation (91). Without loss of generality we can assume A1(h)=1A_{1}^{(h)}=1, because if A1(h)=0A_{1}^{(h)}=0 then all other amplitudes must be zero. Then the components from k=2k=2 through k=N1k=N-1 can be expressed via the following polynomials:

𝒯1(ξ)=1,\displaystyle{\cal T}_{1}(\xi)=1, (92)
𝒯2(ξ)=ξ,\displaystyle{\cal T}_{2}(\xi)=\xi, (93)
𝒯3(ξ)=(1+ξ)𝒯2(ξ)𝒯1(ξ)=ξ2+ξ1,\displaystyle{\cal T}_{3}(\xi)=(1+\xi){\cal T}_{2}(\xi)-{\cal T}_{1}(\xi)=\xi^{2}+\xi-1, (94)
𝒯4(ξ)=(1+ξ)𝒯3(ξ)𝒯2(ξ)=ξ3+2ξ2ξ1,\displaystyle{\cal T}_{4}(\xi)=(1+\xi){\cal T}_{3}(\xi)-{\cal T}_{2}(\xi)=\xi^{3}+2\xi^{2}-\xi-1, (95)
\displaystyle\;\;\;\cdots\cdots\cdots
𝒯k+1(ξ)=(1+ξ)𝒯k(ξ)𝒯k1(ξ).\displaystyle{\cal T}_{k+1}(\xi)=(1+\xi){\cal T}_{k}(\xi)-{\cal T}_{k-1}(\xi). (96)

These polynomials look similar to the well-known Chebyshev polynomials [29] but their recurrence relation is different. See the respective graphs in Fig. 6.

Refer to caption
Refer to caption
Figure 6: Example of several polynomials 𝒯k{\cal T}_{k} (a) and Chebyshev polynomials (b) of the first kind.

For the case of almost flat branches, ω|n|\omega_{-|n|}, q=0q=0, γ0\gamma\neq 0 the horizontal subsystem can be rewritten through these polynomials as functions of ζ=1+βLη(11γ2)\zeta=1+\beta_{L}\eta(1-\sqrt{1-\gamma^{2}}): 𝒯1(ζ)=1{\cal T}_{1}(\zeta)=1, 𝒯2(ζ)=ζ{\cal T}_{2}(\zeta)=\zeta and so on. Since ζ0\zeta\neq 0, the last component must satisfy the boundary condition

ζ𝒯N(ζ)=𝒯N1(ζ).\zeta{\cal T}_{N}(\zeta)={\cal T}_{N-1}(\zeta). (97)

At the first glance, we have a nontrivial solution for this case. However, it is not true because it is impossible to satisfy (97) for ζ>1\zeta>1. Indeed, for ζ>1\zeta>1 all the polynomials 𝒯k{\cal T}_{k} for ζ>1\zeta>1 are monotonically increasing functions (see Fig. 6). The leading term of the 𝒯k(ζ){\cal T}_{k}(\zeta) polynomial for ζ>1\zeta>1 is ζk1\zeta^{k-1}. There is no way that two functions, 𝒯k(ζ){\cal T}_{k}(\zeta) and ζ𝒯k+1(ζ)\zeta{\cal T}_{k+1}(\zeta) would cross anywhere for ζ>1\zeta>1.

In the case of strongly dispersive branches ω|n|(0)\omega_{|n|}(0) we still have to express the solution of the horizontal subsystem (89)-(91) through the same polynomials but arguments of the polynomials will be different. According to (6) the plasmon frequency for the strongly dispersive mode is

ωn(0)=1+1ηβL(1+αn+1),n=1,N1¯.\omega_{n}(0)=\sqrt{1+\frac{1}{\eta\beta_{L}}(1+\alpha_{n+1})},\;\;n=\overline{1,N-1}. (98)

We substitute it instead of ω\omega in Eqs. (89)-(91). As a result, we have the polynomials as functions of

ξn=βLη[1ωn2(0)]+1=αn+1=1+2cosπnN.\xi_{n}=\beta_{L}\eta[1-\omega_{n}^{2}(0)]+1=-\alpha_{n+1}=-1+2\cos\frac{\pi n}{N}. (99)

In some cases we may have αn=0\alpha_{n}=0. Then AN1(h)=0A^{(h)}_{N-1}=0 due to use the boundary condition (90), A2(h)=0A^{(h)}_{2}=0, A3(h)=A2(h)=1A^{(h)}_{3}=-A^{(h)}_{2}=-1 and so on according to the recurrence (96): 𝒯k+1(0)=𝒯k(0)𝒯k1(0){\cal T}_{k+1}(0)={\cal T}_{k}(0)-{\cal T}_{k-1}(0). The components will form the following sequence: (1,0,1,1,0,1,1,0,1,1,0,1,1)(1,0,-1,-1,0,1,1,0,-1,-1,0,1,1\ldots). This can happen only if n/N=1/3n/N=1/3 in (99), thus the dimension of the vector 𝐀(h)\mathbf{A}^{(h)} should be divisible by 33. In that case the last component would be +1+1 or 1-1. Hence, they can be written in the followin general form:

𝐀n(h)(q)=[𝒯1(αn+1)𝒯2(αn+1)𝒯3(αn+1)𝒯N1(αn+1)𝒯N(αn+1)],n=1,N1¯.\mathbf{A}^{(h)}_{n}(q)=\left[\begin{array}[]{c}{\cal T}_{1}(-\alpha_{n+1})\\ {\cal T}_{2}(-\alpha_{n+1})\\ {\cal T}_{3}(-\alpha_{n+1})\\ \cdots\\ {\cal T}_{N-1}(-\alpha_{n+1})\\ {{\cal T}_{N}(-\alpha_{n+1})}\end{array}\right],\;n=\overline{1,N-1}. (100)

It can be shown that the boundary condition (90) is satisfied automatically.

Finally, consider the the general case q0q\neq 0 for all modes ω±n(q)\omega_{\pm n}(q), n=1,N1¯n=\overline{1,N-1}. We substitute Ak(v)A^{(v)}_{k} from Eq. (84) into Eqs. (85)-(87). As a result we get a pair of equations for the top and bottom of the array

A2(h)=ξ±nA1(h),ξ±nAN(h)=AN1(h).\displaystyle A^{(h)}_{2}=\xi_{\pm n}A^{(h)}_{1},\;\xi_{\pm n}A^{(h)}_{N}=A^{(h)}_{N-1}. (101)

Similarly, for the rest of the primitive cell (k1k\neq 1 or kNk\neq N):

Ak1(h)+Ak+1(h)(1+ξ±n)Ak(h)=0,A_{k-1}^{(h)}+A_{k+1}^{(h)}-(1+\xi_{\pm n})A_{k}^{(h)}=0, (102)

where

ξ±n=1+ηβL(1ω±n2)(ω±n2ωi2)ω±n21γ2.\xi_{\pm n}=1+\eta\beta_{L}\frac{(1-\omega_{\pm n}^{2})(\omega_{\pm n}^{2}-\omega_{i}^{2})}{\omega_{\pm n}^{2}-\sqrt{1-\gamma^{2}}}. (103)

Since the dispersion law satisfies the biquadratic equation

(ω±n2ωi2)(ω±n21)αn+1+1ηβL(ω±n2ωi2)2(1+αn+1)(1cosq)ηβL2=0,(\omega_{\pm n}^{2}-\omega_{i}^{2})(\omega_{\pm n}^{2}-1)-\frac{\alpha_{n+1}+1}{\eta\beta_{L}}(\omega_{\pm n}^{2}-\omega_{i}^{2})-\frac{2(1+\alpha_{n+1})(1-\cos q)}{\eta\beta^{2}_{L}}=0, (104)

it is not difficult to figure out that the parameter ξ±n\xi_{\pm n} in (103) does not depend on qq, moreover, it equals:

ξ±n=αn+1=1+2cosπnN,\xi_{\pm n}=-\alpha_{n+1}=-1+2\cos\frac{\pi n}{N}, (105)

exactly in the same way as in the q=0q=0 case. Thus, the horizontal components of the plasmon wave coincide with the q=0q=0 case (100).

The eigenvector components 𝐀(v)\mathbf{A}^{(v)} for the vertical subsystem can be easily written with the help of Eq. (84). In order to have only the real part of the amplitude we compute the following quantity:

12[A(v)+c.c.]=1cosqβL[ω±n2(q)ωi2(q)](A1(h)A2(h)A2(h)A3(h)AN2(h)AN1(h)AN1(h)AN(h))=\displaystyle\frac{1}{2}[A^{(v)}+c.c.]=\frac{1-\cos{q}}{\beta_{L}[\omega^{2}_{\pm n}(q)-\omega^{2}_{i}(q)]}\left(\begin{array}[]{c}A^{(h)}_{1}-A^{(h)}_{2}\\ A^{(h)}_{2}-A^{(h)}_{3}\\ \cdots\\ A^{(h)}_{N-2}-A^{(h)}_{N-1}\\ A^{(h)}_{N-1}-A^{(h)}_{N}\end{array}\right)= (111)
=1cosqβL[ω±n2(q)ωi2(q)][1+αn+1αn+1𝒯3(αn+1)𝒯3(αn+1)𝒯4(αn+1)𝒯N2(αn+1)𝒯N1(αn+1)𝒯N1(αn+1)𝒯N(αn+1)].\displaystyle=\frac{1-\cos{q}}{\beta_{L}[\omega^{2}_{\pm n}(q)-\omega^{2}_{i}(q)]}\left[\begin{array}[]{c}1+\alpha_{n+1}\\ -\alpha_{n+1}-{\cal T}_{3}(-\alpha_{n+1})\\ {\cal T}_{3}(-\alpha_{n+1})-{\cal T}_{4}(-\alpha_{n+1})\\ \cdots\\ {\cal T}_{N-2}(-\alpha_{n+1})-{\cal T}_{N-1}(-\alpha_{n+1})\\ {\cal T}_{N-1}(-\alpha_{n+1})-{\cal T}_{N}(-\alpha_{n+1})\end{array}\right]. (118)

As a final remark, we would like to note that the polynomials (92)-(96) are orthogonal due to the Farvard’s theorem [31].

References

  • [1] S. F. D. Leykam, A. Andreanov, Artificial flat band systems: from lattice models to experiments, Adv. in Phys. 3 (2018) 1473052.
  • [2] B. Sutherland, Localization of electronic wave functions due to local topology, Phys. Rev. B 34 (1986) 5208–5211.
  • [3] E. H. Lieb, Two theorems on the hubbard model, Phys. Rev. Lett. 62 (1989) 1201–1204.
  • [4] V. A. Khodel’, V. R. Shaginyan, Superfluidity in system with fermion condensate, JETP Lett. 51 (1990) 553–555.
  • [5] I. M. Pop, K. Hasselbach, O. Buisson, W. Guichard, B. Pannetier, I. Protopopov, Measurement of the current-phase relation in josephson junction rhombi chains, Phys. Rev. B 78 (2008) 104504.
  • [6] A. Andreanov, M. Fistul, Resonant frequencies and spatial correlations in frustrated arrays of josephson type nonlinear oscillators, J. Phys. A: Math. Theor. 52 (2019) 105101.
  • [7] T. Heikkilä, G. Volovik, Dimensional crossover in topological matter: Evolution of the multiple dirac point in the layered system to the flat band on the surface, JETP Lett. 93 (2011) 59–65.
  • [8] G. E. Volovik, Graphite, graphene, and the flat band superconductivity, JETP Lett. 107 (2018) 516–517.
  • [9] E. V. Gorbar, V. P. Gusynin, D. O. Oriekhov, Gap generation and flat band catalysis in dice model with local interaction, Phys. Rev. B 103 (2021) 155155.
  • [10] R. A. Vicencio, C. Cantillano, L. Morales-Inostroza, B. Real, C. Mejía-Cortés, S. Weimann, A. Szameit, M. I. Molina, Observation of localized states in lieb photonic lattices, Phys. Rev. Lett. 114 (2015) 245503.
  • [11] O. Derzhko, A. Honecker, J. Richter, Low-temperature thermodynamics for a flat-band ferromagnet: Rigorous versus numerical results, Phys. Rev. B 76 (2007) 220402.
  • [12] W. Yu, K. H. Lee, D. Stroud, Vortex motion in Josephson-junction arrays near f=0 and f=1/2, Phys. Rev. B 47 (1993) 5906–5914.
  • [13] S. G. Lachenmann, T. Doderer, D. Hoffmann, R. P. Huebener, P. A. A. Booi, S. P. Benz, Observation of vortex dynamics in two-dimensional josephson-junction arrays, Phys. Rev. B 50 (1994) 3158–3164.
  • [14] L. M. Floria, J. L. Marín, P. J. Martinez, F. Falo, S. Aubry, Intrinsic localization in the dynamics of a josephson-junction ladder, Europhys. Lett. 36 (1996) 539.
  • [15] E. Trías, J. J. Mazo, T. P. Orlando, Discrete breathers in nonlinear lattices: Experimental detection in a josephson array, Phys. Rev. Lett. 84 (4) (2000) 741–744.
  • [16] P. Binder, D. Abraimov, A. V. Ustinov, S. Flach, Y. Zolotaryuk, Observation of breathers in josephson ladders, Phys. Rev. Lett. 84 (4) (2000) 745–748.
  • [17] A. E. Miroshnichenko, S. Flach, M. V. Fistul, Y. Zolotaryuk, J. B. Page, Breathers in josephson junction ladders: Resonances and electromagnetic wave spectroscopy, Phys. Rev. E 64 (6) (2001) 066601.
  • [18] D. Abraimov, P. Caputo, G. Filatrella, M. V. Fistul, G. Y. Logvenov, A. V. Ustinov, Broken symmetry of row switching in 2d josephson junction arrays, Phys. Rev. Lett. 83 (1999) 5354–5357.
  • [19] S. Flach, M. Spicci, Rotobreather dynamics in underdamped josephson junction ladders, J. Phys.: Cond. Matt. 11 (1) (1999) 321–334.
  • [20] D. Bukatova, Y. Zolotaryuk, Flat and almost flat bands in the quasi-one-dimensional josephson junction array, J. Phys.: Condens. Matte 34 (2022) 175402.
  • [21] E. Dagotto, Experiments on ladders reveal a complex interplay between a spin-gapped normal state and superconductivity, Reports on Progress in Physics 62 (11) (1999) 1525.
  • [22] G. Schmiedinghoff, L. Müller, U. Kumar, G. Uhrig, B. Fauseweh, Three-body bound states in antiferromagnetic spin ladders, Communications Physics 5 (2022) 218.
  • [23] O. O. Vakhnenko, Nonlinear integrable dynamics of coupled vibrational and intra-site excitations on a regular one-dimensional lattice, Physics Letters, Section A: General, Atomic and Solid State Physics 405 (2021) 127431.
  • [24] O. Vakhnenko, V. Vakhnenko, Development and analysis of novel integrable nonlinear dynamical systems on quasi-one-dimensional lattices. parametrically driven nonlinear system of pseudo-excitations on a two-leg ladder lattice, Ukr. J. Phys. 69 (8) (2024) 577–590.
  • [25] A. Barone, G. Paterno, Physics and Applications of the Josephson Effect, Wiley, New York, 1982.
  • [26] M. Barahona, S. Watanabe, Row-switched states in two-dimensional underdamped josephson-junction arrays, Physical Review B 57 (5 1998).
  • [27] S. Watanabe, H. S. J. van der Zant, S. H. Strogatz, T. P. Orlando, Dynamics of circular arrays of josephson junctions and the discrete sine-gordon equations, Physica D 97 (1996) 429–470.
  • [28] A. V. Ustinov, M. Cirillo, B. A. Malomed, Fluxon dynamics in one-dimensional josephson-junction arrays, Phys. Rev. B 47 (1993) 8357–8360.
  • [29] M. Abramowitz, I. Stegun, Pocketbook of Mathematical Functions, Verlag Harri Deutsch, Frankfurt/Main, 1984.
  • [30] Y. Zolotaryuk, J. C. Eilbeck, , Phys. Rev. B, Analytical approach to the Davydov-Scott theory with on-site potential, Phys. Rev. B 63 (2001) 054302.
  • [31] J. Favard, Sur les polynomes de Tchebicheff, C. R. Acad. Sci. Paris (in French) 200 (1935) 2052–2053.