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

Non-Hermitian topological phases and exceptional lines in topolectrical circuits

S M Rafi-Ul-Islam e0021595@u.nus.edu Department of Electrical and Computer Engineering, National University of Singapore, Singapore    Zhuo Bin Siu elesiuz@nus.edu.sg Department of Electrical and Computer Engineering, National University of Singapore, Singapore    Mansoor B.A. Jalil elembaj@nus.edu.sg Department of Electrical and Computer Engineering, National University of Singapore, Singapore
Abstract

We propose a scheme to realize various non-Hermitian topological phases in a topolectrical (TE) circuit network consisting of resistors, inductors, and capacitors. These phases are characterized by topologically protected exceptional points and lines. The positive and negative resistive couplings RgR_{g} in the circuit provide loss and gain factors which break the Hermiticity of the circuit Laplacian. By controlling RgR_{g}, the exceptional lines of the circuit can be modulated, e.g., from open curves to closed ellipses in the Brillouin zone. In practice, the topology of the exceptional lines can be detected by the impedance spectra of the circuit. We also considered finite TE systems with open boundary conditions, the admittance spectrum of which exhibits highly tunable zero-admittance states demarcated by boundary points (BPs). The phase diagram of the system shows topological phases which are characterized by the number of their BPs. The transition between different phases can be controlled by varying the circuit parameters and tracked via impedance readout between the terminal nodes. Our TE model offers an accessible and tunable means of realizing different topological phases in a non-Hermitian framework, and characterizing them based on their boundary point and exceptional line configurations.

.1 Introduction

There is growing interest in studying topological states in various platforms such as topological insulatorsFu and Kane (2007); Moore and Balents (2007), cold atomsLiu et al. (2013); Zhang et al. (2018), photonics systems Lu et al. (2014); Khanikaev et al. (2013), superconductorsSato and Ando (2017), and optical latticesGoldman et al. (2016); Lang et al. (2012) due to their extraordinary properties such as topologically protected edge states and unconventional transport characteristics Rafi-Ul-Islam et al. (2020a, b). Such topological states have been studied in Hermitian and lossless systemsRoy (2009), where the eigenenergies are always real. However, Hermitian systems do not exhibit many interesting phenomena such as exceptional pointsAlvarez et al. (2018a), the skin effectYao and Wang (2018); Lee and Thomale (2019); Rafi-Ul-Islam et al. (2021), biorthogonal bulk polarization Kunst et al. (2018), and wave amplification and attenuation El-Ganainy et al. (2018); Jin and Song (2019). In the pursuit of more exotic characteristics in topological phases, researchers have shifted attention from Hermitian to non-Hermitian systemsLiu et al. (2019). In contrast to Hermitian systems, non-Hermitian systems in general exhibit complex eigenvalues unless the system obeys some specific symmetries such as the 𝒫𝒯\mathcal{PT} symmetry where P\mathit{P} and 𝒯\mathcal{T} are the parity and time reversal operations, respectively. One iconic feature of non-Hermitian systems is the existence of exceptional points, where two or more eigenvectors coalesce and the Hamiltonian becomes nondiagonalizable. This feature leads to many novel transport phenomena, such as unidirectional transparencyLin et al. (2011); Feng et al. (2013), unconventional reflectivity Zhu et al. (2018a), and super sensitivityLiu et al. (2016). Generally, the exchange of energy or particles between lattice points and the surrounding environment induces non-Hermiticity in the system. One way to induce non-Hermiticity is to add imaginary onsite potentials at different sublattices that represent gain or loss in the system depending on the sign of the potentials. In addition, asymmetric sublattice couplings may also induce non-Hermiticity in the system Hamiltonian Rafi-Ul-Islam et al. (2021). However, realizing non-Hermitian systems in condensed matterAlvarez et al. (2018b), acoustic metamaterialsAchilleos et al. (2017), and optical structuresZhang et al. (2016); Jin (2017) is in practice difficult because of the limited control over sublattice couplings, instability of the complex eigenspectraYuce and Oztas (2018a), and limitations in experimental accessibility. In pursuit of alternative platforms to overcome the aforementioned experimental limitations, topolectrical (TE) circuits Rafi-Ul-Islam et al. (2020c); Lee et al. (2018); Imhof et al. (2018); Helbig et al. (2019); Rafi-Ul-Islam et al. (2020d) have emerged as an ideal platform to not only realize non-Hermitian systems, but also to investigate many emerging phenomena such as Chern insulatorsHaenel et al. (2019), the quantum spin Hall effectZhu et al. (2019); Sun et al. (2020); Zhu et al. (2018b); Sun et al. (2019), higher-order topological insulatorsEzawa (2018); Peterson et al. (2018), topological corner modesImhof et al. (2018); Serra-Garcia et al. (2019), Klein tunneling Rafi-Ul-Islam et al. (2020a, d) and perfect reflection Rafi-Ul-Islam et al. (2020a, b). Appropriately designed TE circuits can emulate the topological properties of materials and offer unparalleled degrees of tunability and experimental flexibility through the conceptual shift from conventional materials system to artificial electrical networks. The freedom in design and control over lattice couplings allow us to investigate electronic structures beyond the limitations of condensed matter systems. Moreover, TE circuit networks are not constrained by the physical dimension or distance between two lattice (nodal) points but are described solely by the mutual connectivity of the circuit nodes. Besides, the freedom of choice in the connections at each node and long-range hopping make TE systems easier to fabricate compared to real material systems. Therefore, it is also possible to design an equivalent circuit network in lower dimensions that resembles the characteristics of higher-dimensional circuits Rafi-Ul-Islam et al. (2020d). Unlike real material systems, an infinite real system can be mimicked by finite-sized circuit networks in TE circuits. Therefore, TE networks enable us to design a non-Hemitian system in a RLC\mathrm{RLC} circuit network with better measurement accessibility as all the characteristic variables such as the admittance bandstructure and the density of states can be evaluated through electrical measurementsHofmann et al. (2019)(e.g. impedance, voltage, and current readings). In this paper, we investigate the exceptional lines, i.e. loci of exceptional points (EP), in a non-Hermitian TE system consisting of electrical components such as resistors, inductors and capacitors as a function of the non-Hermitian parameter (i.e. resistance). We show that by introducing non-Hermiticity to the circuit Laplacian through the insertion of positive or negative resistances between the voltage nodes and the ground to induce imaginary onsite potentials in the lattice sites and tuning the non-Hermitian parameters appropriately through using by using appropriate resistance values, the loci of the EPs may be easily switched to take the form of either a line or closed curves such as ellipses in the Brillouin zone. We show that a unique property of the TE platform over earlier works Rui et al. (2019); Choi et al. (2019); Stegmaier et al. (2020), viz. the impedance spectrum, is that the exceptional lines can be detected and easily tracked by measuring the impedance spectrum of the circuit. We further investigate finite systems with open boundary conditions along one direction, and find that these finite systems possess a rich phase diagram with different phases possessing up to four pairs of BPs, depending on the circuit parameters. We also show that edge states in Hermitian or pure LC\mathrm{LC} circuits, become hybridized with the bulk modes in non-Hermitian RLC circuits. In summary, our TE model provides an experimentally accessible means to investigate the phase diagram and various topological phases of non-Hermitian systems.

.2 Theoretical Model

We consider a two-dimensional TE circuit, shown in Fig. 1, which is composed of capacitors, inductors, positive resistors (loss elements) and negative resistors (gain elements). The circuit has a unit cell consisting of two sublattice nodes AA and BB. Each node is connected by a capacitor of ±C1\pm C_{1} and CyC_{y} to its nearest neighbour in the xx and yy-direction, respectively, and a parallel combination of a common capacitor CC and inductor LL to the ground. The inductance LL can be varied to modulate the resonance condition. To explore the effects of loss and gain in the TE network, we consider an imaginary onsite potential iRgiR_{g} on the AA nodes and iRg-iR_{g} on the BB nodes. This onsite potential can be realized by connecting resistors of resistance rar_{a} ( ra-r_{a}) between each AA (BB) node and the ground. The onsite potential are related to the resistors by Rg=1/ωraR_{g}=1/{\omega r_{a}}, where ω\omega denotes the frequency of the driving alternating current. The negative imaginary onsite potential at the BB-type nodes can be obtained by using op-amp-based negative resistance converters with current inversion (INRC) (see Supplemental section 1 for details) Alternatively, similar gain and loss terms can obtained in our TE model via using only two unequal positive resistances connected to ground for each type of nodes instead of using negative resistance converters, which yields the mathematically similar Hamiltonain as Eq. 1 except for a global shift of the imaginary part of all the admittance eigenenergies (see Supplementary Note 1 for details). The advantage of using grounding resistors with different values of (positive) resistance is the dynamic stability of the circuit because the absence of op-amps avoids the possibility that the voltage profiles may get over-amplified by the negative resistance.

Refer to caption
Figure 1: a) Schematic of the two-dimensional TE lattice in the xx-yy plane. The blue and magenta circles represent sublattices AA and BB respectively. The alternating sublattice sites AA and BB are connected to each other in the xx-direction by an inductor C1-C_{1} and capacitor C1C_{1} for the intracell and intercell connections, respectively (the dashed rectangle delineates a unit cell). Along the yy-direction, nodes of different sublattices are connected diagonally by a capacitor CyC_{y}. b) Schematic circuit of a negative resistance converter, which introduces a π\pi-phase difference and therefore converts a loss resistive term rar_{a} to a gain term ra-r_{a}. The combination of two resistors having the same resistance R1R_{1} along with an ideal operational amplifier with supply voltages +Vdd+V_{dd} and Vdd-V_{dd} results in current inversion and hence acts as a negative resistance converter. c) Grounding mechanism of the TE circuit. All nodes are connected to ground via a parallel combination of a common capacitor (CC) and inductor (LL). Furthermore, in parallel to these, a positive resistor rar_{a} (loss term) and negative resistor ra-r_{a} (gain term) is connected to ground from the AA and BB nodes, respectively. The negative resistor ra-r_{a} is implemented by means of INRC depicted in (b).

The Laplacian of the TE circuit over the xx-yy plane can be expressed as

H2D(kx,ky)=(C1(1+coskx)+2Cycosky)σxC1sinkxσy+iRgσz,\displaystyle H_{2D}({k_{x}},{k_{y}})=-(C_{1}(1+\cos{k_{x}})+2C_{y}\cos{k_{y}})\sigma_{x}-C_{1}\sin{k_{x}}\sigma_{y}+iR_{g}\sigma_{z}, (1)

where the σi\sigma_{i}s denote the Pauli matrices in the sublattice space. In the absence of resistances, Eq. 1 exhibits both the chiral symmetry, i.e., 𝒞H2D(kx,ky)𝒞1=H2D(kx,ky)\mathcal{C}H_{2D}(k_{x},k_{y})\mathcal{C}^{-1}=-H_{2D}(k_{x},k_{y}) and inversion symmetry, i.e., nH2D(kx,ky)n1=H2D(kx,ky)\mathcal{I}_{n}H_{2D}(k_{x},k_{y})\mathcal{I}_{n}^{-1}=H_{2D}(-k_{x},-k_{y}), where the chirality and inversion operators are 𝒞=σz\mathcal{C}=\sigma_{z} and n=σx\mathcal{I}_{n}=\sigma_{x}, respectively. However, a finite RgR_{g} would break the inversion symmetry of the Laplacian in Eq. 1 although the parity time (𝒫𝒯\mathcal{PT}) symmetry as defined by 𝒫𝒯H2D(kx,ky)𝒯1𝒫1=H2D(kx,ky)\mathcal{P}\mathcal{T}H_{2D}(k_{x},k_{y})\mathcal{T}^{-1}\mathcal{P}^{-1}=H_{2D}^{*}(k_{x},k_{y}), would still be preserved (here, 𝒫=σx\mathcal{P}=\sigma_{x} is the parity operator and 𝒯\mathcal{T} is complex conjugation). Therefore, the Laplacian in Eq. 1 can still have real eigenvalues despite its non-Hermiticity Lee (2016). More specifically, a non-zero RgR_{g} in Eq. 1 corresponds to the insertion of alternating iRgiR_{g} and iRg-iR_{g} terms on the diagonal of the Laplacian, which preserves the commutation with the 𝒫𝒯\mathcal{PT} operator Weimann et al. (2017). In contrast to the Hermitian case, 𝒫𝒯\mathcal{PT} symmetry in the non-Hermitian TE system eigenmodes can be broken depending on the model parameters, i.e., the eigenmodes of Eq. 1 are not necessarily the eigenstates of the 𝒫𝒯\mathcal{PT} operator Yuce and Oztas (2018b) even when the Laplacian itself respects 𝒫𝒯\mathcal{PT} symmetry. In this situation, complex admittance spectra emerge with exceptional points where both the hole- and particle-like admittance bands coalesce. Therefore, the complex admittance dispersion for the circuit model take the form of

E2D(kx,ky)\displaystyle E_{2D}(k_{x},k_{y}) =\displaystyle= 2C12(1+coskx)+4Cy2cos2ky+4C1Cycosky(1+coskx)Rg2,\displaystyle\sqrt{2C_{1}^{2}(1+\cos{k_{x}})+4C_{y}^{2}\cos^{2}{k_{y}}+4C_{1}C_{y}\cos{k_{y}}(1+\cos{k_{x}})-R_{g}^{2}}, (2)

where the ±\pm refers to the two admittance bands, respectively. By tuning the circuit parameters, we can obtain different numbers of real solutions for k\vec{k} for E2D=0E_{2D}=0 in Eq. (2), which translates into different number of exceptional points in the Brillouin zone (BZ). The exceptional points occur at 𝒘ex=(kx,ky)=(π,±arccos(Rg/2Cy))\boldsymbol{w}_{ex}=(k_{x},k_{y})=(\pi,\pm\arccos(R_{g}/2C_{y})) and 𝒘ex=(π,π±arccos(Rg/2Cy))\boldsymbol{w}_{ex}=(\pi,\mp\pi\pm\arccos(R_{g}/2C_{y})).

Refer to caption
Figure 2: Absolute value, real part, and imaginary part of the complex admittance as a function of kyk_{y} with the parameters C1=0.78C_{1}=0.78 mF\mathrm{mF}, Cy=0.39C_{y}=0.39 mF\mathrm{mF}, and kx=πk_{x}=\pi. We consider three representative values of the grounding resistance, i.e., (a) Rg=0R_{g}=0 mΩ1Hz1\mathrm{m\Omega^{-1}{Hz}^{-1}}, (b) Rg=0.5R_{g}=0.5 mΩ1Hz1\mathrm{m\Omega^{-1}{Hz}^{-1}}, and (c) Rg=0.78R_{g}=0.78 mΩ1Hz1\mathrm{m\Omega^{-1}{Hz}^{-1}}. Note that case (c) corresponds to the critical value of the non-Hermitian parameter Rg=2Cy=0.78R_{g}=2C_{y}=0.78 mΩ1Hz1\mathrm{m\Omega^{-1}{Hz}^{-1}}, beyond which the admittance spectrum becomes purely imaginary. All the exceptional points are represented by open red circles.

To illustrate the effect of non-Hermitian gain or loss, we plot the admittance spectrum as a function of wavevector kyk_{y} and fix kx=πk_{x}=\pi for three representative values of RgR_{g}. In the absence of gain or loss (i.e., Rg=0R_{g}=0), the admittance spectrum becomes purely real with two Dirac points or exceptional points (see Fig. 2a). Because the Laplacian in Eq. 1 obeys chiral symmetry, for a given kyk_{y}, the admittance eigenvalues always come in pairs with equal magnitude but opposite signs. For non-zero values of RgR_{g}, the admittance dispersion becomes complex and the band-touching degeneracy points split into pairs of band-touching exceptional points. For instance, in Fig. 2b, the splitting of the degeneracy points is most evident in the admittance plots in the left-most and middle columns. Here, the degeneracy point with zero admittance at ky1.6k_{y}\approx-1.6 in Fig. 2a splits into two exceptional points (EPs) at ky2.3k_{y}\approx-2.3 and ky0.9k_{y}\approx-0.9 in Fig. 2b. The admittance spectrum is then either real (for some range of kyk_{y}) with 𝒫𝒯\mathcal{PT}-symmetrical eigenmodes or purely imaginary (in the complementary range of kyk_{y}), in which case the eigenmodes break the 𝒫𝒯\mathcal{PT} symmetry. The boundaries between the real and imaginary admittances are defined by the EPs, where all the eigenmodes coalesce at the eigenvalue of zero. Two of the four EPs are located at ky=π±arccos(Rg/2Cy)k_{y}=\mp\pi\pm\arccos(R_{g}/2C_{y}) while the other two are at ky=±arccos(Rg/2Cy)k_{y}=\pm\arccos(R_{g}/2C_{y}). As the magnitude of RgR_{g} increases, the range of kyk_{y} corresponding to the real (imaginary) part of the admittance spectrum shrinks (expands). At some critical value given by Rc=2CyR_{c}=2C_{y}, the whole spectrum becomes purely imaginary with a thre EPs (see Fig. 2c). When RgR_{g} exceeds RcR_{c}, the two admittance bands will become gapped and no EP exists in the Brillouin zone (not shown in Fig. 2). In this case, the admittance eigenmodes break the 𝒫𝒯\mathcal{PT} symmetry for the entire range of kyk_{y} in the BZ. As can be seen from Fig. 2c, the real part of the admittance spectrum vanishes at the critical resistance RcR_{c}. In summary, we can obtain a variable number of EPs depending on the RgR_{g} parameter, i.e., two, four, three, and zero EPs for Rg=0R_{g}=0, Rg<2CyR_{g}<2C_{y}, Rg=2CyR_{g}=2C_{y}, and Rg>2CyR_{g}>2C_{y}, respectively.

To obtain the exceptional lines (the loci of the exceptional points), we use the equation for the degeneracy points of the admittance spectrum:

(C1(1+coskx)+2Cycosky)2+(C1sinkx)2=Rg2.\left(C_{1}(1+\cos{k_{x}})+2C_{y}\cos{k_{y}}\right)^{2}+(C_{1}\sin{k_{x}})^{2}=R_{g}^{2}. (3)

Eq. 3 governs the loci of the exceptional points in the kxk_{x}-kyk_{y} plane. A finite RgR_{g} will transform a single pair of band-touching points into exceptional or nodal lines on the kz=0k_{z}=0 plane characterized by Eq. 3. On these exceptional lines, both the real and imaginary parts of the eigenvalues vanish. The top panels of Fig. 3a and b show the exceptional lines for the parameter sets in Fig. 2b and c , respectively. The exceptional points shown in Fig. 2b and c then correspond to the kyk_{y} cross sections of the exceptional lines in Fig. 3 at kx=πk_{x}=\pi.

Refer to caption
Figure 3: The top panels show the loci of the exceptional points in 3D non-Hermitian systems, i.e., systems with finite resistive couplings, at C1=0.78C_{1}=0.78 mF\mathrm{mF}, Cy=0.39C_{y}=0.39 mF\mathrm{mF}, kz=π/2k_{z}=\pi/2, and a. Rg=0.5mFR_{g}=0.5\leavevmode\nobreak\ \mathrm{mF} , and b. Rg=0.78mFR_{g}=0.78\ \mathrm{mF}. The lower panels show the corresponding impedance spectra for the corresponding three-dimensional systems with the same values of C1,CyC_{1},C_{y} and RgR_{g} as the two-dimensional systems. (Note that the quantities plotted are the base 10 logarithms of the impedances in Ohms.)

In addition to the above analysis of the exceptional lines based on the admittance spectrum, an alternative visual representation of the exceptional lines can also be obtained from the impedance spectrum of the TE circuit. In general, the impedance between any two arbitrary nodes pp and qq in the circuit can be measured by connecting an external current source providing a fixed current IpqI_{pq} to the two nodes and measuring the resulting voltages at the two nodes VpV_{p} and VqV_{q}. The impedance in the circuit is then given by

Zpq=VpVqIpq=i|ψi,pψi,q|2λi,Z_{pq}=\frac{V_{p}-V_{q}}{I_{pq}}=\sum_{i}\frac{|\psi_{i,p}-\psi_{i,q}|^{2}}{\lambda_{i}}, (4)

where ψi,a\psi_{i,a} and λi\lambda_{i} are the voltage at node aa and the eigenvalue of the ithi^{\mathrm{th}} eigenmode of the (finite-width) Laplacian. One of the key characteristics of Eq. 4 is that the impedance diverges (increases to a large value) in the vicinity of the zero-admittance modes (λi=0\lambda_{i}=0 ) for non-zero eigen-mode voltages ψi,p\psi_{i,p} and ψi,q\psi_{i,q}. Therefore, the locus of the high impedance readout can be used to mark out the exceptional lines in momentum space. The lower panels of Fig. 3a and b depict the corresponding impedance spectra for the parameter sets in Fig. 2b and c respectively. The plotted impedance is that across the two nodes in a unit cell (i.e. with nodes pp and qq chosen to be terminal points at either end of the circuit). (For the case of C1=2Cy=RgC_{1}=2C_{y}=R_{g} plotted in Fig. 3b, the ky=±πk_{y}=\pm\pi lines are also exceptional lines.) For both resistive values, the locus of high impedance readouts coincides exactly with the exceptional lines in the admittance spectrum (compare upper and lower panels of Fig. 3). This suggests the possible electrical detection of exceptional lines in the TE system via impedance measurements.

.3 Zero-admittance states in finite system

To gain further insight into the zero-admittance states of a non-Hermitian system, we will study a 2D dissipative TE system described by Eq. 1 , which is finite along the xx-direction, i.e., having open boundary conditions along that direction, but is infinite in the yy-direction. Before investigating the properties of the finite system, we will first explain some properties of the infinite-sized 2D system. For this system, Kirchoff’s current law at the AA and BB nodes at resonance can be written as

EVx,yA=C1Vx,yB+C1Vx1,yBCy(Vx,y+1B+Vx,y1B)+iRgVx,yA,-EV_{x,y}^{A}=-C_{1}V_{x,y}^{B}+C_{1}V_{x-1,y}^{B}-C_{y}(V_{x,y+1}^{B}+V_{x,y-1}^{B})+iR_{g}V_{x,y}^{A}, (5)

and

EVx,yB=C1Vx,yA+C1Vx+1,yACy(Vx,y+1A+Vx,y1A)iRgVx,yB.-EV_{x,y}^{B}=-C_{1}V_{x,y}^{A}+C_{1}V_{x+1,y}^{A}-C_{y}(V_{x,y+1}^{A}+V_{x,y-1}^{A})-iR_{g}V_{x,y}^{B}. (6)

By substituting the ansatz Vx,y=λeikx+ikyV_{x,y}=\lambda e^{ik_{x}+ik_{y}} in Eqs. 5 and 6, we obtain

(EiRgσz)|λ=(C1(1+χx(σxiσy)+χx1(σx+iσy))+2C2coskyσx)|λ(E-iR_{g}\sigma_{z})|\lambda\rangle=(C_{1}(1+\chi_{x}(\sigma_{x}-i\sigma_{y})+\chi_{x}^{-1}(\sigma_{x}+i\sigma_{y}))+2C_{2}\cos{k_{y}}\sigma_{x})|\lambda\rangle (7)

where χx=eikx\chi_{x}=e^{ik_{x}} and λ=(λA,λB)T\lambda=(\lambda_{A},\lambda_{B})^{\mathrm{T}}. For a given EE and kyk_{y}, χx\chi_{x} can be solved from Eq. 7 as

χx=(t2+C12p2)±Δ22tC1,\chi_{x}=\frac{-(t^{2}+C_{1}^{2}-p^{2})\pm\sqrt{\Delta^{2}}}{2tC_{1}}, (8)

where t=C1+2Cycoskyt=C_{1}+2C_{y}\cos{k_{y}} and p2=E2+Rg2p^{2}=E^{2}+R_{g}^{2}, and

Δ2(t2+C12p2)2(2tC1)2.\Delta^{2}\equiv(t^{2}+C_{1}^{2}-p^{2})^{2}-(2tC_{1})^{2}. (9)

When Δ2<0\Delta^{2}<0,

χx;(Δ2<0)=(t2+C12p2)±i|Δ2|2tC1.\chi_{x;(\Delta^{2}<0)}=\frac{-(t^{2}+C_{1}^{2}-p^{2})\pm i\sqrt{|\Delta^{2}|}}{2tC_{1}}. (10)

It can be readily seen that |χx;(Δ2<0)|=1|\chi_{x;(\Delta^{2}<0)}|=1, and because χxexp(ikx)\chi_{x}\equiv\exp(ik_{x}), this indicates that the corresponding values of kxk_{x} would be real when Δ2<0\Delta^{2}<0. In this case, kx=arg(χx;(Δ2<0))=±arctan((Δ2)/(t2+C12p2))k_{x}=\arg(\chi_{x;(\Delta^{2}<0)})=\pm\arctan(\sqrt{(-\Delta^{2})}/(t^{2}+C_{1}^{2}-p^{2})). On the other hand, when Δ2>0\Delta^{2}>0, χx;(Δ2>0)\chi_{x;(\Delta^{2}>0)} is real, and, in general, |χx;(Δ2>0)|1|\chi_{x;(\Delta^{2}>0)}|\neq 1. This indicates that the corresponding values of kxk_{x} would be imaginary. At the boundary between the two cases, we have

Δ2=0\displaystyle\Delta^{2}=0 \displaystyle\Rightarrow (t2+C12p2)2=(2tC1)2\displaystyle(t^{2}+C_{1}^{2}-p^{2})^{2}=(2tC_{1})^{2}
\displaystyle\Rightarrow χx;(Δ2=0)=(2tC1)22tC1=sign(2tC1),\displaystyle\chi_{x;(\Delta^{2}=0)}=-\frac{\sqrt{(2tC_{1})^{2}}}{2tC_{1}}=-\mathrm{sign}(2tC_{1}),

so that the corresponding kx=0,πk_{x}=0,\pi depending on the sign of 2tC12tC_{1}. When Δ2<0\Delta^{2}<0, the ±i|Δ2|\pm i\sqrt{|\Delta^{2}|} terms in Eq. 10 result in a finite separation along the kxk_{x} axis between two points on the same equal admittance contours (EACs) for a given kyk_{y} value (see Fig. 4). The two values of kxk_{x} meet when Δ2=0\Delta^{2}=0. Figure 4a depicts the case where χx=+1\chi_{x}=+1 when Δ2=0\Delta^{2}=0. Here, for a given kyk_{y}, kxk_{x} on the EAC becomes single-valued at kx=0k_{x}=0. Figure 4b depicts the case where χx=1\chi_{x}=-1 when Δ2=0\Delta^{2}=0, and here kxk_{x} becomes single-valued at kx=πk_{x}=\pi. For both cases, the values of kyk_{y} where Δ2=0\Delta^{2}=0 mark the boundaries for the existence of real kxk_{x}.

Refer to caption
Figure 4: The equal admittance contours (EACs) at E=0E=0 for (left) with parameters C1=0.78C_{1}=0.78 mF\mathrm{mF}, Cy=0.1C_{y}=0.1 mF\mathrm{mF}, and Rg=1.5R_{g}=1.5 mΩ1Hz1\mathrm{m\Omega^{-1}{Hz}^{-1}} , and (right) with parameters C1=0.39C_{1}=0.39 mF\mathrm{mF}, Cy=0.78C_{y}=0.78 mF\mathrm{mF}, Rg=0.0R_{g}=0.0 mΩ1Hz1\mathrm{m\Omega^{-1}{Hz}^{-1}} . The regions in the Brillouin zone where Δ2>0\Delta^{2}>0 and Δ2<0\Delta^{2}<0 are indicated.

We now consider the nanoribbon geometry with a finite width along the xx-direction. We show in the Supplementary Materials that for the nanoribbon geometry with non-zero |Rg||R_{g}|, the zero-admittance exceptional points do not occur within the bulk energy gaps, but in the bulk bands where real values of kxk_{x} exist for E=0E=0 in the infinite bulk system. Due to the finite width of the nanoribbons, these zero-admittance points occur as quantized bulk states. The points on the kyk_{y} axis that mark the threshold for the existence of the zero admittance states for the nanoribbon system coincide with the kyk_{y} values where the solutions of χx\chi_{x} at E=0E=0 in Eq. 8 are equal, i.e., when Δ2=0\Delta^{2}=0. These values of kyk_{y} would mark the ‘boundary points’ (BPs) of the system. The values of kyk_{y} corresponding to the BPs are given by the solutions of the following equation:

C1+2Cycosky+ηC1=ζRg,C_{1}+2C_{y}\cos{k_{y}}+\eta C_{1}=\zeta R_{g}, (11)

where η=±1\eta=\pm 1 and ζ=±1\zeta=\pm 1 independently. The four possible combinations of η\eta and ζ\zeta and the two possible signs of kyk_{y} in Eq. (11) therefore provide up to eight real solutions for kyk_{y}. The BPs come in pairs, and hence, depending on the choice of the coupling capacitances C1C_{1} and CyC_{y} and resistance RgR_{g}, we can obtain up to a maximum of four pairs of BPs (see later discussions on phase diagram). Note that the BPs set the boundaries for the existence of exceptional points. In a nanoribbon, the exceptional points do not necessarily appear exactly at the BPs, but in between alternating pairs of BPs. To illustrate the role of the capacitive and resistive coupling parameters in determining the number of BPs, we plot the admittance spectra for finite TE circuits with N=20N=20 unit cells along the xx-direction in Fig. 5. In the Hermitian limit (i.e. Rg=0R_{g}=0), the admittance spectrum is purely real, and each pair of quantized bands is symmetric about the E=0E=0 axis. The Hermitian TE circuit can host two or four BPs depending upon the relative strength of the capacitive couplings. If C1>CyC_{1}>C_{y}, we only have a pair of BPs occurring at ky=±π/2k_{y}=\pm\pi/2. However, an additional pair of BPs emerges at ky=arccos(±C1/Cy)k_{y}=\arccos(\pm C_{1}/C_{y}) in the spectrum if C1<CyC_{1}<C_{y} (which is the case illustrated in Fig. 5a). The Hermitian system possesses chiral and time-reversal symmetries and belongs to to the BDI Altland-Zimbauer class Kawabata et al. (2019). Treating kyk_{y} as a parameter to an effectively one-dimensional model, the effective 1D model along the xx direction is mathematically identical to the SSH model, for which the topological invariant is the winding number. The band-touching points B1 to B4 will then correspond to the values of kyk_{y} at which the bulk band gap closes and the system transits between topologically trivial phases and topologically non-trivial phases. The emergence of the non-trivial states in Fig. 5a) span over (π/2-\pi/2 to arccos(C1/Cy)-\arccos(C_{1}/C_{y})) and (π/2\pi/2 to arccos(C1/Cy)\arccos(C_{1}/C_{y})) in the kyk_{y} axis. (Although the bands appear to be flat at the scale of the figure, they actually disperse weakly and touch only at isolated points.) The transition points between non-trivial (with edge states) and trivial regions (without edge states) are marked by the onset of high impedance states (shown in the rightmost plot of Fig. 5a).

Refer to caption
Figure 5: Evolution of edge states and boundary points (BP) which mark the transition between non-zero and zero admittance surface states in a non-Hermitian TE circuit, as the resistive and capacitive couplings are varied. We consider a finite TE circuit with N=20N=\mathrm{20} unit cells along the xx-direction. The first and second columns depict the real and imaginary parts of the admittance spectra, while the third column plots the corresponding spectra for the impedance taken between two terminal points. Parameters used for Panel a: C1=0.39C_{1}=0.39 mF\mathrm{mF}, Cy=0.78C_{y}=0.78 mF\mathrm{mF}, and Rg=0R_{g}=0 mF\mathrm{mF}; Panel b: C1=0.78C_{1}=0.78 mF\mathrm{mF}, Cy=0.39C_{y}=0.39 mF\mathrm{mF}, and Rg=0.5R_{g}=0.5 mF\mathrm{mF}; and Panel c: C1=0.39C_{1}=0.39 mF\mathrm{mF}, Cy=0.78C_{y}=0.78 mF\mathrm{mF}, and Rg=0.5R_{g}=0.5 mF\mathrm{mF}. All the BPs are marked with open red circles in the admittance dispersion and impedance spectra. The kyk_{y} regions between the BPs are denoted as I, III, IIA and IIB regions for ease of reference in the text.

A finite non-Hermitian gain or loss term RgR_{g} results in four different types of admittance regions for a given value of kyk_{y}. These admittance regions can host purely real, purely imaginary, complex spectra with exceptional points, and complex admittances without E2D=0E_{2D}=0 states, which are labelled as I, III, IIA and IIB respectively, as shown in Figs. 5b and 5c. (For instance, the states between B2 and B3, and B6 and B7 in Fig. 5c have purely imaginary admittances.) Eigenstates with preserved and broken 𝑃𝑇\mathit{PT} symmetry are present in regions I and III. Additionally, region IIA hosts exceptional points while IIB does not (see Figs. 5b and 5c ). The difference between regions IIA and IIB can be explained via the existence and non-existence of real solutions of χx\chi_{x} in Eq. 8. The transition points between any two successive regions are marked by the BPs. Moreover, the zero-admittance states in region IIA and the E2D0E_{2D}\neq 0 states in the other regions are characterized by high and low impedances respectively (see impedance plots shown in the right-most column of Fig. 5b and c ). For illustration, the EPs in the region IIA are marked by crosses in Fig. 5b in both the impedance and admittance plots. The impedance readout switches between the high and low impedance states occur at the BPs, which mark the boundaries between the IIA and other regions.

Refer to caption
Figure 6: a) Phase diagram of a 2D non-Hermitian TE model showing the effect of the capacitive coupling C1C_{1} and the gain/loss parameter RgR_{g} on the number of BPs, and hence the number of zero-admittance regions in the admittance spectrum. We consider a finite number of unit cells along xx-direction, and set Cy=1.0C_{y}=1.0 mF\mathrm{mF}. b) and c): (Left) Dispersion relations of Nx=20N_{x}=20 unit cell-wide nanoribbons at b. C1=2C_{1}=2 mF and Rg=0R_{g}=0 mF and c. C1=1mFC_{1}=1\ \mathrm{mF} and Rg=1.5mFR_{g}=1.5\ \mathrm{mF} indicated by the two circles in panel a; (Right) The square of the voltage amplitudes corresponding to the EP with the minimum kyk_{y} value, as indicated by the circles in the dispersion relations.

Fig. 6a shows the phase diagram of the 2D non-Hermitian TE model as a function of the resistive coupling RgR_{g} and the capacitive coupling C1C_{1}. The different phases are characterized by different numbers of BPs and hence different numbers of zero-admittance regions. These phases are separated from one another by phase boundaries that are defined by the existence and number of real solutions in Eq. 11. Note that BPs always occur and annihilate in pairs and they connect two admittance bands together. To study the effect of non-Hermitian term RgR_{g} on the zero-admittance states, we consider a 2D TE system with open-boundary conditions in the xx-direction. In the Hermitian condition, i.e., Rg=0R_{g}=0, the system hosts zero-energy modes in which the square of the voltage amplitudes are localized at its edges, as shown in top panel of Fig. 6b. This corresponds to the usual non-trivial SSH edge state. However, in the non-Hermitian case, i.e., with the introduction of a finite RgR_{g}, we would instead obtain bulk states at E=0E=0 rather than edge states (see Fig. 6c). This is indicated by the square of the voltage amplitude being no longer localized at an edge.

.4 Conclusion

In conclusion, we proposed a topolectrical (TE) circuit model with resistive elements which provide loss and gain factors that break the Hermiticity of the circuit to model and realize various non-Hermitian topological phases. By varying the resistive elements, the loci of the exceptional points (or exceptional lines) of the circuit can be modulated. IWe showed that the topology of the exceptional lines in the Brillouin zone can be traced by the impedance spectra of the circuit. Additionally, we studied a finite TE system where open boundary conditions apply in one of the dimensions. In these finite circuits, we demonstrated the tunability of both the number of exceptional points corresponding to zero-admittance states, as well as that of boundary points (BPs) which delineate the circuit parameter range where these exceptional points exist. The regions separated by the BPs are characterized by high and low values of impedance differing by several orders of magnitude, which are detectable in a practical circuit. We also derived a phase diagram of the finite TE system which delineates different between topological phases that are characterized by different number of BP pairs (up to a maximum of four). The edge state character of zero-admittance states of Hermitian LC circuits are transformed into exceptional points which are hybridized with the bulk modes in the non-Hermitian RLC circuits. In summary, we have proposed a tunable electrical framework consisting of RLC circuit networks as a means to realize different topological phases of non-Hermitian systems, and characterize them based on their impedance output, as well as their BP and exceptional line configurations.

Data availability

The data that support the plots within this paper and other findings of this study are available from the corresponding author upon reasonable request.

Code availability

The computer codes used in the current study are accessible from the corresponding author upon reasonable request.

Acknowledgement

This work is supported by the Ministry of Education (MOE) Tier-II grant MOE2018-T2-2-117 (NUS Grant Nos. R-263-000-E45-112/R-398-000-092-112), MOE Tier-I FRC grant (NUS Grant No. R-263-000-D66-114), and other MOE grants (NUS Grant Nos. C-261-000-207-532, and C-261-000-777-532).

Author contributions

S.M.R-U-I, Z.B.S and M.B.A.J initiated the primary idea. S.M.R-U-I and Z.B.S contributed to formulating the analytical model and developing the code, under the kind supervision of M.B.A.J. All the authors contributed to the data analysis and the writing of the manuscript.

Additional information

Competing interests: The authors declare no competing interests.

References

  • Fu and Kane (2007) L. Fu and C. L. Kane, Phys. Rev. B 76, 045302 (2007).
  • Moore and Balents (2007) J. E. Moore and L. Balents, Phys. Rev. B 75, 121306 (2007).
  • Liu et al. (2013) X.-J. Liu, K.-T. Law, T.-K. Ng,  and P. A. Lee, Phys. Rev. Lett. 111, 120402 (2013).
  • Zhang et al. (2018) D.-W. Zhang, Y.-Q. Zhu, Y. Zhao, H. Yan,  and S.-L. Zhu, Adv. Phys. 67, 253 (2018).
  • Lu et al. (2014) L. Lu, J. D. Joannopoulos,  and M. Soljačić, Nat. Photonics 8, 821 (2014).
  • Khanikaev et al. (2013) A. B. Khanikaev, S. H. Mousavi, W.-K. Tse, M. Kargarian, A. H. MacDonald,  and G. Shvets, Nat. Mater. 12, 233 (2013).
  • Sato and Ando (2017) M. Sato and Y. Ando, Rep. Prog. Phys. 80, 076501 (2017).
  • Goldman et al. (2016) N. Goldman, J. C. Budich,  and P. Zoller, Nat. Phys. 12, 639 (2016).
  • Lang et al. (2012) L.-J. Lang, X. Cai,  and S. Chen, Phys. Rev. Lett. 108, 220401 (2012).
  • Rafi-Ul-Islam et al. (2020a) S. Rafi-Ul-Islam, Z. B. Siu,  and M. B. Jalil, Appl. Phys. Lett. 116, 111904 (2020a).
  • Rafi-Ul-Islam et al. (2020b) S. Rafi-Ul-Islam, Z. B. Siu, C. Sun,  and M. B. Jalil, Phys. Rev. Appl. 14, 034007 (2020b).
  • Roy (2009) R. Roy, Phys. Rev. B 79, 195322 (2009).
  • Alvarez et al. (2018a) V. M. Alvarez, J. B. Vargas,  and L. F. Torres, Phys. Rev. B 97, 121401 (2018a).
  • Yao and Wang (2018) S. Yao and Z. Wang, Phys. Rev. Lett. 121, 086803 (2018).
  • Lee and Thomale (2019) C. H. Lee and R. Thomale, Phys. Rev. B 99, 201103 (2019).
  • Rafi-Ul-Islam et al. (2021) S. Rafi-Ul-Islam, Z. B. Siu,  and M. B. Jalil, Phys. Rev. B 103, 035420 (2021).
  • Kunst et al. (2018) F. K. Kunst, E. Edvardsson, J. C. Budich,  and E. J. Bergholtz, Phys. Rev. Lett. 121, 026808 (2018).
  • El-Ganainy et al. (2018) R. El-Ganainy, K. G. Makris, M. Khajavikhan, Z. H. Musslimani, S. Rotter,  and D. N. Christodoulides, Nat. Phys. 14, 11 (2018).
  • Jin and Song (2019) L. Jin and Z. Song, Phys. Rev. B 99, 081103 (2019).
  • Liu et al. (2019) T. Liu, Y.-R. Zhang, Q. Ai, Z. Gong, K. Kawabata, M. Ueda,  and F. Nori, Phys. Rev. Lett. 122, 076801 (2019).
  • Lin et al. (2011) Z. Lin, H. Ramezani, T. Eichelkraut, T. Kottos, H. Cao,  and D. N. Christodoulides, Phys. Rev. Lett. 106, 213901 (2011).
  • Feng et al. (2013) L. Feng, Y.-L. Xu, W. S. Fegadolli, M.-H. Lu, J. E. Oliveira, V. R. Almeida, Y.-F. Chen,  and A. Scherer, Nat. Mater. 12, 108 (2013).
  • Zhu et al. (2018a) W. Zhu, X. Fang, D. Li, Y. Sun, Y. Li, Y. Jing,  and H. Chen, Phys. Rev. Lett. 121, 124501 (2018a).
  • Liu et al. (2016) Z.-P. Liu, J. Zhang, Ş. K. Özdemir, B. Peng, H. Jing, X.-Y. Lü, C.-W. Li, L. Yang, F. Nori,  and Y.-x. Liu, Phys. Rev. Lett. 117, 110802 (2016).
  • Alvarez et al. (2018b) V. M. Alvarez, J. B. Vargas, M. Berdakin,  and L. F. Torres, Eur. Phys. J. Spec. Top. 227, 1295 (2018b).
  • Achilleos et al. (2017) V. Achilleos, G. Theocharis, O. Richoux,  and V. Pagneux, Phys. Rev. B 95, 144303 (2017).
  • Zhang et al. (2016) Z. Zhang, Y. Zhang, J. Sheng, L. Yang, M.-A. Miri, D. N. Christodoulides, B. He, Y. Zhang,  and M. Xiao, Phys. Rev. Lett. 117, 123601 (2016).
  • Jin (2017) L. Jin, Phys. Rev. A 96, 032103 (2017).
  • Yuce and Oztas (2018a) C. Yuce and Z. Oztas, Sci. Rep. 8, 1 (2018a).
  • Rafi-Ul-Islam et al. (2020c) S. Rafi-Ul-Islam, Z. B. Siu, C. Sun,  and M. B. Jalil, New J. Phys. 22, 023025 (2020c).
  • Lee et al. (2018) C. H. Lee, S. Imhof, C. Berger, F. Bayer, J. Brehm, L. W. Molenkamp, T. Kiessling,  and R. Thomale, Commun. Phys. 1, 1 (2018).
  • Imhof et al. (2018) S. Imhof, C. Berger, F. Bayer, J. Brehm, L. W. Molenkamp, T. Kiessling, F. Schindler, C. H. Lee, M. Greiter, T. Neupert, et al., Nat. Phys. 14, 925 (2018).
  • Helbig et al. (2019) T. Helbig, T. Hofmann, C. H. Lee, R. Thomale, S. Imhof, L. W. Molenkamp,  and T. Kiessling, Phys. Rev. B 99, 161114 (2019).
  • Rafi-Ul-Islam et al. (2020d) S. Rafi-Ul-Islam, Z. B. Siu,  and M. B. Jalil, Commun. Phys. 3, 1 (2020d).
  • Haenel et al. (2019) R. Haenel, T. Branch,  and M. Franz, Phys. Rev. B 99, 235110 (2019).
  • Zhu et al. (2019) W. Zhu, Y. Long, H. Chen,  and J. Ren, Phys. Rev. B 99, 115410 (2019).
  • Sun et al. (2020) C. Sun, S. Rafi-Ul-Islam, H. Yang,  and M. B. Jalil, Phys. Rev. B 102, 214419 (2020).
  • Zhu et al. (2018b) W. Zhu, S. Hou, Y. Long, H. Chen,  and J. Ren, Phys. Rev. B 97, 075310 (2018b).
  • Sun et al. (2019) C. Sun, J. Deng, S. Rafi-Ul-Islam, G. Liang, H. Yang,  and M. B. Jalil, Phys. Rev. Appl. 12, 034022 (2019).
  • Ezawa (2018) M. Ezawa, Phys. Rev. B 98, 201402 (2018).
  • Peterson et al. (2018) C. W. Peterson, W. A. Benalcazar, T. L. Hughes,  and G. Bahl, Nature 555, 346 (2018).
  • Serra-Garcia et al. (2019) M. Serra-Garcia, R. Süsstrunk,  and S. D. Huber, Phys. Rev. B 99, 020304 (2019).
  • Hofmann et al. (2019) T. Hofmann, T. Helbig, C. H. Lee, M. Greiter,  and R. Thomale, Phys. Rev. Lett. 122, 247702 (2019).
  • Rui et al. (2019) W. Rui, M. M. Hirschmann,  and A. P. Schnyder, Phys. Rev. B 100, 245116 (2019).
  • Choi et al. (2019) Y. Choi, C. Hahn, J. W. Yoon,  and S. H. Song, Nat. Commun. 10, 1 (2019).
  • Stegmaier et al. (2020) A. Stegmaier, S. Imhof, T. Helbig, T. Hofmann, C. H. Lee, M. Kremer, A. Fritzsche, T. Feichtner, S. Klembt, S. Höfling, et al., arXiv:2011.14836  (2020).
  • Lee (2016) T. E. Lee, Phys. Rev. Lett. 116, 133903 (2016).
  • Weimann et al. (2017) S. Weimann, M. Kremer, Y. Plotnik, Y. Lumer, S. Nolte, K. G. Makris, M. Segev, M. C. Rechtsman,  and A. Szameit, Nat. Mater. 16, 433 (2017).
  • Yuce and Oztas (2018b) C. Yuce and Z. Oztas, Sci. Rep. 8, 17416 (2018b).
  • Kawabata et al. (2019) K. Kawabata, K. Shiozaki, M. Ueda,  and M. Sato, Phys. Rev. X 9, 041015 (2019).