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

Duality between coherent quantum phase slip and Josephson junction in a nanosheet determined by the dual Hamiltonian method

M. Yoneda Aichi University of Technology, 50-2 Umanori Nishihasama-cho, Aichi 443-0047, Japan    M. Niwa    M. Motohashi School of Engineering, Tokyo Denki University, 5 Senju Asahi-cho, Adachi-ku Tokyo
120-855, Japan
Abstract

The duality between coherent quantum phase slip and Josephson junction in nanosheets was investigated using the dual Hamiltonian method. This is equivalent to the duality between superconductivity and superinsulator in the 2 + 1 dimension at zero temperature. This method proved to be reliable within the Villain approximation. The possibility of the dual Ginzburg-Landau theory, which is the phenomenology of superinsulators and the dual BCS theory, which is a microscopic theory, is also shown.

pacs:
74.20.z

1 Introduction

In recent decades, two types of phenomena, which are considered to be dual states in superconductivity, have attracted attention. One of them is a phenomenon called coherent quantum phase slip, which is mainly known as a dual phenomenon of a Josephson junction (JJJ\!J) in a one-dimensional superconducting system. The other is a phenomenon called superinsulator, which is mainly known as a dual phenomenon to superconductors (SC). A theoretical study of coherent quantum phase slip was submitted by Mooij et al.Mooij et al. (2005)-Mooij (2015) As flux quanta through nano superconducting wires, and an experimental demonstration of coherent quantum phase slip was performed by Astafiev et al.Astafiev (2012) which embedded in a large superconducting loop with InOx Achieved in wire. The concept of superinsulator was first conceived in 1978 by ’t HooftHooft (1978) as a theory explaining quark confinement. The superinsulator in the case of condensed matter was rediscovered by Diamantini et al.Diamantini et al. (1996)in 1996 as a periodic mixed Chern-Simons Abelian gauge theory describing charge-vortex coupling equivalent to planar Josephson junction arrays in a self-dual approximation. Then, in 1998, Krämer and Doniach Doniach et al. (1998) also rediscovered superinsulators using a phenomenological model in which vortices had a finite mass and traveled in a dissipative environment. Furthermore, in 2012, Yoneda et al. Yoneda et al. (2012) rediscovered superinsulators from a superinsulator / superconductor / superinsulator junction that is as the dual junction to Josephson junction. In previous studies, TiNVinokur et al. (2008), InOOvadia et al. (2015) and NbTiNMironov et al. (2018) films have been experimentally observed to be superinsulating materials. Recently, superinsulators have attracted attention as a powerful desktop environment for QCD phenomenaDiamantini et al. (2018, 2019) in order to realize a single-color version of quantum chromodynamics and elucidate quark confinement and asymptotic freedom. In a previous paperYoneda et al (2019), we introduced two Hamiltonians dual to each other for a one-dimensional nanowire-based Josephson junction and a quantum phase slip junction (QPSJQ\!P\!S\!J), and applied dual conditions between the current and voltage of the electric circuit. A general theory to construct a dual system, called the dual Hamiltonian method, was proposed. Furthermore, a general discussion of the superconductor-superinsulator transition in a 1+11+1 dimensional (1+1d1+1d) system at zero temperature from these two Hamiltonians was conducted. Our results show that in the 1+1d1+1d system at zero temperature, coherent quantum phase slip and superinsulator are completely equivalent phenomena. The main purpose of this work was to extend the theory of 1+1d1+1d systems on nanowires shown in the previous paper to the theory of 2+1d2+1d systems on nanosheets. The rest of this paper is organized as follows: In Section 2, we use the dual Hamiltonian method for between the JJ model and the QPSJQ\!P\!S\!J model on a nanosheet at zero temperature to derive a between the phase and the amplitude dual relations, and a dual relations between the various constants. In Section 3, based on the nonlinear Legendre transformation between the Lagrangians and the Hamiltonians with canonical conjugate variables of infinite order in a compact lattice space, and between the phase and the amplitude relationship derived in the previous section, we show that there is exact duality between the JJJ\!J model and the QPSJQ\!P\!S\!J model. Additionally, the phase diagram between the JJJ\!J state (superconducting state) and the QPSQ\!P\!S state (superinsulating state) was discussed. In Section 4, we prove the validity of the results of the previous section by deriving a dual transformation from the anisotropic XY(AXY)X\!Y(A\!X\!Y) model to the gauged dual anisotropic XY(DAXY)X\!Y(D\!A\!X\!Y) model by the Villain approximation in the 2+1d2+1d system. In Section 5, contrary to Section 4, we derive a dual transformation from the DAXYD\!A\!X\!Y model to the AXYA\!X\!Y model with gauge by the Villain approximation in a 2+1d2+1d system. In Section 6, the dual Ginzburg-Landau (DGLD\!G\!L) theory was derived from the mean field approximation of the gauged QPSJQ\!P\!S\!J model, and the possibility of confinement of electric flux in the superinsulator was shown. In Section 7, we calculated the critical value of the QPSQ\!P\!S amplitude by the effective energy approach. In the summary and discussion in Section 8, we summarize the conclusions of this paper, discuss whether the minimum unit of charge confinement in a superinsulator is 2e2e or ee, and, finally, we describe the possibility of dual BCS theory. In Appendix A, the numerical evaluation of the anisotropic massless lattice Green’s function is shown at the origin of x=0x\!=\!0, which is necessary for loop correction. In Appendix B, the effective energy approach for the QPSJQ\!P\!S\!J model is explained.

2 Dual Hamiltonian method between the Josephson junction (JJ) model and the quantum phase slip junction(QPSJ) model on a nanosheet at zero temperature

Refer to caption
Figure 1: A checkered nanosheet consisting of superconductors (black squares) and superinsulators (white squares).

Consider a checkered nanosheet consisting of superconductors and superinsulators as shown in Figure 1. In such a 2d2d checkered nanosheet, when the superconductor region acts strongly, the JJJ\!J state becomes strong, and the Hamiltonian in this case is as follows:

HJJ=Ec𝐱=1M[Nθ(x)]2+EJ𝐱=1Mj=12[1cosjθ(x)],\displaystyle\scalebox{0.85}{$\displaystyle{H_{\!J\!J}}={E_{c}}\!\sum\limits_{\mathbf{x}=1}^{M}{\left[N_{\theta}\left(x\right)\right]}^{2}+{E\!_{J}}\!\sum\limits_{\mathbf{x}=1}^{M}\sum\limits_{j=1}^{2}{\left[1-\cos{{\nabla\!}_{j}}\theta\left(x\right)\right]}$}, (1)

where, Nθ(x)Nθ(𝐱,τ){N_{\theta}}\!\left(x\right)\!\equiv\!{N_{\theta}}\!\left(\mathbf{x},\tau\right) means the number of the particles in the Cooper pair that is the canonical conjugate to the phase θ\theta of the Cooper pair, and the space differenceYoneda et al (2019); Diamantini et al (2017) of the phase θ(x)θ(𝐱,τ)\theta\left(x\right)\!\equiv\!\theta\left(\mathbf{x},\tau\right) is defined by jθ(𝐱,τ)θ(𝐱,τ)θ(𝐱a𝐞,τ){{\nabla}_{j}}\theta\!\left(\mathbf{x},\tau\right)\!\equiv\!\theta\!\left(\mathbf{x},\tau\right)\!-\!\theta\left(\mathbf{x}\!-\!a\mathbf{e},\tau\right) ; τβ\tau\!\equiv\hbar\beta ( β1/kBT\beta\!\equiv\!1/{{k_{B}}T}\; is the reverse temperature), 𝐱=(x1,x2)\mathbf{x}\!=\!\left(x_{1},x_{2}\right) and Ml2/a2M\!\equiv\!{l^{2}}/{a^{2}} (ll and aaare the space length and lattice spacing, respectively)are the imaginary time, the lattice coordinate points, and the total number of lattices in the two-dimensional lattice space, respectively; and Ec(2e)2/2C{E_{c}}\!\equiv\!{{\left(2e\right)}^{2}}/{2C} and EJΦ0Ic/2π{E\!_{J}}\!\equiv\!{{\Phi_{0}}{I_{c}}}/{2\pi} are the charging energy per Cooper pair and the Josephson energy, respectively, where CC,Ic{I_{c}} and Φ0=h/2e{\Phi_{0}}\!=\!h/2e are the capacitance, the critical current and the magnetic flux-quantum, respectively. From the Hamiltonian of Eq. (2), Josephson’s equations are as follows:

V(t)=i2eθ(x)τ=2Nθ(x)2eEc,\displaystyle\scalebox{0.9}{$\displaystyle V\left(t\right)=i\frac{\hbar}{2e}\frac{\partial\theta\left(x\right)}{\partial\tau}=\frac{2{N_{\theta}}\left(x\right)}{2e}{E_{c}}$},
(2)

where, V(x)V\left(x\right) and Ij(x){I_{j}}\left(x\right) (j=1,2j\!=\!1,2) are the voltage and the jj components of the current in the JJJ\!J, respectively. On the other hand, when the superinsulator region acts strongly, the QPSJQ\!P\!S\!J state becomes strong, and the Hamiltonian in this case is as follows:

HQPS=EL𝐱=1M[N~θ~(x)]2+ESj=12𝐱=1M[1cosjθ~(x)],\displaystyle\scalebox{0.85}{$\displaystyle{H_{Q\!P\!S}}={E_{L}}\sum\limits_{\mathbf{x}=1}^{M}{{{\left[{\tilde{N}_{\tilde{\theta}}}\left(x\right)\right]}^{2}}}+{E_{S}}\sum\limits_{j=1}^{2}{\sum\limits_{\mathbf{x}=1}^{M}{\left[1-\cos{{\nabla}_{j}}\tilde{\theta}\left(x\right)\right]}}$}, (3)

where, N~θ~(x)N~θ~(𝐱,τ){{\tilde{N}}_{\tilde{\theta}}}\left(x\right)\equiv{\tilde{N}_{\tilde{\theta}}}\left(\mathbf{x},\tau\right) means the magnetic flux quantum numbers θ~(x)θ~(𝐱,τ)\tilde{\theta}\left(x\right)\equiv\tilde{\theta}\left(\mathbf{x},\tau\right) of the magnetic flux quantum field. ELΦ02/2L\,{E_{L}}\equiv{{{\Phi}_{0}}^{2}}/{2L}\; and ES2eVc/2π\,{E_{S}}\equiv{2e{V_{c}}}/{2\pi}\; are the inductive energy per magnetic flux quantum and the QPSQ\!P\!S amplitude, respectively, where Vc\,{V_{c}} is the critical voltage. From the Hamiltonian of Eq.(3), the dual Josephson equations are as follows:

V~(x)=iΦ0θ~(x)τ=2N~θ~(x)Φ0EL,\displaystyle\scalebox{0.9}{$\displaystyle\tilde{V}\left(x\right)=i\frac{\hbar}{{\Phi}_{0}}\frac{\partial\tilde{\theta}\left(x\right)}{\partial\tau}=\frac{2{\tilde{N}_{\tilde{\theta}}}\left(x\right)}{\Phi_{0}}{E_{L}}$},
(4)

where, V~(x)\tilde{V}\left(x\right) and I~j(x){\tilde{I}_{j}}\left(x\right) (j=1,2j\!=\!1,2) are the dual voltage and the j components of dual current in the QPSJQ\!P\!S\!J, respectively. As the first step of the dual Hamiltonian method, two dual conditionsYoneda et al. (2012); Yoneda et al (2019) between Eq.(2) and the dual equations of Eq.(2) are assumed as follows:

V(x)I~(x), I(x)V~(x),.\displaystyle\scalebox{0.9}{$\displaystyle V\left(x\right)\equiv\tilde{I}\left(x\right),\text{ }I\left(x\right)\equiv\tilde{V}\left(x\right),$}. (5)

where, II12+I22I\!\equiv\!\sqrt{{I_{1}}\!^{2}+{I_{2}}\!^{2}} and I~I~12+I~22\tilde{I}\!\equiv\!\sqrt{{\tilde{I}_{1}}^{2}+{\tilde{I}_{2}}^{2}} are the intensity of the current in the JJJ\!J and the intensity of the dual current in the QPSJQ\!P\!S\!J, respectively. When the condition of Eq.(5) is imposed between Eqs.(2) and (2), the following two relational expressions between the phase and the number of particles between dual systems are derived as shown below. One of them is the relationship between the phase θ\theta of the Cooper pair and the magnetic flux quantum numbers N~θ~{\tilde{N}_{\tilde{\theta}}}, and the other is the relationship between the phase θ~\tilde{\theta} of the magnetic flux quantum and the number Nθ{N_{\theta}} of the Cooper pair, as follows:

Nθ(x)=12πj=12sin2jθ~(x),\displaystyle\scalebox{0.85}{$\displaystyle{{N}_{\theta}}\left(x\right)=\frac{1}{2\pi}\sqrt{\sum\limits_{j=1}^{2}{{{\sin}^{2}}{\nabla_{j}}\tilde{\theta}\left(x\right)}}$},
N~θ~(x)=12πj=12sin2jθ(x),\displaystyle\scalebox{0.85}{$\displaystyle{{\tilde{N}}_{\tilde{\theta}}}\left(x\right)=\frac{1}{2\pi}\sqrt{\sum\limits_{j=1}^{2}{{{\sin}^{2}}{\nabla_{j}}\theta\left(x\right)}}$}, (6)

In Eq.(2), the leftmost equation shows that the Cooper pair number is proportional to the strength of the flux quantum current, and the next equation Is shown that the flux quantum number is proportional to the strength of the Cooper pair current. Since the Cooper pair current and the flux quantum current form a closed loop, the number of Cooper pairs and the flux quantum number can be considered as the winding number of the current loop formed from each dual current. If the relationships described in Eq.(2) are satisfied, the relationship between the QPSQ\!P\!S amplitude and the charging energy per single charge, and the relationship between the Josephson energy and the inductive energy per magnetic flux-quantum, are as follows:

ES=Ec2π2,EJ=EL2π2.\displaystyle\scalebox{0.95}{$\displaystyle{E_{S}}=\frac{E_{c}}{2{\pi}^{2}},\;\;\;{E_{J}}=\frac{E_{L}}{2{\pi}^{2}}$}. (7)

Furthermore, the inductance and capacitance are related to the critical current and the critical voltage, respectively, as follows:

L=Φ02πIc,C=2e2πVc,\displaystyle\scalebox{0.95}{$\displaystyle L=\frac{\Phi_{0}}{2\pi{I_{c}}},\;\;\;C=\frac{2e}{2\pi{V_{c}}}$}, (8)

The Lagrangians at zero temperature in Eqs.(1) and Eq.(2) are as follows:

(9)
LQPS=𝐱=1M{ES02[τθ~(x)]2+ESj=12[1cosjθ~(x)]},\displaystyle\scalebox{0.8}{$\displaystyle{L\!_{Q\!P\!S}}\!=\!-\!\sum\limits_{\mathbf{x}=1}^{M}\!{\left\{\frac{E\!_{S}^{0}}{2}{{\left[{{\nabla}_{\tau}}\tilde{\theta}\left(x\right)\right]}^{2}}\!\!+\!{E\!_{S}}\!\sum\limits_{j=1}^{2}{\left[1-\cos{{\nabla}_{j}}\tilde{\theta}\left(x\right)\right]}\right\}}$}, (10)

where EJ0E_{J}^{0} and ES0E_{S}^{0} can be considered as the imaginary time components of the JJJ\!J and the QPSJQ\!P\!S\!J, respectively, and are defined as follows:

(11)

where a0τmax/Mτ{a_{0}}\equiv{\tau}_{\max}/M_{\tau} is an imaginary time spacing in the time dimension; and τmax{\tau}_{\max} and MτM_{\tau} are an imaginary time length and an imaginary time division number, respectively. Eqs.(7) and (11) can be summarized as a relationship between the Josephson energy and the QPSJQ\!P\!S\!J energy as follows:

(12)

where EJEJa0/{{E^{\prime}}\!_{J}}\equiv{{E_{J}}{a_{0}}}/{\hbar}, EJ0EJ0a0/{E^{\prime}}\!_{J}^{0}\equiv{E_{J}^{0}{a_{0}}}/{\hbar}, ESESa0/{{E^{\prime}}\!_{S}}\equiv{{E_{S}}{a_{0}}}/{\hbar} and ES0ES0a0/{E^{\prime}}\!_{S}^{0}\equiv{E_{S}^{0}{a_{0}}}/{\hbar} represent the nondimension energies, respectively.The first terms of Eq.(9) and (10) are expressed in a quadratic form for the imaginary time difference of each phase, however, the rightmost terms in Eqs.(9) and (10) are expressed in a cosine form for the spatial difference of each phase. Here, the cosine form also introduces approximately to the leftmost terms of Eqs.(9) and (10) in consideration of the periodicity in the lattice space as follows:

LAXY=𝐱=1M{EJ0[1cosτθ(x)]+EJj=12[1cosjθ(x)]},\displaystyle\scalebox{0.75}{$\displaystyle{L\!_{A\!X\!Y}}\!=\!-\!\sum\limits_{\mathbf{x}=1}^{M}{\left\{\!E_{J}^{0}\Bigl{[}1-\cos{{\nabla}_{\tau}}\theta\left(x\right)\Bigr{]}\!+\!{E_{J}}\sum\limits_{j=1}^{2}{\Bigl{[}1-\cos{{\nabla}_{j}}\theta\left(x\right)\Bigr{]}}\right\}}$}, (13)
LDAXY=𝐱=1M{ES0[1cosτθ~(x)]+ESj=12[1cosxθ~(x)]}.\displaystyle\scalebox{0.75}{$\displaystyle{L_{D\!A\!X\!Y}}\!=\!-\!\sum\limits_{\mathbf{x}=1}^{M}{\left\{\!E_{S}^{0}\left[1-\cos{{\nabla}_{\tau}}\tilde{\theta}\left(x\right)\right]\!+\!{E_{S}}\sum\limits_{j=1}^{2}{\left[1-\cos{{\nabla}_{x}}\tilde{\theta}\left(x\right)\right]}\right\}}$}. (14)

where, LAXY{L_{A\!X\!Y}} and LDAXY{L_{D\!A\!X\!Y}} are equivalent to LJJ{L_{J\!J}} and LQPS{L_{Q\!P\!S}}, respectively, and the AXYA\!X\!Y and DAXYD\!A\!X\!Y model, respectively. From the Lagrangian in Eq.(13) the partition function of the JJJ\!J state (superconducting state) at zero temperature is as follows:

ZAXYexp{(EJ0+EJ)MMτ}ZAXY,\displaystyle\scalebox{0.8}{$\displaystyle Z\!_{A\!X\!Y}\equiv\exp\left\{-\Bigl{(}{E^{\prime}}\!_{J}^{0}+{E^{\prime}}\!_{J}\Bigr{)}\!M{M_{\tau}}\right\}{{Z^{\prime}}\!_{A\!X\!Y}}$},\quad\quad\;\;\;
ZAXYDθexpx,τ[EJ0cosτθ(x)+EJj=12cosjθ(x)].\displaystyle\scalebox{0.8}{$\displaystyle{Z^{\prime}}\!_{A\!X\!Y}\!\!\equiv\!\!\int\!\!{D\theta}\exp\!\sum\limits_{x,\tau}\!{\left[{E^{\prime}}_{J}^{0}\cos{\nabla\!_{\tau}}\theta\left(x\right)+{E^{\prime}}\!_{J}\sum\limits_{j=1}^{2}{\cos{\nabla\!_{j}}\theta\left(x\right)}\right]}$}. (15)

where xτ=1Mτ𝐱=1M\sum\limits_{x}\!\!\!\equiv\!\!\sum\limits_{\tau\!=\!1}^{M_{\tau}}\!{\sum\limits_{\mathbf{x}\!=\!1}^{M}} and Dθτ=1Mτ𝐱=1Mππdθ(𝐱,τ)2π\int\!\!{D\theta\!}\!\!\equiv\!\!\prod\limits_{\tau\!=\!1}^{M_{\tau}}{\!\prod\limits_{\mathbf{x}\!=\!1}^{M}{\!\!\int\limits_{-\pi}^{\pi}\!\!{\frac{d\theta\left(\mathbf{x},\tau\right)}{2\pi}}}} are sums and path integrals, respectively, in 2+1d2+1d lattice space x=(x,τ)x\!=\!(x,\tau). Similarly, from the Lagrangian in Eq.(14), the partition function of the QPSJQ\!P\!S\!J state (superinsulating state) at zero temperature is as follows:

ZDAXYexp{(ES0+ES)MτMx}ZDAXY,\displaystyle\scalebox{0.8}{$\displaystyle{Z_{D\!A\!X\!Y}}\equiv\exp\left\{-\!\left({E^{\prime}}_{S}^{0}+{E^{\prime}}_{S}\right)\!{M_{\tau}}{M_{x}}\right\}{{Z^{\prime}}\!_{D\!A\!X\!Y}}$},\quad\quad
ZDAXYDθ~expx[ES0cosτθ~(x)+ESj=12cosjθ~(x)].\displaystyle\scalebox{0.8}{$\displaystyle{{Z^{\prime}}\!_{D\!A\!X\!Y}}\!\!\equiv\!\!\int\!\!{D\tilde{\theta}}\exp\!\sum\limits_{x}\!{\left[{E^{\prime}}_{S}^{0}\cos{\nabla\!_{\tau}}\tilde{\theta}\left(x\right)+{E^{\prime}}\!_{S}\!\sum\limits_{j=1}^{2}{\cos{\nabla\!_{j}}\tilde{\theta}\left(x\right)}\right]}$}. (16)

The partition functions ZAXYZ_{A\!X\!Y} and ZDAXYZ_{D\!A\!X\!Y} in Eqs.(2) and (2) are the starting points for the analysis in the following sections.

3 Dual transformation between the JJ model and the QPSJ model on the nanosheet

This section describes the nonlinear Legendre transformation between the Hamiltonian and the Lagrangian with canonical conjugate variables of the infinite order in a compact 2+1d2+1d lattice space on a nanosheet. Herein, we show that there is exact duality between the JJJ\!J model and the QPSJQ\!P\!S\!J model. For the partition functions ZJJ{Z\!_{J\!J}} in the Lagrangian of Eq.(9), the auxiliary field N(x)N\left(x\right) by Hubbard-Stratonovich transformation is introduced as follows:

.EJj=12[1cosjθ(x)]},\displaystyle\scalebox{0.9}{$\displaystyle\Bigl{.}-\!{{E^{\prime}}\!_{J}}\!\sum\limits_{j=1}^{2}{\left[1\!-\!\cos{{\nabla}\!_{j}}\theta\left(x\right)\right]}\Bigr{\}}$},\quad (17)

It can be seen that the auxiliary field N(x)N\left(x\right) is the same as the number Nθ(x){N_{\theta}}\left(x\right) of the Cooper pairs introduced in Eq.(1).The canonical conjugate momentum pθ(x){{p}_{\theta}}\left(x\right) with respect to θ(x)\theta\left(x\right) in the Lagrangian of the lattice space of Eq.(9) is defined as follows:.

ipθ(x)iN(x)=a0EJ0τθ(x).\displaystyle\scalebox{0.98}{$\displaystyle i{p_{\theta}}\!\left(x\right)\!\equiv\!i\hbar{N}\!\left(x\right)\!=\!-{a_{0}}E_{J}^{0}{\nabla_{\tau}}\theta\left(x\right)$}. (18)

On the other hand, for the partition functions ZAXYZ_{A\!X\!Y} in Eq.(2), the auxiliary field N(x)N\!\left(x\right) by Hubbard–Stratonovich transformation is introduced as follows:

.EJj=12[1cosjθ(x)]},\displaystyle\scalebox{0.83}{$\displaystyle\Bigl{.}-{{E^{\prime}}\!_{J}}\sum\limits_{j=1}^{2}{\left[1\!-\!\cos{\nabla\!_{j}}\theta\!\left(x\right)\right]}\Bigr{\}}$},\quad (19)

Again, it can be seen that the auxiliary field N(x)N\left(x\right) is the same as the number Nθ(x){N_{\theta}}\left(x\right) of the Cooper pairs introduced in Eq.(1). The canonical conjugate momentum pθ(x){{p}_{\theta}}\!\left(x\right) with respect to θ(x)\theta\!\left(x\right) in the Lagrangian of the compact lattice space of Eq.(13) is defined as follows:

ipθ(x)iNθ(x)=2a0EJ0sin[τθ(x)/2],\displaystyle\scalebox{0.95}{$\displaystyle i{p_{\theta}}\left(x\right)\equiv i\hbar{N_{\theta}}\left(x\right)=-2{a_{0}}E_{J}^{0}\sin\Bigl{[}{\nabla_{\tau}\theta\left(x\right)}/2\Bigr{]}$}, (20)

In Eq.(20), if the linear approximation of sin(τθ/2)τθ/2\sin\!\left({\nabla\!_{\tau}\theta\!}/2\right)\!\!\simeq\!\!{{\nabla_{\tau}}\!\theta\!}/2 holds, it matches the canonical conjugate momentum in Eq.(18).Therefore, the contents of the curly bracket of the exponential function between Eqs.(2) and (3) can be considered as a “nonlinear Legendre transformation” introduced by the canonical conjugate momentum of Eq.(20). Similarly, for the partition function ZDAXYZ\!_{D\!A\!X\!Y} in Eq.(2), the auxiliary field N~(x)\tilde{N}\left(x\right) by Hubbard–Stratonovich transformation is introduced as follows:

.ESj=12[1cosjθ~(x)]},\displaystyle\scalebox{0.83}{$\displaystyle\Bigl{.}-{{E^{\prime}}\!_{S}}\sum\limits_{j=1}^{2}\!{\left[1\!-\!\cos{\nabla\!_{j}}\tilde{\theta}\!\left(x\right)\right]}\Bigr{\}}$},\quad (21)

It can be seen that the auxiliary field N~(x)\tilde{N}\left(x\right) is the same as the magnetic flux quantum numbers N~θ~(x)\tilde{N}_{\tilde{\theta}}\left(x\right) in Eq.(3). The canonical conjugate momentum p~θ~(x){\tilde{p}_{\tilde{\theta}}}\left(x\right) with respect to θ~(x)\tilde{\theta}\left(x\right) in the Lagrangian of the compact lattice space of Eq.(14) is defined as follows:

ip~θ~(x)iN~θ~(x)=2a0ES0sin[τθ~(x)/2],\displaystyle\scalebox{0.95}{$\displaystyle i{\tilde{p}_{\tilde{\theta}}}\left(x\right)\equiv i\hbar{\tilde{N}_{\tilde{\theta}}}\left(x\right)=-2{a_{0}}E_{S}^{0}\sin\!\left[{\nabla_{\tau}\tilde{\theta}\left(x\right)}/2\right]$}, (22)

Therefore, the contents of the curly brackets of the exponential function between Eqs.(2) and (3) can also be considered as the“nonlinear Legendre transformation” introduced by the canonical conjugate momentum of Eq.(22). Since Eqs.(20) and (22) have a half-angle relationship, a half-angle version of Eq.(2) is introduced as follows:

Nθ(x)=12πj=12sin2[jθ~(x)/2],\displaystyle\scalebox{0.82}{$\displaystyle{N_{\theta}}\left(x\right)=\frac{1}{2\pi}\sqrt{\sum\limits_{j=1}^{2}{{\sin}^{2}\left[{{\nabla_{j}}\tilde{\theta}\left(x\right)}/2\right]}}$},
N~θ~(x)=12πj=12sin2[jθ(x)/2],\displaystyle\scalebox{0.82}{$\displaystyle\tilde{N}_{\tilde{\theta}}\left(x\right)=\frac{1}{2\pi}\sqrt{\sum\limits_{j=1}^{2}{{\sin}^{2}\Bigl{[}{{\nabla_{j}}\theta\left(x\right)}/2\Bigr{]}}}$}, (23)

Eq.(3) is equivalent to Eq.(2) within the range of linear approximation. From Eqs.(20), (3), and (3), the relationship between the phase θ(x)\theta\left(x\right) of the Cooper pair and the phase θ~(x)\tilde{\theta}\left(x\right) of the magnetic flux quantum field is as follows:

Nθ2(x)=4(a0EJ0)22sin2[τθ(x)/2]=1π2j=12sin2[jθ~(x)/2],\displaystyle\scalebox{0.82}{$\displaystyle N_{\theta}^{2}\!\left(x\right)\!=\!-\frac{4{\left({a_{0}}E_{J}^{0}\right)}^{2}}{\hbar^{2}}{\sin^{2}}\Bigl{[}\nabla_{\tau}\theta\left(x\right)/2\Bigr{]}\!=\!\frac{1}{\pi^{2}}\!\sum\limits_{j=1}^{2}{{\sin^{2}}\left[{{\nabla}_{j}\tilde{\theta}\left(x\right)}/2\right]}$},
N~θ~2(x)=4(a0ES0)22sin2[τθ~(x)/2]=1π2j=12sin2[jθ(x)/2],\displaystyle\scalebox{0.82}{$\displaystyle\tilde{N}_{\tilde{\theta}}^{2}\!\left(x\right)\!=\!-\frac{4{\left({a_{0}}E_{S}^{0}\right)}^{2}}{\hbar^{2}}{\sin^{2}}\left[{\nabla_{\tau}\tilde{\theta}\left(x\right)}/2\right]\!=\!\frac{1}{\pi^{2}}\!\sum\limits_{j=1}^{2}{{\sin^{2}}\Bigl{[}{\nabla_{j}}\theta\left(x\right)/2\Bigr{]}}$}, (24)

Using Eqs.(7), (11) and (3), the following relationships are obtained:

Ec𝐱[Nθ(x)]2=EJ0𝐱[1cosτθ(x)]=ES𝐱j=12[1cosjθ~(x)],\displaystyle\scalebox{0.75}{$\displaystyle E_{c}\!\sum\limits_{\mathbf{x}}{{\left[N_{\theta}\left(x\right)\right]}^{2}}\!\!=\!\!-E_{J}^{0}\sum\limits_{\mathbf{x}}\!{\Bigl{[}1\!-\!\cos{\nabla_{\tau}}\theta\left(x\right)\Bigr{]}}\!=\!E_{S}\!\sum\limits_{\mathbf{x}}{\sum\limits_{j=1}^{2}{\left[1\!-\!\cos{\nabla_{j}}\tilde{\theta}\left(x\right)\right]}}$}, (25)
EL𝐱[N~θ~(x)]2=ES0𝐱[1cosτθ~(x)]=EJ𝐱j=12[1cosjθ(x)].\displaystyle\scalebox{0.75}{$\displaystyle E_{L}\!\sum\limits_{\mathbf{x}}{{\left[\tilde{N}_{\tilde{\theta}}\left(x\right)\right]}^{2}}\!\!\!\!=\!\!-E_{S}^{0}\sum\limits_{\mathbf{x}}\!{\left[1\!-\!\cos{\nabla_{\tau}}\tilde{\theta}\left(x\right)\right]}\!=\!E_{J}\sum\limits_{\mathbf{x}}{\sum\limits_{j=1}^{2}{\Bigl{[}1\!-\!\cos{{\nabla}_{j}}\theta\left(x\right)\Bigr{]}}}$}. (26)

Eq.(25) means that the charging energy by the charge 2e2e in the JJJ\!J is equal to the QPSJQ\!P\!S\!J energy, that is, the condensation energy of the magnetic flux. Eq.(26) means that the charging energy by the flux quantum Φ0\Phi_{0} in the QPSJQ\!P\!S\!J is equal to the JJJ\!J energy, that is, the condensation energy of the Cooper pair. By establishing the relationship between Eqs.(25) and (26), between the Hamiltonian Eqs.(1) and (3), and, between the Lagrangian Eqs.(13) and (14), it can be shown that each of them is self-dual. Thus, from the canonical conjugate momentum of Eqs.(20) and (22), and the relationship between the phase and amplitude of Eq.(2) derived in the previous section, an exact duality between the JJJ\!J and QPSJQ\!P\!S\!J models has been proven. From the Josephson’s equations Eq.(2) and the dual Josephson’s equations Eq.(2), the electrical conductance GG can be derived as followsYoneda et al. (2012); Yoneda et al (2019):

G=GQdNθ(x)dN~θ~(x)=GQEJESN~θ~(x)Nθ(x),\displaystyle\scalebox{0.85}{$\displaystyle G=-{G_{Q}}\frac{d{N_{\theta}}\left(x\right)}{d{\tilde{N}_{\tilde{\theta}}}\left(x\right)}={G_{Q}}\frac{E_{J}}{E_{S}}\frac{{\tilde{N}_{\tilde{\theta}}}\left(x\right)}{N_{\theta}\left(x\right)}$}, (27)

where GQ(2e)2/h{G_{Q}}\equiv{{\left(2e\right)}^{2}}/{h} is the quantum conductance. From Eq.(27), the following elliptical orbit can be drawn:

g1Nθ2+gN~θ~2=1η,\displaystyle\scalebox{1.0}{$\displaystyle g^{-1}{N_{\theta}}^{2}+g\tilde{N}_{\tilde{\theta}}^{2}=\frac{1}{\eta}$},\;\;\;\;\;\;\;\;\;\;\;
g(EJ/ES)1/2,ηEJES/Γ,\displaystyle\scalebox{1.0}{$\displaystyle g\equiv{\left(E\!_{J}/E\!_{S}\right)}^{1/2},\;\eta\equiv\sqrt{{E\!_{J}}{E\!_{S}}}/{\Gamma}$}, (28)

where Γ\Gamma is an arbitrary integral constant with an energy dimension. From Eq.(27), one can draw an elliptical orbit as shown in Figure 2.

Refer to caption
         (a) EJESE_{J}\gg E_{S}         (b) EJ=ESE_{J}=E_{S}
Refer to caption
 (c) EJESE_{J}\ll E_{S}
Figure 2: Elliptical orbit with the number Nθ{N_{\theta}} of the Cooper pairs versus the number N~θ~{\tilde{N}_{\tilde{\theta}}} of the magnetic flux quantum

Figure 2 (a) shows a JJJ\!J junction state (superconducting state) for EJESE_{J}\gg E_{S}, Figure 2 (b) shows a quantum Hall state (Bose semiconductor state) for EJ=ESE_{J}=E_{S}, and Figure 2 (c) shows a QPSQ\!P\!S junction state (superinsulating state) for EJESE_{J}\ll E_{S}.The results shown in Figure 2 are very similar to those in referencesDiamantini et al. (2018)Diamantini et al (2017)Diamantini et al. (2019). Compare the result of Eq.(27) with the result of the quantum Hall effect shown below:.

G=GQν,\displaystyle\scalebox{0.85}{$\displaystyle G={G_{Q}}\nu$}, (29)

where ν\nu is the Landau level occupancy and is defined as:

νNeNϕ,\displaystyle\scalebox{0.85}{$\displaystyle\nu\equiv\frac{N_{e}}{N_{\phi}}$}, (30)

where NeN_{e} and NθN_{\theta} are the electron number and the number of magnetic fluxes (vortex number), respectively. In our model, the correspondence is NeNθ{N_{e}}\!\to\!N_{\theta} and NϕN~θ~N_{\phi}\!\to\!\tilde{N}_{\tilde{\theta}}. When the Hall conductivity of the quantum Hall effect in the bulk is derived using the Kubo formula, the Landau level occupancy ν\nu can also be expressed as the topological quantum number, also called the Chern number. From Eqs.(3), (27), (29) and (30), it can be seen that the quantum Hall state is obtained if the following relational expression is satisfied:

EJ[1cosθ(x)]=ES[1cosθ~(x)],\displaystyle\scalebox{0.85}{$\displaystyle E_{J}\Bigl{[}1-\cos\theta\left(x\right)\Bigr{]}=E_{S}\left[1-\cos\tilde{\theta}\left(x\right)\right]$}, (31)

Eq.(31) means that the energy of the Josephson junction is equal to the energy of the quantum phase slip. In the vicinity of the region where the Landau level occupancy ν\nu is represented by the Chern number, it is expected that not only the quantum Hall phase but also a topological insulator phase and a topological metal phase exist. If η>1\eta>1 and g(EJ/ES)1/21g\!\equiv\!\left(E_{J}/E_{S}\right)^{1/2}\!\leqq\!1 , the region means Bose insulatorDiamantini et al. (2019)-Zhang and Schilling (2018), and if η>1\eta>1 and g(EJ/ES)1/21g\equiv\left(E_{J}/E_{S}\right)^{1/2}\geqq 1, the region means Bose metalDiamantini et al. (2019); Das and Doniach (1999)-Phillips (2019). Figure 3 shows a schematic phase diagram for gg versus η\eta.

Refer to caption
Figure 3: Schematic phase diagram for g(EJ/ES)1/2g\equiv{\left(E\!_{J}/E\!_{S}\right)}^{1/2}-ηEJES/Γ\eta\equiv\sqrt{{E\!_{J}}{E\!_{S}}}/{\Gamma}. Schematic phase diagram with gg on the horizontal axis and η\eta on the vertical axis.

4 Dual transformation from the AXY model to the gauged DAXY model by Villain approximation in a 2+1d system

In this section, we perform the dual transformation between the AXYA\!X\!Y model and the DAXYD\!A\!X\!Y model from the Villain approximation.Kleinert (1989)-Janke and Kleinert (1986).First, we apply the Villain approximation to ZAXY{Z^{\prime}}\!_{A\!X\!Y} introduced in Eq.(2) as follows:

.+(EJ)v2j=12(jθ2πnj)2},\displaystyle\scalebox{0.88}{$\displaystyle\Bigl{.}\!+\!\frac{-\left({E^{\prime}}_{J}\right)_{v}}{2}\sum\limits_{j=1}^{2}{\Bigl{(}\nabla_{j}\theta-2\pi{n_{j}}\Bigr{)}}^{2}\Bigr{\}}$}, (32)

where ZQVZ\!_{Q\!V} is the Villain approximation of the partition function ZAXY{Z^{\prime}}\!_{A\!X\!Y} , and RQV[Rv(EJ)2Rv(EJ0)]MMτR\!_{Q\!V}\!\!\equiv\!\!{[R_{v}{\left(E_{J}\right)^{2}}R_{v}({E^{\prime}}_{J}^{0})]^{MM_{\tau}}} is the Villain’s normalization parameter,, where Rv(E)2π(E)vI0(E){R_{v}}\!\left(E\right)\!\!\equiv\!\!\sqrt{2\pi{(E)_{v}}}{I_{0}}\!\left(E\right) , and (E)v{2ln[I0(E)/I1(E)]}1(E)_{v}\!\!\equiv\!\!{\left\{-2\ln[I_{0}(E)/I_{1}(E)]\right\}}^{-1} . I0(E){I_{0}}\!\left(E\right) and I1(E){I_{1}}\!\left(E\right) represents the modified Bessel functions of order zero and order one, respectively. The summation symbols {n}x,τj=02nj(x,τ)=\sum\limits_{\left\{n\right\}}\!\!\equiv\!\!\prod\limits_{x,\tau}\!\prod\limits_{j=0}^{2}\sum\limits_{n_{j}\left(x,\tau\right)=-\infty}^{\infty} are used for the integer fields nj(x,τ)n_{j}\left(x,\tau\right) . The partition function in Eq.(4) is equivalent to the Euclidean version of the quantum vortex dynamics in a film of superfluid helium introduced by KleinertKleinert (1989, 1985). For Eq.(4) the following identities associated with the Jacobi theta function are used:

n=exp{E2(θ2πn)2}=l=12πEexp(b22E+ibθ),\displaystyle\scalebox{0.84}{$\displaystyle\sum\limits_{n=-\infty}^{\infty}\!\!\!\exp\!\left\{\frac{-E}{2}\left(\theta-2\pi n\right)^{2}\right\}\!=\!\!\!\sum\limits_{l=-\infty}^{\infty}\!\!\!{\frac{1}{\sqrt{2\pi E}}\!\exp\!\left(\frac{-b^{2}}{2E}+ib\theta\right)}$}, (33)

As a result, Eq.(4) can be rewritten as follows:

ZQV=CQV{b}δjbj,0expx,τ[b02(x)2(EJ0)v+j=12bj2(x)2(EJ)v],\displaystyle\scalebox{0.85}{$\displaystyle Z\!_{QV}\!=\!C\!_{QV}\!\!\sum\limits_{\left\{b\right\}}{\delta_{{\nabla_{j}}b_{j},0}}\!\exp\!\sum\limits_{x,\tau}\!{\left[\frac{-b_{0}^{2}\left(x\right)}{2{{\left({E^{\prime}}\!_{J}^{0}\right)}_{v}}}+\sum\limits_{j=1}^{2}\frac{-b_{j}^{2}\left(x\right)}{2{{\left({E^{\prime}}\!_{J}\right)}_{v}}}\right]}$}, (34)

where CQV{C_{QV}} is a normalization parameter defined by [I0(EJ)2I0(EJ0)]MMτ{{[{I_{0}}{{\left({E^{\prime}}_{J}\right)}^{2}}{I_{0}}({E^{\prime}}_{J}^{0})]}^{MM_{\tau}}} , and bi(x){b_{i}}\left(x\right) represents auxiliary magnetic fields with integer values. The integer dual vector potentials a~i(x){\tilde{a}_{i}}\!\left(x\right) (i=0,1,2i=0,1,2) is introduced as follows Villain (1975):

bi(x)=εijlja~l(x)=(×𝐚~)i(x),\displaystyle\scalebox{0.98}{$\displaystyle b_{i}\!\left(x\right)={{\varepsilon}_{ijl}}{\nabla_{j}}{\tilde{a}_{l}}\!\left(x\right)={\left(\nabla\times\mathbf{\tilde{a}}\right)_{i}}\!\left(x\right)$}, (35)

where εijl{\varepsilon_{ijl}} is the Levi–Civita symbol of three dimensions. By using the dual transformations of Eq.(35), the following Eq.(36) can be obtained:

ZQVCQV{a~}δjεijlja~l,0expx[(×𝐚~)022(EJ0)v+j=12(×𝐚~)j22(EJ)v],\displaystyle\scalebox{0.73}{$\displaystyle Z\!_{Q\!V}\!\equiv\!{C\!_{Q\!V}}\!\!\sum\limits_{\left\{{\tilde{a}}\right\}}{\delta_{{\nabla_{j}}{\varepsilon_{ijl}}{\nabla_{j}}{\tilde{a}_{l}},0}}\!\exp\!\sum\limits_{x}\left[\frac{-{{\left(\nabla\times\mathbf{\tilde{a}}\right)}_{0}}^{2}}{2{{\left({E^{\prime}}_{J}^{0}\right)}_{v}}}\!+\!\sum\limits_{j=1}^{2}\frac{-{\left(\nabla\times\mathbf{\tilde{a}}\right)_{j}}^{2}}{2{{\left({{E^{\prime}}_{J}}\right)}_{v}}}\right]$}, (36)

Substituting Poisson’s formula in Eq.(37) to (36):

(a~j)=f(a~j)=Dα~jf(α~j)(l~j)=δjl~j,0expx(i2πj=02l~jα~j),\displaystyle\scalebox{0.8}{$\displaystyle\sum\limits_{\left({\tilde{a}_{j}}\right)\!=\!-\infty}^{\infty}\!\!\!\!\!\!f\left({\tilde{a}_{j}}\right)\!=\!\!\!\!\int\limits_{-\infty}^{\infty}\!\!\!D{{\tilde{\alpha}^{\prime}}_{j}}f\!(\tilde{\alpha^{\prime}}_{j})\!\!\!\!\!\!\!\sum\limits_{\left({\tilde{l}_{j}}\right)=-\infty}^{\infty}\!\!\!\!\!\!{\delta_{{\nabla_{j}}{\tilde{l}_{j}},0}\!\exp\!\sum\limits_{x}\!\!{\left(i2\pi\sum\limits_{j=0}^{2}{{\tilde{l}_{j}}{\tilde{\alpha^{\prime}}_{j}}}\right)}}$}, (37)

then Eq..(4) is as follows:

+j=12(×α~)j2(x)2(EJ)v+i2πj=02l~j(x)α~j(x)},\displaystyle\scalebox{0.8}{$\displaystyle\left.+\sum\limits_{j=1}^{2}\frac{-\left(\nabla\times\mathbf{\tilde{\alpha^{\prime}}}\right)_{j}^{2}\left(x\right)}{2{{\left({{E^{\prime}}_{J}}\right)}_{v}}}\!+i2\pi\!\sum\limits_{j=0}^{2}{\tilde{l}_{j}\left(x\right)\tilde{\alpha^{\prime}}_{j}\left(x\right)}\right\}$}, (38)

Poisson’s formula in Eq.(37) converts the integer-valued vector potentials a~i\tilde{a}_{i} to the continuous-valued vector potentials α~i{\tilde{\alpha}^{\prime}}_{i}. Following the quantum vortex dynamics Kleinert (1989, 1985), the Euclidean Lagrangian density of α~i{\tilde{\alpha}^{\prime}}_{i} is as follows:

LQV(x)=β~02(x)2(EJ0)v+i=12β~i2(x)2(EJ)vi2πj=02l~j(x)α~j(x),\displaystyle\scalebox{0.85}{$\displaystyle L_{QV}\!\!\left(x\right)\!=\!\frac{\tilde{\beta}_{0}^{2}\left(x\right)}{2{{\left({E^{\prime}}_{J}^{0}\right)}_{v}}}+\sum\limits_{i=1}^{2}\frac{\tilde{\beta}_{i}^{2}\left(x\right)}{2{{\left({E^{\prime}}_{J}\right)}_{v}}}-i2\pi\sum\limits_{j=0}^{2}\!{{\tilde{l}_{j}}\!\left(x\right){{\tilde{\alpha^{\prime}}}_{j}}\!\left(x\right)}$}, (39)

where, β~i{\tilde{\beta}_{i}}(i=1,2i=1,2) and β~0{\tilde{\beta}_{0}} can be considered as a dual electric field and a dual magnetic field in a 2+1d2+1d dual electromagnetic field, respectively, and are defined as follows:

β~0(x)1α~2(x)2α~1(x),\displaystyle\scalebox{0.9}{$\displaystyle{\tilde{\beta}_{0}}\left(x\right)\equiv{\nabla_{1}}{\tilde{\alpha^{\prime}}_{2}}\left(x\right)-{{\nabla}_{2}}{\tilde{\alpha^{\prime}}_{1}}\left(x\right)$},
β~1(x)2α~0(x)0α~2(x),\displaystyle\scalebox{0.9}{$\displaystyle{\tilde{\beta}_{1}}\left(x\right)\equiv{\nabla_{2}}{\tilde{\alpha^{\prime}}_{0}}\left(x\right)-{{\nabla}_{0}}{\tilde{\alpha^{\prime}}_{2}}\left(x\right)$},
β~2(x)0α~1(x)1α~0(x),\displaystyle\scalebox{0.9}{$\displaystyle{\tilde{\beta}_{2}}\left(x\right)\equiv{\nabla_{0}}{\tilde{\alpha^{\prime}}_{1}}\left(x\right)-{{\nabla}_{1}}{\tilde{\alpha^{\prime}}_{0}}\left(x\right)$}, (40)

If the 1,2 components e~1{\tilde{e}_{1}} and e~2{\tilde{e}_{2}} of the dual electric field are set as e~1β~2{\tilde{e}_{1}}\equiv{\tilde{\beta}_{2}}, and e~2β~1{\tilde{e}_{2}}\equiv-{\tilde{\beta}_{1}}, respectively, the dual Maxwell’s equations from Lagrangian of Eq.(39) are as follows:

1(EJ)v[1e~1(x)+2e~2(x)]=i2πl~0(x),\displaystyle\scalebox{0.98}{$\displaystyle\frac{1}{\left({E^{\prime}}\!_{J}\right)_{v}}\left[{\nabla_{1}}{\tilde{e}_{1}}\left(x\right)+{\nabla_{2}}{\tilde{e}_{2}}\left(x\right)\right]=i2\pi{\tilde{l}_{0}}\left(x\right)$},
1(EJ0)v2β~0(x)1(EJ)v0e~1(x)=i2πl~1(x),\displaystyle\scalebox{0.98}{$\displaystyle\frac{1}{\left({E^{\prime}}\!_{J}^{0}\right)_{v}}{\nabla_{2}}{\tilde{\beta}_{0}}\left(x\right)-\frac{1}{\left({E^{\prime}}\!_{J}\right)_{v}}{\nabla_{0}}{\tilde{e}_{1}}\left(x\right)=i2\pi{\tilde{l}_{1}}\left(x\right)$},
1(EJ)v0e~2(x)1(EJ0)v1β~0(x)=i2πl~2(x),\displaystyle\scalebox{0.98}{$\displaystyle-\frac{1}{\left({E^{\prime}}\!_{J}\right)_{v}}{\nabla_{0}}{\tilde{e}_{2}}\left(x\right)-\frac{1}{\left({E^{\prime}}\!_{J}^{0}\right)_{v}}{\nabla_{1}}{\tilde{\beta}_{0}}\left(x\right)=i2\pi{\tilde{l}_{2}}\left(x\right)$}, (41)

For Eq.(4), in order to integrate out for continuous-valued vector potentials α~i\tilde{\alpha^{\prime}}_{i}, the partition function when the axial gauge-fixing condition α~0=0\tilde{\alpha^{\prime}}_{0}=0 is imposed is as follows:

.×x,xα~i(x)Dij(x,x)α~j(x)+i2πj=02l~j(x)α~i(x)],\displaystyle\scalebox{0.82}{$\displaystyle\Bigl{.}\times\sum\limits_{x,x^{\prime}}{\tilde{\alpha^{\prime}}_{i}}^{\bot}\left(x\right)D_{ij}^{\bot}\left(x,{x^{\prime}}\right){\tilde{\alpha^{\prime}}_{j}}^{\bot}\left(x^{\prime}\right)\!+\!i2\pi\!\sum\limits_{j=0}^{2}{{\tilde{l}_{j}}^{\bot}\!\left(x\right){\tilde{\alpha^{\prime}}_{i}}^{\bot}\left(x\right)}\Bigr{]}$},
Dij(x,x)δijgab¯ab+¯ij,\displaystyle\scalebox{0.82}{$\displaystyle D_{ij}^{\bot}\left(x,x^{\prime}\right)\equiv-{{\delta}_{ij}}^{\bot}{{g}^{ab}}{{\bar{\nabla}}_{a}}{{\nabla}_{b}}+{{\bar{\nabla}}_{i}}^{\bot}{{\nabla}_{j}}^{\bot}$}, (42)

where the orthogonal symbol \bot as a superscript indicates that only the components α~1{{\tilde{\alpha}^{\prime}}_{1}} and α~2{{\tilde{\alpha}^{\prime}}_{2}} exist, and the metric gabg^{ab} is defined as follows:

gab(γ00010001),γ(EJ0)v(EJ)v,\displaystyle\scalebox{0.8}{$\displaystyle g^{ab}\equiv\begin{pmatrix}\gamma&0&0\\[-12.0pt] 0&1&0\\[-12.0pt] 0&0&1\\ \end{pmatrix},\;\;\;\gamma\equiv\frac{\left(E_{J}^{0}\right)_{v}}{\left(E_{J}\right)_{v}}$}, (43)

where γ\gamma is an anisotropic parameter in the JJJ\!J model. Integrating over the continuous-valued gauge fields α~1{{\tilde{\alpha}^{\prime}}_{1}} and α~2{{\tilde{\alpha}^{\prime}}_{2}} for Eq.(4) yields following the partition function:

.2π2(EJ)vl~0(x)𝒱0(xx)l~0(x)],\displaystyle\scalebox{0.78}{$\displaystyle\Bigl{.}-2{{\pi}^{2}}{{({E^{\prime}}_{J})}_{v}}{\tilde{l}_{0}}\left(x\right){{\cal{V}}_{0}}\left(x-x^{\prime}\right){\tilde{l}_{0}}\left(x^{\prime}\right)\Bigr{]}$}, (44)

where CQVCQV[det(ηab¯ab)]12[det(¯00)]12{C^{\prime}_{QV}}\equiv{C_{QV}}{{\left[\det\left(-{{\eta}^{ab}}{\bar{\nabla}_{a}}{\nabla_{b}}\right)\right]}^{\frac{-1}{2}}}{{\left[\det\left(-{\bar{\nabla}_{0}}{\nabla_{0}}\right)\right]}^{\frac{-1}{2}}} , and the anisotropic massless lattice potential (or lattice Green’s function)Kleinert (1989) 𝒱0(x){{\cal{V}}_{0}}\left(x\right) is defined as:

𝒱0(x)1gab¯ab(x),\displaystyle\scalebox{0.85}{$\displaystyle{{\cal{V}}_{0}}\left(x\right)\equiv\frac{-1}{g^{ab}\bar{\nabla}_{a}\nabla_{b}}\left(x\right)$},\quad (45)

Now, from this lattice potential, the ”split lattice potential” 𝒱0(x){{\cal{V^{\prime}}}_{0}}\left(x\right) obtained by dividing the ”core lattice potential” 𝒱0(0){\cal{V}}_{0}\left(0\right) and ”split difference operator”i{{\nabla^{\prime}}_{i}} are introduced as follows:

(46)

Thus, we have the following identity:

=CDα~iexpx[(×α~)022(EJ0)v+j=12(×α~)j22(EJ)v],\displaystyle\scalebox{0.64}{$\displaystyle\!=\!C^{\prime}\!\!\!\int{D\tilde{\alpha^{\prime}}_{i}}\exp\!\sum\limits_{x}{\left[\frac{-\left({\nabla^{\prime}}\times\mathbf{\tilde{\alpha^{\prime}}}\right)_{0}^{2}}{2\left({E^{\prime}}_{J}^{0}\right)_{v}}+\sum\limits_{j=1}^{2}\frac{-{\left({\nabla^{\prime}}\times\mathbf{\tilde{\alpha^{\prime}}}\right)_{j}^{2}}}{2\left({E^{\prime}}_{J}\right)_{v}}\right]}$},\quad (47)

where C[det(¯~l~l)]12[det(¯~0~0)]12C^{\prime}\!\equiv\!{{\left[\det\left(-\tilde{\bar{\nabla^{\prime}}}_{l}\tilde{\nabla^{\prime}}_{l}\right)\right]}^{\frac{1}{2}}}{{\left[\det\left(-\tilde{\bar{\nabla^{\prime}}}_{0}{\tilde{\nabla}^{\prime}}_{0}\right)\right]}^{\frac{1}{2}}} . Substituting Eqs.(4) and (4) in Eq.(4) gives:

×expx[2π2𝒱0(0)(EJ0)vl~j22π2𝒱0(0)(EJ)vl~02+i2πj=02l~jα~j],\displaystyle\scalebox{0.7}{$\displaystyle\!\times\!\exp\!\sum\limits_{x}{\left[-2{{\pi}^{2}}{{\cal{V}}_{0}}\left(0\right){{\left({E^{\prime}}_{J}^{0}\right)}_{v}}{{\tilde{l}}_{j}}^{2}-2{{\pi}^{2}}{{\cal{V}}_{0}}\left(0\right){{\Bigl{(}{E^{\prime}}_{J}\Bigr{)}}_{v}}{\tilde{l}_{0}}^{2}+i2\pi\sum\limits_{j=0}^{2}{{{\tilde{l}}_{j}}{{{\tilde{\alpha^{\prime}}}}_{j}}}\right]}$}, (48)

where CQV′′CQV[det(¯~l~l)]12[det(¯~0~0)]12{C^{\prime\prime}_{QV}}\!\equiv\!{C^{\prime}_{QV}}{{\left[\det\left(-{\tilde{\bar{\nabla^{\prime}}}_{l}}{\tilde{\nabla^{\prime}}_{l}}\right)\right]}^{\frac{1}{2}}}{{\left[\det\left(-{\tilde{\bar{\nabla^{\prime}}}_{0}}{\tilde{\nabla^{\prime}}\!_{0}}\right)\right]}^{\frac{1}{2}}} , and the metric fab{f^{ab}} is defined as::

fab(1000γ000γ).\displaystyle\scalebox{0.9}{$\displaystyle f^{ab}\equiv\begin{pmatrix}1&0&0\\[-12.0pt] 0&\gamma&0\\[-12.0pt] 0&0&\gamma\end{pmatrix}$}. (49)

Furthermore, from Eqs.(2), (4), and (34), the following identity can be obtained:

Dθ~exp(1)x{[1cos(0θ~2πα~0)]4π2𝒱0(0)(EJ)v+j=12[1cos(jθ~2πα~j)]4π2𝒱0(0)(EJ0)v},\displaystyle\scalebox{0.62}{$\displaystyle\approx\!\!\!{\int\!\!D\tilde{\theta}}\exp\!\left(-1\right)\!\sum\limits_{x}\!{\left\{\frac{\left[1-\cos\left({\nabla_{0}}\tilde{\theta}-2\pi{\tilde{\alpha^{\prime}}_{0}}\right)\right]\!\!}{4{\pi}^{2}{{\cal{V}}_{0}}\left(0\right){\Bigl{(}{{E^{\prime}}_{J}}\Bigr{)}}_{v}}+\frac{\sum\limits_{j=1}^{2}{\left[1-\cos\left({\nabla}_{j}\tilde{\theta}-2\pi{\tilde{\alpha^{\prime}}_{j}}\right)\right]\!\!}}{4{{\pi}^{2}}{{\cal{V}}_{0}}\left(0\right){{\Bigl{(}{E^{\prime}}_{J}^{0}\Bigr{)}}_{v}}}\right\}}$}, (50)

As an analogy to the relational expression between the JJJ\!J energy and the QPSJQ\!P\!S\!J energy in Eq.(12),the following relational expression is obtained from Eq.(4).

(51)

Eqs.(12) and (51) differ only in the coefficients of the core lattice potential V0(0){V_{0}}\left(0\right), therefore, Eq.(51) is a reasonable result. Substituting Eqs.(4) and (51) in Eq.(4) gives:

×x{(ES0)v[1cos(0θ~2πα~0)]+(ES)vj=12[1cos(jθ~2πα~j)]},\displaystyle\scalebox{0.66}{$\displaystyle\times\!\sum\limits_{x}{\left\{\!\!{({E^{\prime}}_{S}^{0})}_{v}\!\!\left[1\!-\!\cos\left({\nabla_{0}}\tilde{\theta}-2\pi{\tilde{\alpha^{\prime}}_{0}}\right)\right]\!+\!{{({E^{\prime}}\!_{S})}_{v}}\!\!\sum\limits_{j=1}^{2}{\left[1\!-\!\cos\!\left({\nabla}_{j}\tilde{\theta}-2\pi{{\tilde{\alpha^{\prime}}}_{j}}\right)\right]}\right\}}$}, (52)

Eq.(4) represents the gauge-coupled QPSJQ\!P\!S\!J partition function by the dual gauge field α~i\tilde{\alpha^{\prime}}_{i}. For the coefficient (EJ0)v{{\left({E^{\prime}}\!_{J}^{0}\right)}\!_{v}} of the dual gauge field energy, use the relationships in Eqs.(51) and (7), and next, when these non-dimensional energy constants have been converted to the original energy dimension, Eq. (and next, when these non-dimensional energy constants have been converted to the original energy dimension, Eq. (52) will be as follows:) will be as follows:

×exp(a0)τ,𝒙{(ES0)v[1cos(0θ~2πα~0)]+(ES)vj=12[1cos(jθ~2πα~j)]},\displaystyle\scalebox{0.62}{$\displaystyle\times\!\!\exp\!\!\left(\!\frac{-a_{0}}{\hbar}\!\right)\!\!\sum\limits_{\tau,\bm{x}}{\!\left\{\!(E_{S}^{0})_{v}\!\!\left[\!1\!-\!\cos\!\left({{\nabla}_{0}}\tilde{\theta}\!-\!2\pi{{\tilde{\alpha^{\prime}}}_{0}}\right)\!\right]\!+\!(E_{S})_{v}\!\!\sum\limits_{j=1}^{2}\!{\left[\!1\!-\!\cos\!\left({{\nabla}_{j}}\tilde{\theta}\!-\!2\pi{{\tilde{\alpha^{\prime}}}_{j}}\right)\!\right]}\!\right\}}$}, (53)

Furthermore, by scaling 2e2e to the non-dimensional dual gauge field α~i\tilde{\alpha^{\prime}}\!_{i}, we introduce a new dual gauge field α~i\tilde{\alpha}_{i} as follows:

2eα~i(x)α~i(x),\displaystyle\scalebox{0.95}{$\displaystyle 2e{\tilde{\alpha^{\prime}}_{i}}\left(x\right)\equiv{\tilde{\alpha}_{i}}\left(x\right)$}, (54)

By the transformation of Eq.(54), Eq.(4) is transformed as follows:

×τ,𝒙{ES0[1cos(0θ~qmα~0)]+ESj=12[1cos(jθ~qmα~j)]},\displaystyle\scalebox{0.7}{$\displaystyle\times\!\sum\limits_{\tau,\bm{x}}\!\left\{E_{S}^{0}\!\left[1\!-\!\cos\!\left({\nabla}_{0}\tilde{\theta}-q_{m}{\tilde{\alpha}_{0}}\right)\right]\!+\!{E_{S}}\!\sum\limits_{j=1}^{2}{\left[1\!-\!\cos\!\left({\nabla}_{j}\tilde{\theta}-q_{m}{\tilde{\alpha}_{j}}\right)\right]}\!\right\}$}, (55)

where μ′′~C/V0(0)\tilde{\mu^{\prime\prime}}\!\equiv\!C/V_{0}\!\left(0\right) , and qm2π/2e=Φ0/q_{m}\!\equiv\!2\pi\!/2e\!=\!\Phi_{0}\!/\hbar represents a unit magnetic charge. Eq.(4) shows that the pure AXYA\!X\!Y model ( JJJ\!J model without gauge coupling) has been dual transformed to the gauged DAXYD\!A\!X\!Y model (gauged QPSJQ\!P\!S\!J model) by the Villain approximation in the 2+1d2+1d system. In other words, the gauge QPSJQ\!P\!S\!J model is a “frozen lattice dual superconductor”Kleinert (1989); Herbut (2007)-Neuhaus et al. (2003), for the JJJ\!J model without gauge coupling.

5 Dual transformation from the DAXY model to the gauged AXY model by Villain approximation in a 2 + 1 d system

In this section, contrary to the previous section, we show the dual transformation from the DAXYD\!A\!X\!Y model to the AXYA\!X\!Y model by the Villain approximation. First, apply the Villain approximation to ZDAXY{Z^{\prime}}_{D\!A\!X\!Y} introduced in Eq.(2) as follows:

ZQDVRQDVDθ~{n~}expx[(ES0)v2(τθ~2πn~0)2+(ES)v2j=12(jθ~2πn~j)2],\displaystyle\scalebox{0.63}{$\displaystyle Z_{Q\!D\!V}\!\!\equiv\!\!{R_{Q\!D\!V}}\!\!\!\int\!\!\!D\tilde{\theta}\sum\limits_{\left\{\tilde{n}\right\}}\exp\!\sum_{x}\!\!\left[\frac{\!\!-\!{\left(\!{E^{\prime}}_{S}^{0}\!\right)_{v}}}{2}{{\Bigl{(}\!{\nabla}_{\tau}\tilde{\theta}\!-\!2\pi{\tilde{n}_{0}}\!\Bigr{)}}\!^{2}}\!\!+\!\!\frac{-{{\Bigl{(}\!{E^{\prime}_{S}}\!\Bigr{)}}_{v}}}{2}\!\!\sum\limits_{j=1}^{2}\!{{\Bigl{(}\!{{\nabla}_{j}}\tilde{\theta}\!-2\pi{\tilde{n}_{j}}\!\Bigr{)}}\!^{2}}\!\right]$}, (56)

where RDQV[Rv(ES)2Rv(ES0)]MMτR_{D\!Q\!V}\!\equiv\!{{[R_{v}\!\left({E^{\prime}}\!_{S}\right)}^{2}{R_{v}}\!({E^{\prime}}\!_{S}^{0})]}^{M{M_{\tau}}} , and ZQDVZ\!_{Q\!D\!V} represents the Villain approximations of the partition function ZDAXY{Z^{\prime}}\!_{D\!A\!X\!Y}. Using the Jacobi theta function of Eq.(33), Eq.(56) can be rewritten as follows:

ZQDV=CQDV{b~}δjb~j,0expx[b~02(x)2(ES0)v+j=12b~j2(x)2(ES)v],\displaystyle\scalebox{0.85}{$\displaystyle Z\!_{Q\!D\!V}\!=\!C\!_{Q\!D\!V}\!\!\sum\limits_{\left\{\tilde{b}\right\}}{\delta_{{\nabla_{j}}\tilde{b}_{j},0}}\exp\sum\limits_{x}\!{\left[\frac{-\tilde{b}_{0}^{2}\left(x\right)}{2{{\left({E^{\prime}}\!_{S}^{0}\right)}_{v}}}+\sum\limits_{j=1}^{2}\frac{-\tilde{b}_{j}^{2}\left(x\right)}{2{{\left({E^{\prime}}\!_{S}\right)}_{v}}}\right]}$}, (57)

where CQDV[I0(EJ)2I0(EJ0)]MMτ{C_{Q\!D\!V}}\!\equiv\!{{[{I_{0}}{{\left({E^{\prime}}_{J}\right)}^{2}}{I_{0}}({E^{\prime}}_{J}^{0})]}^{M{M_{\tau}}}} , and b~i(x){\tilde{b}_{i}}\left(x\right) represents auxiliary dual magnetic fields with integer values. Integer vector potentials ai(x){a_{i}}\left(x\right)(i=0,1,2i\!=\!0,1,2) are introduced as follows :

b~i(x)=εijljal(x)=(×𝐚)i(x),\displaystyle\scalebox{0.98}{$\displaystyle\tilde{b}_{i}\!\left(x\right)\!=\!{{\varepsilon}_{ijl}}{\nabla_{j}}{a_{l}}\!\left(x\right)={\left(\nabla\times\mathbf{a}\right)_{i}}\!\left(x\right)$}, (58)

By using the dual transformations of Eq.(58), the following Eq.(59) is obtained:

ZQDV=CQDV{a}δjεijljal,0expx[(×𝐚)022(ES0)v+j=12(×𝐚)j22(ES)v],\displaystyle\scalebox{0.76}{$\displaystyle Z\!_{Q\!D\!V}\!=\!C\!_{Q\!D\!V}\!\!\sum\limits_{\left\{{a}\right\}}\!{\delta_{{\nabla\!_{j}}{\varepsilon\!_{ijl}}{\nabla\!_{j}}{a_{l}},0}}\!\exp\!\sum\limits_{x}\!\!\left[\frac{-{{\left(\nabla\!\times\!\mathbf{a}\right)}_{0}}^{2}}{2{{\left({E^{\prime}}_{S}^{0}\right)}_{v}}}\!+\!\sum\limits_{j=1}^{2}\!\frac{-{\left(\nabla\!\times\!\mathbf{a}\right)_{j}}^{2}}{2{{\left({E^{\prime}}_{S}\right)}_{v}}}\right]$}, (59)

Using Poisson’s formula in the following Eq.(37) for Eq.(59):

.+12(ES0)v(×α)02+i2πj=02ljαj],\displaystyle\scalebox{0.8}{$\displaystyle\Bigl{.}+\frac{-1}{2{{\left({E^{\prime}}_{S}^{0}\right)}_{v}}}\left(\nabla\times\mathbf{\alpha^{\prime}}\right)_{0}^{2}+i2\pi\!\sum\limits_{j=0}^{2}{{l_{j}}{{\alpha^{\prime}}_{j}}}\Bigr{]}$}, (60)

The Euclidean Lagrangian density of αj{\alpha^{\prime}}\!_{j} is as follows:

LQDV(x)=12(ES)vi=12βi2(x)+12(ES0)vβ02(x)i2πj=02lj(x)αj(x),\displaystyle\scalebox{0.75}{$\displaystyle L\!_{Q\!D\!V}\!\!\left(x\right)\!\!=\!\!\frac{1}{2{{\left({E^{\prime}}\!_{S}\right)}_{v}}}\!\!\sum\limits_{i=1}^{2}{{\beta}_{i}^{2}\left(x\right)}\!+\!\frac{1}{2{{\left({E^{\prime}}\!_{S}^{0}\right)}_{v}}}{\beta}_{0}^{2}\left(x\right)-i2\pi\sum\limits_{j=0}^{2}\!{{l_{j}}\!\left(x\right){{\alpha^{\prime}}_{j}}\!\left(x\right)}$}, (61)

where,βi{\beta_{i}}(i=1,2i=1,2) and β0{\beta_{0}} can be considered as an electric field and a magnetic field in a 2+1d2+1d electromagnetic field, respectively, and are defined as follows:

β0(x)1α2(x)2α1(x),\displaystyle\scalebox{0.9}{$\displaystyle{\beta}_{0}\left(x\right)\equiv{\nabla_{1}}{{\alpha^{\prime}}_{2}}\left(x\right)-{{\nabla}_{2}}{{\alpha^{\prime}}_{1}}\left(x\right)$},
β1(x)2α0(x)0α2(x),\displaystyle\scalebox{0.9}{$\displaystyle{\beta}_{1}\left(x\right)\equiv{\nabla_{2}}{{\alpha^{\prime}}_{0}}\left(x\right)-{{\nabla}_{0}}{{\alpha^{\prime}}_{2}}\left(x\right)$},
β2(x)0α1(x)1α0(x),\displaystyle\scalebox{0.9}{$\displaystyle{\beta}_{2}\left(x\right)\equiv{\nabla_{0}}{{\alpha^{\prime}}_{1}}\left(x\right)-{{\nabla}_{1}}{{\alpha^{\prime}}_{0}}\left(x\right)$}, (62)

If the 1,2 components e1{{e}_{1}} and e2{{e}_{2}} of the electric field are set as e1β2{{e}_{1}}\equiv{{\beta}_{2}}, and e2β1{{e}_{2}}\equiv-{{\beta}_{1}}, respectively, the Maxwell’s equations from the Lagrangian of Eq.(61) are as follows:

1(ES)v[1e1(x)+2e2(x)]=i2πl0(x),\displaystyle\scalebox{0.9}{$\displaystyle\frac{1}{\left({E^{\prime}}\!_{S}\right)_{v}}\left[{\nabla_{1}}{e_{1}}\left(x\right)+{\nabla_{2}}{e_{2}}\left(x\right)\right]=i2\pi{l_{0}}\left(x\right)$},
1(ES0)v2β0(x)1(ES)v0e1(x)=i2πl1(x),\displaystyle\scalebox{0.9}{$\displaystyle\frac{1}{\left({E^{\prime}}\!_{S}^{0}\right)_{v}}{\nabla_{2}}{\beta_{0}}\left(x\right)-\frac{1}{\left({E^{\prime}}\!_{S}\right)_{v}}{\nabla_{0}}{e_{1}}\left(x\right)=i2\pi{l_{1}}\left(x\right)$},
1(ES)v0e2(x)1(ES0)v1β0(x)=i2πl2(x),\displaystyle\scalebox{0.9}{$\displaystyle-\frac{1}{\left({E^{\prime}}\!_{S}\right)_{v}}{\nabla_{0}}{e_{2}}\left(x\right)-\frac{1}{\left({E^{\prime}}\!_{S}^{0}\right)_{v}}{\nabla_{1}}{{\beta}_{0}}\left(x\right)=i2\pi{l_{2}}\left(x\right)$}, (63)

Integrating over the continuous-valued gauge fields α1{{\alpha^{\prime}}\!_{1}} and α2{{\alpha^{\prime}}\!_{2}} for Eq.(5) yields the following partition function:

.2π2(EJ)vl0(x)𝒱~0(xx)l0(x)],\displaystyle\scalebox{0.78}{$\displaystyle\Bigl{.}\!-\!2{{\pi}^{2}}{{\Bigl{(}{E^{\prime}}_{J}\Bigl{)}}_{v}}\!{l_{0}}\left(x\right)\!{\tilde{\cal{V}}_{0}}\!\left(x-x^{\prime}\right)\!{l_{0}}\left(x^{\prime}\right)\Bigr{]}$}, (64)

where CQDVCQDV[det(ηab¯ab)]12[det(¯00)]12{{C^{\prime}}\!_{Q\!D\!V}}\equiv{C\!_{Q\!D\!V}}{{\left[\det\left(-{{\eta}^{ab}}{{\bar{\nabla}}_{a}}{{\nabla}_{b}}\right)\right]}^{\frac{-1}{2}}}{{\left[\det\left(-{{\bar{\nabla}}_{0}}{{\nabla}_{0}}\right)\right]}^{\frac{-1}{2}}} , and the anisotropic massless lattice potential (or lattice Green’s function) V~0(x)\tilde{V}_{0}\left(x\right) is defined as:

g~ab(γ~00010001),γ~(ES0)v(ES)v,\displaystyle\scalebox{0.83}{$\displaystyle{{\tilde{g}}^{ab}}\equiv\left(\begin{matrix}{\tilde{\gamma}}&0&0\\[-12.0pt] 0&1&0\\[-12.0pt] 0&0&1\end{matrix}\right),\;\;\tilde{\gamma}\equiv\frac{{\left(E_{S}^{0}\right)_{v}}}{\left({E_{S}}\right)_{v}}$}, (65)

From this lattice potential, we introduce a ”split lattice potential” 𝒱~0(x)𝒱~0(x)𝒱~0(0)δx,0\tilde{\cal{V^{\prime}}}_{0}\left(x\right)\equiv{\tilde{\cal{V}}_{0}}\left(x\right)\!-\!{{\tilde{\cal{V}}}_{0}}\left(0\right){{\delta}_{x,0}} which is obtained by dividing the ”core lattice potential” 𝒱~0(0){\tilde{\cal{V}}}_{0}\left(0\right) and ”split difference operator” j{\nabla^{\prime}}_{j} , as follows:

Similar from Eq.(4) to (51) in the previous section, the partition function ZDQV{Z_{DQV}} can be written as follows:

×x{(EJ0)v[1cos(0θ~2πα0)]+(EJ)vj=12[1cos(jθ~2παj)]},\displaystyle\scalebox{0.65}{$\displaystyle\!\times\!\sum\limits_{x}{\left\{\!\!({E^{\prime}}_{J}^{0})_{v}\!\!\left[1\!-\!\cos\left({\nabla_{0}}\tilde{\theta}-2\pi{{\alpha^{\prime}}\!_{0}}\right)\right]\!+\!({E^{\prime}}\!_{J})_{v}\!\!\sum\limits_{j=1}^{2}{\left[1\!-\!\cos\!\left({\nabla}_{j}\tilde{\theta}-2\pi{{\alpha^{\prime}}\!_{j}}\right)\right]}\right\}}$}, (66)

where C′′QDVCQDV[det(¯~l~l)]12[det(¯~0~0)]12{{C^{\prime\prime}}\!_{Q\!D\!V}}\!\equiv\!{{C^{\prime}}\!_{Q\!D\!V}}{{\left[\det\!\left(-{{\tilde{\bar{\nabla^{\prime}}}}_{l}}{{\tilde{\nabla^{\prime}}}_{l}}\right)\right]}^{\frac{1}{2}}}{{\left[\det\!\left(-{{\tilde{\bar{\nabla^{\prime}}}}_{0}}{{\tilde{\nabla^{\prime}}}_{0}}\right)\right]}^{\frac{1}{2}}} , and the metric f~ab{{\tilde{f}}^{ab}} is defined as:

(67)

where (EJ0)v({E^{\prime}}_{J}^{0})\!_{v} and (EJ)v({E^{\prime}}_{J})\!_{v} are defined as follows:

(68)

If 𝒱0(0)𝒱~0(0){{\cal{V}}_{0}}\left(0\right)\!\equiv\!{\tilde{\cal{V}}_{0}}\left(0\right) holds, Eqs.(68) and (51) are completely equivalent. Eq.(5) represents the gauge-coupled JJJJ partition function by the gauge field αj{\alpha^{\prime}}\!_{j}. For the coefficient (ES0)v{{({E^{\prime}}\!_{S}^{0})}_{v}} of the gauge field energy, use the relationship between Eqs.(68) and (7), and then, when these non-dimensional energy constants are converted to the original energy dimension, Eq. (68) will be as follows:

×exp(a0)τ,𝒙{(EJ0)v[1cos(0θ2πα0)]+(EJ)vj=12[1cos(jθ2παj)]},\displaystyle\scalebox{0.64}{$\displaystyle\times\!\!\exp\!\!\left(\!\frac{-a_{0}}{\hbar}\!\right)\!\!\sum\limits_{\tau,\bm{x}}{\!\left\{\!(E_{J}^{0})_{v}\!\Bigl{[}1\!-\!\cos\!\left({{\nabla}\!_{0}}{\theta}\!-\!2\pi{{\alpha^{\prime}}\!\!_{0}}\right)\Bigr{]}\!+\!(E_{J})_{v}\!\sum\limits_{j=1}^{2}\!{\Bigl{[}1\!-\!\cos\!\left({{\nabla}\!\!_{j}}{\theta}\!-\!2\pi{{\alpha^{\prime}}\!\!_{j}}\right)\Bigr{]}}\!\right\}}$}, (69)

Furthermore, by scaling Φ0{{\Phi}_{0}} to the non-dimensional gauge field αi{{\alpha^{\prime}}_{i}}, we introduce a new gauge field αi{{\alpha}_{i}} as follows:

Φ0αi(x)αi(x),\displaystyle\scalebox{0.95}{$\displaystyle{\Phi_{0}}{{\alpha^{\prime}}\!_{i}}\left(x\right)\!\equiv\!{\alpha_{i}}\left(x\right)$}, (70)

Eq.(5) is transformed as follows:

×τ,𝒙{(EJ0)v[1cos(0θ2qα0)]+(EJ)vj=12[1cos(jθ2qαj)]},\displaystyle\scalebox{0.67}{$\displaystyle\!\times\!\!\sum\limits_{\tau,\bm{x}}{\!\left\{(E_{J}^{0})_{v}\!\Bigl{[}1\!-\!\cos\!\left({{\nabla}_{0}}{\theta}-2q{{\alpha}_{0}}\right)\Bigr{]}\!+\!(E_{J})_{v}\!\sum\limits_{j=1}^{2}{\Bigl{[}1\!-\!\cos\!\left({{\nabla}_{j}}{\theta}-2q{{\alpha}_{j}}\right)\Bigr{]}}\!\right\}}$}, (71)

where μL/𝒱~0(0)\mu\!\equiv\!L/{\tilde{\cal{V}}_{0}\left(0\right)} , 2q2e/=2π/Φ02q\equiv{2e}/{\hbar}\!=\!{2\pi}/{\Phi_{0}} represents a unit Cooper pair charge, which is twice the unit charge qe/=2π/2Φ0q\equiv{e}/{\hbar}\!=\!{2\pi}/{2{\Phi_{0}}} . Eq.(70) shows that the pure DAXYD\!A\!X\!Y model (JJJ\!J model without gauge coupling) has been dual transformed to the gauged AXYA\!X\!Y model (gauged JJJ\!J model) by the Villain approximation in the 2+1d2+1d system. In other words, the gauge JJJ\!J model is a “frozen lattice dual superconductor” for the QPSQ\!P\!S model without gauge coupling.

6 Mean field analysis of the gauged QPSJ model on the nanosheet

In this section, we introduce the mean field approximation to the partition function of Eq.(4) and discuss its phase transition. From Eq. (4), the partition function excluding the constant part is newly defined as ZGQPJ{Z_{G\!Q\!P\!J}}of the gauged QPSJQ\!P\!S\!J model, and using unit vectors of two real components U~i=[cosθ~,sinθ~]{\tilde{U}_{i}}=[\cos\tilde{\theta},\sin\tilde{\theta}], Eq. (4) is rewritten as follows:

(72)

where μ~μ′′~/a0{\tilde{\mu^{\prime}}}\equiv{\hbar{\tilde{\mu^{\prime\prime}}}}/{{a_{0}}}, and the lattice difference operator R~α~{{\tilde{R}}_{{\tilde{\alpha}}}} is defined as:

R~α~1+12d~(i=12D~¯iD~i+γ~D~¯τD~τ),γ~ES0ES,\displaystyle\scalebox{0.8}{$\displaystyle{\tilde{R}}_{\tilde{\alpha}}\equiv 1+\frac{1}{2\tilde{d}}\left(\sum\limits_{i=1}^{2}{{\bar{\tilde{D}}_{i}}{\tilde{D}}_{i}}+\tilde{\gamma}{\bar{\tilde{D}}_{\tau}}{\tilde{D}}_{\tau}\right),\tilde{\gamma}\equiv\frac{{E^{\prime}}\!_{S}^{0}}{{E^{\prime}}\!_{S}}$}, (73)

where d~2+γ~\tilde{d}\equiv 2+\tilde{\gamma} represents anisotropic dimensional constants of the gauged QPSJQ\!P\!S\!J; and D~i{{\tilde{D}}_{i}} and D~¯i{\bar{\tilde{D}}}_{i} are forward and backward covariant lattice derivatives, respectively, and are defined, for example, for a complex field U~=U~1+iU~2\tilde{U}={\tilde{U}}_{1}\!+\!i{\tilde{U}}_{2} , as follows:

D~iU~(𝐱,τ)U~(𝐱+𝐢,τ)eiqmα~i(𝐱,τ)U~(𝐱,τ),,\displaystyle\scalebox{0.8}{$\displaystyle{{\tilde{D}}_{i}}\tilde{U}\left(\mathbf{x},\tau\right)\equiv\tilde{U}\left(\mathbf{x}+\mathbf{i},\tau\right){{e}^{-i{{q}_{m}}{{{\tilde{\alpha}}}_{i}}\left(\mathbf{x},\tau\right)}}-\tilde{U}\left(\mathbf{x},\tau\right),$},
D~¯iU~(𝐱,τ)U~(𝐱,τ)U~(𝐱𝐢,τ)eiqmα~i(𝐱𝐢,τ),\displaystyle\scalebox{0.8}{$\displaystyle{\bar{\tilde{D}}_{i}}\tilde{U}\left(\mathbf{x},\tau\right)\equiv\tilde{U}\left(\mathbf{x},\tau\right)-\tilde{U}\left(\mathbf{x}-\mathbf{i},\tau\right){{e}^{i{{q}_{m}}{{\tilde{\alpha}}_{i}}\left(\mathbf{x}-\mathbf{i},\tau\right)}}$}, (74)

the same applies to D~τ{\tilde{D}}_{\tau} and D~¯τ{\bar{\tilde{D}}}_{\tau}. In Eq.(6), we introduce two sets of real two-component fields u~l{\tilde{u}}_{l} and ψ~l{\tilde{\psi}}_{l} (l=1,2l=1,2) which satisfy the following identity:

du~1du~2dψ~1dψ~2(2πi)2exp{ψ~l(u~1U~l)}=1,\displaystyle\scalebox{0.8}{$\displaystyle\int_{-\infty}^{\infty}{\text{d}{{\tilde{u}}_{1}}\text{d}{{\tilde{u}}_{2}}}\int_{-\infty}^{\infty}{\frac{\text{d}{{\tilde{\psi}}_{1}}\text{d}{{\tilde{\psi}}_{2}}}{{{\left(2\pi i\right)}^{2}}}\exp\left\{-{{\tilde{\psi}}_{l}}\left({{\tilde{u}}_{1}}-{{\tilde{U}}_{l}}\right)\right\}=1}$}, (75)
×expxl=12[ESd~u~lR~α~u~lψ~lu~l+lnI0(|ψ~l|)],\displaystyle\scalebox{0.8}{$\displaystyle\!\times\!\exp\!\sum\limits_{x}\!{\sum\limits_{l=1}^{2}\!{{\left[{E^{\prime}}_{S}\tilde{d}{{\tilde{u}}_{l}}{{\tilde{R}}_{\tilde{\alpha}}}{{\tilde{u}}_{l}}-{{\tilde{\psi}}_{l}}{{\tilde{u}}_{l}}+\ln{I_{0}}(|{{\tilde{\psi}}_{l}}|)\right]}}}$},\quad (76)

where we have defined the functional integrals of θ~\tilde{\theta} as follows:

xππdθ~2πexp{xl=12ψ~lU~l}=expx,l{lnI0(|ψ~l|)},\displaystyle\scalebox{0.8}{$\displaystyle\prod\limits_{x}\!\!{\int_{-\pi}^{\pi}\!\!{\frac{d\tilde{\theta}}{2\pi}}}\!\exp\!\left\{\sum\limits_{x}{\sum\limits_{l=1}^{2}{{{\tilde{\psi}}_{l}}{{\tilde{U}}_{l}}}}\right\}\!=\!\exp\!\sum\limits_{x,l}{\left\{\ln{I_{0}}(|{{\tilde{\psi}}_{l}}|)\right\}}$}, (77)

where I0(|ψ~|){I_{0}}(|\tilde{\psi}|) ( |ψ~|ψ~12+ψ~22|{\tilde{\psi}}|\equiv\sqrt{{{\tilde{\psi}}_{1}}^{2}+{{\tilde{\psi}}_{2}}^{2}} ) represents the modified Bessel functions of a zeroth-order integer. In Eq.(6), performing the integrals over u~l{\tilde{u}}_{l} fields, we obtain the partition function by the complex field ψ~ψ~1+iψ~2\tilde{\psi}\equiv{{\tilde{\psi}}_{1}}+i{{\tilde{\psi}}_{2}} and ψ~ψ~1iψ~2{{\tilde{\psi}}^{*}}\equiv{{\tilde{\psi}}_{1}}-i{{\tilde{\psi}}_{2}}.

ψ~^(x)R~α~12ψ~(x),\displaystyle\scalebox{0.9}{$\displaystyle\hat{\tilde{\psi}}\left(x\right)\equiv{{\tilde{R}}_{{\tilde{\alpha}}}}^{\frac{1}{2}}\tilde{\psi}\left(x\right)$},\quad\quad\quad\quad\quad (78)

In Eq.(6), since ψ~\tilde{\psi} and ψ~{\tilde{\psi}}^{*} can be regarded as the order parameters of the superinsulator (i.e., the disorder parameters of the superconductor), the non-dimensional free energy F(ψ~,ψ~,α~i){F^{\prime}}(\tilde{\psi},{{\tilde{\psi}}^{*}},{\tilde{\alpha}}_{i}) can be Landau expanded for terms up to |ψ|4{\left|\psi\right|}^{4}, |Diψ|2{\left|{D_{i}}\psi\right|}^{2} and |Dτψ|2{\left|{D_{\tau}}\psi\right|}^{2} as followsKleinert (1989):

+18d~(i=12|Diψ~|2+γ~|Dτψ~|2)+14(1ESd~1)|ψ~|2+164|ψ~|4},\displaystyle\scalebox{0.75}{$\displaystyle\left.\!+\!\frac{1}{8\tilde{d}}\left(\sum\limits_{i=1}^{2}{{|{D_{i}}\tilde{\psi}|}^{2}}\!+\!\tilde{\gamma}{{\left|{D_{\tau}}\tilde{\psi}\right|}^{2}}\right)\!+\!\frac{1}{4}\left(\frac{1}{{E^{\prime}}_{S}\tilde{d}}\!-\!1\right){{|\tilde{\psi}|}^{2}}\!+\!\frac{1}{64}{{|\tilde{\psi}|}^{4}}\right\}$}, (79)

FDGL{F^{\prime}}\!_{D\!G\!L}is the non-dimensional DGL energy of the superinsulator or QPSJQ\!P\!S\!J model on the nanosheet in d~2+γ~\tilde{d}\equiv 2+\tilde{\gamma} dimension at zero temperature. Therefore, the critical values ESMF{E^{\prime}}_{S}^{MF} according to the mean field approximation of QPSQ\!P\!S amplitude ES{E^{\prime}}_{S} are as follows:

ESMF1d~=12+γ~,\displaystyle\scalebox{0.76}{$\displaystyle{E^{\prime}}_{S}^{MF}\equiv\frac{1}{{\tilde{d}}}=\frac{1}{2+\tilde{\gamma}}$}, (80)

The continuous limit in Eq.(6), is as follows:
FDGL(ψ~,ψ~,A~i)dx0d2x{fab2μ~(×𝐀~)a(×𝐀~)b+12mΦi=12|(iiΦ0A~i)ψ~|2\displaystyle F\!_{D\!G\!L}(\tilde{\psi},{{\tilde{\psi}}^{*}},{{\tilde{A}}_{i}})\!\equiv\!\!\int\!\!{d{x_{0}}}\!\!\int\!\!{{d^{2}}x}\!\left\{\frac{f^{ab}}{2\tilde{\mu}}{{({\partial^{\prime}}\times\mathbf{\tilde{A}})}_{a}}{{({\partial^{\prime}}\times\mathbf{\tilde{A}})}_{b}}\!+\!\frac{1}{2m_{\Phi}}\sum\limits_{i=1}^{2}{{{\left|\left(-i\hbar{\partial}_{i}-{{\Phi}_{0}}{{\tilde{A}}_{i}}\right)\tilde{\psi}\right|}^{2}}}\right.
+γ~2mΦ|(i0Φ0A~0)ψ~|2+α~ε~|ψ~|2+β~|ψ~|4}\displaystyle\left.\quad\quad\quad\quad\quad\quad\quad+\frac{\tilde{\gamma}}{2m_{\Phi}}{{\left|\left(-i\hbar{\partial}_{0}-{\Phi}_{0}{\tilde{A}}_{0}\right)\tilde{\psi}\right|}^{2}}+{\tilde{\alpha}}{\tilde{\varepsilon}}{{|{\tilde{\psi}}|}^{2}}+\tilde{\beta}{{|\tilde{\psi}|}^{4}}\right\}

α~4a0a2,ε~ESMFESES,β~64a0a2,\displaystyle\scalebox{0.8}{$\displaystyle\tilde{\alpha}\equiv\frac{\hbar}{4{a_{0}}{a^{2}}},\;\tilde{\varepsilon}\equiv\frac{{E^{\prime}}_{S}^{M\!F}-{E^{\prime}}_{S}}{{E^{\prime}}_{S}},\;\tilde{\beta}\equiv\frac{\hbar}{64{a_{0}}{a^{2}}}$}, (81)

where μ~C/a2𝒱0(0)\tilde{\mu}\!\!\equiv\!\!{C}/{a^{2}{{\cal{V}}_{0}}\left(0\right)}, mΦ4d~a0m_{\Phi}\!\!\equiv\!\!4\tilde{d}\hbar{a_{0}}is a pseudo-mass of magnetic flux having a dimension of [Js2]\left[\text{J}\cdot{{\text{s}}^{2}}\right], and A~μ{\tilde{A}}_{\mu} represents dual vector potentials having a dimension of [C/m]\left[\text{C}/\text{m}\right]. Therefore, the order parameter ψ~\tilde{\psi} of the superinsulator is gauge coupled to the U(1) dual gauge field A~μ{\tilde{A}}_{\mu} by a unit magnetic charge qm2π/2e=Φ0/q_{m}\!\!\equiv\!\!2\pi/2e\!\!=\!\!{\Phi}_{0}/\hbar, and when ψ~\tilde{\psi} is in the condensed state, As shown in Figure 4, the pair of the Cooper pair and the anti-Cooper pair12, 13 is confined within the superinsulator by the electric flux-tubes.

Refer to caption
Figure 4: Schematic diagram of a superinsulator on a nanosheet at zero temperature.

Figure 4 shows the confinement of electric flux in the superinsulator between the Cooper pair and the anti-Cooper pair. Whether this confinement picture in which the charge of 2e2e for a superinsulator is the smallest unit of charge is correct will be discussed again in Section 8 (Summary and discussion). From Eq.(81), three DGL equations are derived as follows:

jΦi=1μ~ηia(×𝐁~)a=Φ02mΦ{ψ~iψ~(iψ~)ψ~}Φ02mΦ|ψ~|2A~i,\displaystyle\scalebox{0.78}{$\displaystyle{j_{\Phi}}^{i}\!=\!\frac{1}{\tilde{\mu}}{{\eta^{\prime}}^{ia}}{{\left({\partial^{\prime}}\!\times\!\mathbf{\tilde{B}}\right)}_{a}}\!=\!\frac{\hbar{{\Phi}_{0}}}{2m_{\Phi}}\left\{{{\tilde{\psi}}^{*}}{{\partial}_{i}}\tilde{\psi}\!-\!\left({\partial}_{i}{{\tilde{\psi}}^{*}}\right)\tilde{\psi}\right\}\!-\!\frac{{{\Phi}_{0}}^{2}}{m_{\Phi}}|\tilde{\psi}|^{2}\tilde{A}_{i}$}, (82)
jΦ0=1μ~η0a(×𝐁~)a=γΦ02mΦ{ψ~0ψ~(0ψ~)ψ~}Φ02mΦ|ψ~|2A~0,\displaystyle\scalebox{0.76}{$\displaystyle{j_{\Phi}}^{0}\!=\!\frac{1}{\tilde{\mu}}{{\eta^{\prime}}^{0a}}{{\left({\partial^{\prime}}\!\times\!\mathbf{\tilde{B}}\right)}_{a}}\!=\!\frac{\gamma\hbar{{\Phi}_{0}}}{2m_{\Phi}}\left\{{{\tilde{\psi}}^{*}}{{\partial}_{0}}\tilde{\psi}-\left({\partial}_{0}{{\tilde{\psi}}^{*}}\right)\tilde{\psi}\right\}\!-\!\frac{{{\Phi}_{0}}^{2}}{m_{\Phi}}|\tilde{\psi}|^{2}\tilde{A}_{0}$}, (83)
12mΦi=12(iiΦ0A~i)2ψ~+γ2mΦ(i0Φ0A~0)2ψ~+2β|ψ~|2ψ~=αεψ~,\displaystyle\scalebox{0.76}{$\displaystyle\frac{1}{2{m}_{\Phi}}\!\!\sum\limits_{i=1}^{2}\!{{{\left(\!-\!i\hbar{{\partial}_{i}}\!-\!{{\Phi}_{0}}{\tilde{A}_{i}}\right)}^{2}}\tilde{\psi}}\!+\!\frac{\gamma}{2{m}_{\Phi}}{{\left(\!-i\hbar{{\partial}_{0}}\!-\!{{\Phi}_{0}}{\tilde{A}_{0}}\right)}^{2}}\tilde{\psi}\!+\!2\beta|\tilde{\psi}|^{2}\tilde{\psi}\!=\!-\!\alpha\varepsilon\tilde{\psi}$}, (84)

where B~i(x)=εiabaA~b(x){\tilde{B}_{i}}\left(x\right)={\varepsilon_{iab}}{{\partial^{\prime}}_{a}}{\tilde{A}_{b}}\left(x\right) represents a dual magnetic flux density. Eq.(82) and (83) represent the current density and Eq.(84) represents the nonlinear Schrödinger equation for superinsulators. The thermodynamic critical dual magnetic field H~c{\tilde{H}_{c}} for the dual magnetic field H~0=B~0/μ~{\tilde{H}_{0}}={{{\tilde{B}}_{0}}}/{\tilde{\mu}} is as follows:

H~c(ES)=α2ε22βμ~=α2βμ~(1ESMFES),\displaystyle\scalebox{0.76}{$\displaystyle\tilde{H}_{c}\left({E^{\prime}}_{S}\right)=\sqrt{\frac{{{\alpha}^{2}}{{\varepsilon}^{2}}}{2\beta\tilde{\mu}}}=\frac{\alpha}{\sqrt{2\beta\tilde{\mu}}}\left(1-\frac{{E^{\prime}}_{S}^{MF}}{{{E^{\prime}}_{S}}}\right)$}, (85)

where H~c\tilde{H}_{c} is the dimension of the voltage. Both have a dimension of length, the penetration depth λ~(ES)\tilde{\lambda}({E^{\prime}}\!_{S}) and the coherent length ξ~(ES)\tilde{\xi}({E^{\prime}}\!_{S}) in the super insulator, which both have the dimension of length, are as follows:

λ~(ES)=βmΦ2πμ~αΦ02(ESESESMF)12,\displaystyle\scalebox{0.85}{$\displaystyle\tilde{\lambda}\left({E^{\prime}}\!_{S}\right)=\sqrt{\frac{\beta m_{\Phi}}{2\pi\tilde{\mu}\alpha{\Phi_{0}}^{2}}}{{\left(\frac{{E^{\prime}}\!_{S}}{{E^{\prime}}_{S}-{E^{\prime}}\!_{S}^{M\!F}}\right)}^{\frac{1}{2}}}$},\quad (86)
ξ~(ES)=2mΦα(ESESESMF)12.\displaystyle\scalebox{0.85}{$\displaystyle\tilde{\xi}\left({E^{\prime}}\!_{S}\right)=\frac{\hbar}{\sqrt{2{{m}_{\Phi}}\alpha}}{{\left(\frac{{E^{\prime}}\!_{S}}{{E^{\prime}}_{S}-{E^{\prime}}\!_{S}^{M\!F}}\right)}^{\frac{1}{2}}}$}. (87)

From Eqs.(86) and (87), the dual Ginzburg–Landau parameter κ\kappa is as follows:

κ~(ES)=λ~(ES)ξ~(ES)=mΦΦ0βπμ~=2π2e8πB~c(ES)λ~2(ES),\displaystyle\scalebox{0.8}{$\displaystyle\tilde{\kappa}\left({E^{\prime}}_{S}\right)\!=\!\frac{\tilde{\lambda}\left({E^{\prime}}_{S}\right)}{\tilde{\xi}\left({E^{\prime}}_{S}\right)}=\frac{m_{\Phi}}{\hbar{{\Phi}_{0}}}\sqrt{\frac{\beta}{\pi\tilde{\mu}}}\!=\!\frac{2\pi}{2e}\sqrt{8\pi}\tilde{B}_{c}\!\left({E^{\prime}}_{S}\right){{\tilde{\lambda}}^{2}}\!\left({E^{\prime}}_{S}\right)$}, (88)

where B~c=μ~H~c{\tilde{B}_{c}}=\tilde{\mu}{\tilde{H}_{c}} represents the thermodynamic critical dual magnetic flux density. From the analogy of the classification of type I and type II superconductors, the following classifications by type I and type II superinsulators are formed from the sign of the surface energy σ~SN{\tilde{\sigma}_{SN}} and the value of the Ginzburg–Landau parameter κ\kappa at the superinsulator–normal insulator boundary.
i) type-I superinsulator

κ~<12,σ~SN>0.\tilde{\kappa}\textless\frac{1}{\sqrt{2}},\;\;\;\;\;\;{\tilde{\sigma}_{SN}}\textgreater 0.

ii) intermediate of type I and type II superinsulator

κ~=12,σ~SN=0.\tilde{\kappa}=\frac{1}{\sqrt{2}},\;\;\;\;\;\;{{\tilde{\sigma}}_{SN}}=0.

iii) typeII superinsulator

κ~>12,σ~SN<0.\tilde{\kappa}\textgreater\frac{1}{\sqrt{2}},\;\;\;\;\;\;{{\tilde{\sigma}}_{SN}}\textless 0.

From the analogy with the mixed state of the type-II superconductor, the possibility of the existence of the mixed state in the case of the type-II superinsulator is expected, and, from the analogy with the Abrikosov magnetic flux lattice of the type-II superconductor, the existence of an electric flux lattice is also expected in the case of the type-II superinsulator on a nanosheet.

7 Estimating the critical value of the QPS amplitude by the effective energy approach

In the previous section, we derived the thermodynamic critical dual magnetic field H~c\tilde{H}_{c}, the penetration depth λ~\tilde{\lambda}, and the coherent length ξ~\tilde{\xi} from mean field analysis, all of which depended on the difference between the QPSQ\!P\!S amplitude ES{E^{\prime}}\!_{S} and its mean field critical value ESMF{E^{\prime}}\!_{S}^{M\!F} . Since the mean field approximation is a very rough approximation, in this section, we show the calculation result of the mean energy approximation with the contribution of fluctuations up to the two loop corrections by the effective energy approach shows in Appendix B. The critical value (ES)c2loop\left({E^{\prime}}\!_{S}\right)_{c}^{2\text{loop}} of the QPSQ\!P\!S amplitude ES{E^{\prime}}\!_{S} in the mean field approximation with up to the two loop corrections is given to Eq.(B) as a function of the anisotropy parameter γ~\tilde{\gamma}. The results are plotted in Figure 5. Similarly, Figure 5 shows the results for the tricritical point (ES)tri{{\left({E^{\prime}}\!_{S}\right)}^{tri}} of the QPSQ\!P\!S amplitude given in Eq.(113).

Refer to caption
Figure 5: The critical value and the tricritical point as a function of anisotropy parameter in mean field approximation with up to two loop corrections. Where, γ~\tilde{\gamma} is the anisotropy parameter, and (ES)c2loop{{\left({E^{\prime}}\!_{S}\right)}_{c}^{2loop}} and (ES)tri{{\left({E^{\prime}}\!_{S}\right)}^{tri}} are the critical value and the tricritical point of QPSQ\!P\!S amplitud, respectively.

From Eqs. (51) and (11), the QPSQ\!P\!S amplitude ES{{E^{\prime}}\!_{S}} has the following relationship with the charging energy Ec{{E^{\prime}}\!_{c}} :

Ec=2π2𝒱0(0)(ES)v,\displaystyle\scalebox{0.95}{$\displaystyle{E^{\prime}}\!_{c}=2{{\pi}^{2}}{{\cal{V}}_{0}}\left(0\right){{\left({E^{\prime}}\!_{S}\right)}_{v}}$}, (89)

where 𝒱0(0){{\cal{V}}_{0}}\left(0\right) is a massless lattice Green’s function having an anisotropy parameter γ\gamma in the JJJ\!J model defined by Eq.(43).Using Eq.(5), we have plotted in Figure 6 the critical value (Ec)c{{\left({E^{\prime}}_{c}\right)}_{c}} of the charging energy Ec{E^{\prime}}_{c} for various values of the anisotropy parameter γ\gamma in JJJ\!J models as a function of γ~\tilde{\gamma}

Refer to caption
Figure 6: The critical value of the charging energy for various values of the anisotropy parameter γ\gamma in Josephson junction (JJJ\!J) models as a function of γ~\tilde{\gamma}.

8 Summary and discussion

The conclusions of this paper are summarized below. First, using the dual Hamiltonian method, the phase and amplitude relationship between the JJJ\!J and QPSJQ\!P\!S\!J models without gauge coupling on 2+1d2+1d nanosheets at zero temperature was determined, and the relationships between various constants were derived. Furthermore, the exact duality between the JJJ\!J model and the QPSJQ\!P\!S\!J model based on the nonlinear Legendre transformation between the Lagrangian and the Hamiltonian using canonical conjugate variables of infinite order in a compact 2+1d2+1d lattice space was demonstrated. A dual transformation from the AXYA\!X\!Y model to the gauged DAXYD\!A\!X\!Y model by the Villain approximation in the 2+1d2+1d system was derived.there are two main differences between the dual transformation by the dual Hamiltonian method and the dual transformation by the Villain approximation. One is that Eqs.(12) and (51) differ only in the core potential 𝒱0(0){\cal{V}}_{0}\left(0\right). Another difference is that, in the case of the dual transformation by the Villain approximation, there is a gauge coupling by the dual gauge field α~i{{\tilde{\alpha^{\prime}}}_{i}}, however, in the dual transformation by the dual Hamiltonian method, there is no gauge coupling by the dual gauge field. The gauge coupling by the dual gauge field α~i{{\tilde{\alpha^{\prime}}}_{i}} is gauge coupled with the U(1)U(1) dual gauge field by the unit magnetic charge qm2π/2e{q_{m}}\!\equiv\!{2\pi}/{2e} by the scaling of 2e2e introduced in Eq.(54), and, as shown in Figure 4, it was shown that the electric flux of the Cooper pair and the anti-Cooper pair in units of 2e2e was confined in the superinsulator. In the following, we consider whether the superinsulator’s picture of confinement in 2e2e units that is, confinement by a pair consisting of a Cooper pair and an anti-Cooper pair is correct or incorrect. In the case of the magnetic flux confinement for the superconductor, the magnetic flux quantum Φ0=h/2e{{\Phi}_{0}}=h/{2e}, which is the minimum unit of magnetic flux, is confined. However, if Eq.(55) is correct, in the confinement of charges in the superinsulator, 2e=h/Φ02e=h/{{{\Phi}_{0}}}, which is twice the elementary charge e=h/(2Φ0)e=h/{\left(2{{\Phi}_{0}}\right)},which is the minimum unit of charge, is confined. This is clearly inconsistent with the superconducting case. Therefore, in the case of a superinsulator, it should also be considered correct to assume that confinement occurs in units of the elementary charge e=h/(2Φ0)e\!=\!h/{\left(2{{\Phi}_{0}}\right)}, which is the minimum unit of charge. In other words, since the scaling by 2e2e in Eq.(54) in Section 4 was completely artificial, one could change Eq. (54) to scaling by ee as follows:

eα~i(x)α~i(x),\displaystyle\scalebox{0.95}{$\displaystyle e{{\tilde{\alpha^{\prime}}}_{i}}\left(x\right)\equiv{{\tilde{\alpha}}_{i}}\left(x\right)$}, (90)

By the transformation of Eq.(90), Eq.(4) is transformed as follows:

×τ,𝒙{ES0[1cos(0θ~2qmα~0)]+ESj=12[1cos(jθ~2qmα~j)]},\displaystyle\scalebox{0.68}{$\displaystyle\times\!\sum\limits_{\tau,\bm{x}}\!\left\{E_{S}^{0}\!\left[1\!-\!\cos\!\left({\nabla}_{0}\tilde{\theta}-2q_{m}{\tilde{\alpha}_{0}}\right)\right]\!+\!{E_{S}}\!\sum\limits_{j=1}^{2}{\left[1\!-\!\cos\!\left({\nabla}_{j}\tilde{\theta}-2q_{m}{\tilde{\alpha}_{j}}\right)\right]}\!\right\}$}, (91)

When Eq.(8) is compared with Eq.(4), the coupling magnetic charge is 2qm=2Φ0/=2π/e2q_{m}\!=\!2{\Phi}_{0}/\hbar\!=\!2\pi/e, which is twice the unit magnetic charge qm=Φ0/=2π/2eq_{m}\!=\!{\Phi}_{0}/\hbar\!=\!2\pi/{2e}. Figure 7 shows the confinement of the electric flux between the positive and the negative elementary charges in the superinsulator, that is, it differs from Figure 4.

Refer to caption
Figure 7: Schematic diagram of a superinsulator with an elementary charge ee on a nanosheet at zero temperature.

To obtain the result of Eq.(8), EcE_{c} introduced in Eq.(1) is not (2e)2/2C{\left(2e\right)}^{2}/2C but rather e2/2Ce^{2}/{2C}. In this case, the DGLD\!G\!L free energy for Eq.(8) is as follows: FDGL(ψ~,ψ~,A~i)dx0d2x{fab2μ~(×𝐀~)a(×𝐀~)b+12mΦi=12|(ii2Φ0A~i)ψ~|2\displaystyle F\!_{D\!G\!L}(\tilde{\psi},{{\tilde{\psi}}^{*}},{{\tilde{A}}_{i}})\!\!\equiv\!\!\!\int\!\!{d{x_{0}}}\!\!\int\!\!{{d^{2}}\!x}\!\left\{\frac{f^{ab}}{2\tilde{\mu}}{{({\partial^{\prime}}\!\times\!\mathbf{\tilde{A}})}_{a}}{{({\partial^{\prime}}\!\times\!\mathbf{\tilde{A}})}_{b}}\!+\!\frac{1}{2m_{\Phi}}\!\sum\limits_{i=1}^{2}\!{{{\left|\left(\!-i\hbar{\partial}_{i}\!-\!2{{\Phi}_{0}}{{\tilde{A}}_{i}}\right)\tilde{\psi}\right|}^{2}}}\right.

(92)

where m2Φ{{m}_{2\Phi}} is the effective mass of the flux pair (vortex pair in the same rotation direction = superinsulator) and ψ~\tilde{\psi} is the wave function of the flux pair. On the other hand, the Ginzburg–Landau free energy for Eq.(5) is as follows: FGL(ψ,ψ,Ai)dx0d2x{f~ab2μ(×𝐀)a(×𝐀)b+12mci=12|(iiΦ0Ai)ψ~|2\displaystyle F\!_{G\!L}(\psi,{\psi}^{*},A_{i})\!\equiv\!\!\int\!\!{d{x_{0}}}\!\!\int\!\!{{d^{2}}x}\!\left\{\frac{{\tilde{f}}^{ab}}{2\mu}{{({\partial^{\prime}}\times\mathbf{A})}_{a}}{{({\partial^{\prime}}\times\mathbf{A})}_{b}}\!+\!\frac{1}{2m_{c}}\sum\limits_{i=1}^{2}{{{\left|\left(-i\hbar{\partial}_{i}-{{\Phi}_{0}}{A_{i}}\right)\tilde{\psi}\right|}^{2}}}\right.
+γ2mc|(i0Φ0A0)ψ~|2+α~ε|ψ|2+β~|ψ|4}\displaystyle\left.\quad\quad\quad\quad\quad\quad\quad+\frac{\gamma}{2m_{c}}{{\left|\left(-i\hbar{\partial}_{0}-{\Phi}_{0}{A}_{0}\right)\tilde{\psi}\right|}^{2}}+\tilde{\alpha}\varepsilon{{|{\psi}|}^{2}}+\tilde{\beta}{{|\psi|}^{4}}\right\}

εEJMFEJEJ,\displaystyle\scalebox{0.8}{$\displaystyle\varepsilon\equiv\frac{{E^{\prime}}_{J}^{M\!F}-{E^{\prime}}_{J}}{{E^{\prime}}_{J}}$}, (93)

where mcm_{c} is the effective mass of the Cooper pair and ψ\psi is the wave function of the Cooper pair. It is known that the microscopic theory of superconductivity with respect to the Ginzburg–Landau theory can be described by the following BCS Hamiltonian (HBCSH_{B\!C\!S}):

|g|2σ,σΩd3xφσ+(x)φσ+(x)φσ(x)φσ(x),\displaystyle\scalebox{0.8}{$\displaystyle-\frac{\left|g\right|}{2}\!\sum\limits_{\sigma,{\sigma^{\prime}}}\!{\int\limits_{\Omega}{\!{d^{3}}x}}\varphi_{\sigma}^{+}\!\left(x\right)\varphi_{\sigma^{\prime}}^{+}\!\left(x\right)\varphi_{\sigma^{\prime}}\!\left(x\right)\varphi_{\sigma}\!\left(x\right)$},\quad (94)

where me{m_{e}} is the effective mass of the electron, and φσ(x){{\varphi}_{\sigma}}\left(x\right) is the electron field having a spin subscript σ\sigma and is a fermion satisfying the anti-commutation relation

{φσ(x),φσ+(x)}+=δσσδ3(xx).{{\left\{{{\varphi}_{\sigma}}\left(x\right),\varphi_{{\sigma^{\prime}}}^{+}\left(x^{\prime}\right)\right\}}_{+}}={{\delta}_{\sigma{\sigma^{\prime}}}}{{\delta}^{3}}\left(x-x^{\prime}\right).

From Eqs.(93) and (8), in microscopic theory, the matter field φσ{\varphi}_{\sigma}is a fermion field which is coupled to the gauge field by an elementary charge ee. On the other hand, in the Ginzburg–Landau theory, the material field ψ~\tilde{\psi} is a boson field which is coupled to the gauge field by 2e2e. The magnetic flux quantum in the superconductor have a repulsive force, form a vortex lattice, and never intersect with each other, so they can be regarded as elementary excitations of fermions.Therefore, from the analogy between the BCS theory and the Ginzburg–Landau theory of superconductivity, it is expected that the microscopic theory of superinsulators with respect to the DGL theory in Eq.(92) can be described by the following dual BCS Hamiltonian:

|g~|2σ~,σ~Ωd3xφ~σ~+(x)φ~σ~+(x)φ~σ~(x)φ~σ~(x),\displaystyle\scalebox{0.8}{$\displaystyle-\frac{\left|\tilde{g}\right|}{2}\!\sum\limits_{\tilde{\sigma},{\tilde{\sigma}^{\prime}}}\!{\int\limits_{\Omega}{\!{d^{3}}x}}\tilde{\varphi}_{\tilde{\sigma}}^{+}\!\left(x\right)\tilde{\varphi}_{\tilde{\sigma}^{\prime}}^{+}\!\left(x\right)\tilde{\varphi}_{\tilde{\sigma}^{\prime}}\!\left(x\right)\tilde{\varphi}_{\tilde{\sigma}}\!\left(x\right)$},\quad (95)

where mΦ{m_{\Phi}} is the magnetic flux (vortex) quantum, and φ~σ{{\tilde{\varphi}}_{\sigma}} is the magnetic flux quantum field having the pseudo-spin subscript σ~\tilde{\sigma}, and is a fermion satisfying the anti-commutation relationship

{φ~σ~(x),φ~σ~+(x)}+=δσ~σ~δ3(xx).{{\left\{{{\tilde{\varphi}}_{\tilde{\sigma}}}\left(x\right),\tilde{\varphi}_{{\tilde{\sigma}^{\prime}}}^{+}\left(x^{\prime}\right)\right\}}_{+}}={{\delta}_{\tilde{\sigma}{\tilde{\sigma}^{\prime}}}}{{\delta}^{3}}\left(x-x^{\prime}\right).

Similar to the relationship between Eqs.(93) and (8), in the microscopic theory of superinsulators of Eq.(8), the material field φ~σ{{\tilde{\varphi}}_{\sigma}} is a fermion field which is coupled to the gauge field by the flux quantum Φ0{{\Phi}_{0}}. On the other hand, in the DGL theory of Eq.(92), the material field ψ~\tilde{\psi} is a boson field and is coupled to the gauge field by 2Φ0{{\Phi}_{0}}. The construction of the dual BCS theory, which is a microscopic theory of superinsulators on nanosheets, not only elucidates the microscopic mechanisms for superinsulators and quantum phase slips, but also the microscopic mechanisms for quark confinement and asymptotic freedom. This theory is expected to play an important role as a powerful model of the QCD phenomenon on desktopDiamantini et al. (2018, 2019).

9 Acknowledgments

I would like to thank all the faculty and staff of Aichi University of Technology.

Appendix A Anisotropic lattice Green’s function 𝒱m(𝟎){\cal{V}}_{m}\left(\mathbf{0}\right) at the source x=0x=0

Perform a numerical evaluation of the anisotropic massive lattice Green’s function (lattice potential)𝒱m(𝟎){{\cal{V}}_{m}}\left(\mathbf{0}\right) at the origin x=0x=0.

𝒱m(𝟎)1gμν¯μν+m2=n=0hn[m2+ 2d~]n+1,\displaystyle\scalebox{1.0}{${\cal{V}}_{m}\left(\mathbf{0}\right)\!\equiv\!\frac{1}{-g^{\mu\nu}{\bar{\nabla}}_{\mu}{\nabla}_{\nu}+m^{2}}=\sum\limits_{n=0}^{\infty}\frac{h_{n}}{{\left[\;m^{2}\;+\;2\tilde{d}\;\right]}^{n+1}}$}, (96)

where, d~2+γ~\tilde{d}\equiv 2+\tilde{\gamma} is the anisotropic dimension, and hnh_{n} is the anisotropic hopping coefficient of the anisotropic massive lattice Green’s function 𝒱m(𝟎){\cal{V}}_{m}\left(\mathbf{0}\right) at the origin x=0x=0, and is introduced as follows:

hnn!j=0,2,4nγ~nj{[(nj)2]!}2j!Hj,\displaystyle\scalebox{1.0}{$h_{n}\equiv n!\sum\limits_{j=0,2,4}^{n}{\frac{{{\tilde{\gamma}}^{n-j}}}{{{\left\{\left[\frac{\left(n-j\right)}{2}\right]!\right\}}^{2}}j!}}{H_{j}}$}, (97)

where HnH_{n} represents the isotropic hopping coefficientsKleinert (1989), for example, in the two-dimensional case, H0=1H_{0}=1, H2=4H_{2}=4, H6=36H_{6}=36, H8=400H_{8}=400,…, and in the three-dimensional case, H0=1H_{0}=1, H2=6H_{2}=6, H6=90H_{6}=90, H8=1860H_{8}=1860,…,. TABLEI\rm{\,I\,} lists examples of γ~=0.1,0.2,.,0.8,0.9\tilde{\gamma}=0.1,0.2,....,0.8,0.9, and 1.01.0.

Table 1: Values of the anisotropic hopping coefficient hnh_{n} for values of n up to 10 for values of the anisotropic parameter γ~=0\tilde{\gamma}=0 to 1 in the case of 2<d~32<\tilde{d}\leq 3, where d~2+γ~\tilde{d}\equiv 2+\tilde{\gamma}.
γ~\tilde{\gamma} h2h_{2} h4h_{4} h6h_{6} h8h_{8} h10h_{10}
  0 4 36 400 4900 63504
  0.1 4.02 36.003 410.83602 5125.514241 67964.55133
  0.2 4.08 36.048 443.77728 5820.335539 81960.10908
  0.3 4.18 36.243 500.13058 7040.109553 107387.453
  0.4 4.32 36.768 582.09792 8880.292915 147593.9992
  0.5 4.5 37.875 692.8125 11480.27344 207665.9648
  0.6 4.72 39.888 836.38912 15029.23717 294849.9426
  0.7 4.98 43.203 1017.98898 19773.88112 419126.4121
  0.8 5.28 48.288 1243.89888 26028.09861 593959.5603
  0.9 5.62 55.683 1521.62482 34184.79254 837254.033
  1.0 6 66 1860 44730 1172556

The asymptotic behavior of the anisotropic massive lattice Green’s function 𝒱m(𝟎){\cal{V}}_{m}\!\left(\mathbf{0}\right) at small values of mm in the case of 2<d32<d\leq 3 is as follows:

Δhnn!j=0,2,4nγ~nj{[(nj)/2]!}2j!Hjγ~1/22d~4π(2n)!22n(n!)2(2d~)n+1n+1,\displaystyle\scalebox{0.7}{$\displaystyle\Delta{h_{n}}\!\equiv\!n!\!\!\!\sum\limits_{j=0,2,4}^{n}\!\!{\frac{{{\tilde{\gamma}}^{n-j}}}{{{\left\{\left[{\left(n-j\right)}/{2}\;\right]!\right\}}^{2}}j!}{H_{j}}}\!-\!\frac{{{\tilde{\gamma}}^{-1/2}}\sqrt{2\tilde{d}}}{4\pi}\frac{\left(2n\right)!}{{2^{2n}}{{\left(n!\right)}^{2}}}\frac{{{\left(2\tilde{d}\right)}^{n+1}}}{n+1}$}, (98)

TABLEII\rm{\,II\,} shows the asymptotic anisotropic hopping coefficient Δhn\Delta{h_{n}} for values of nn up to 10 for values of the anisotropic parameter γ~=0\tilde{\gamma}=0 to 1 in the case of 2<d32<d\leq 3.

Table 2: Asymptotic anisotropic hopping coefficient Δhn\Delta{h_{n}} for values of n up to 10 for values of the anisotropic parameter γ~=0\tilde{\gamma}=0 to 1 in the case of 2<d~32<\tilde{d}\leq 3, where d~2+γ~\tilde{d}\equiv 2+\tilde{\gamma}.
γ~\tilde{\gamma} Δh2\Delta{h_{2}} Δh4\Delta{h_{4}} Δh6\Delta{h_{6}} Δh8\Delta{h_{8}} Δh10\Delta{h_{10}}
  0.1 -1.16603 -0.7561 -0.3789 27.6816 549.214
  0.2 -0.64231 0.10562 4.26659 59.7304 786.122
  0.3 -0.4334 0.38867 5.27033 62.4803 769.857
  0.4 -0.32319 0.50921 5.42088 60.563 744.348
  0.5 -0.25823 0.56803 5.36908 59.2432 755.792
  0.6 -0.2182 0.60248 5.34728 60.225 818.965
  0.7 -0.19353 0.6296 5.46036 64.2966 944.418
  0.8 -0.17904 0.65818 5.7662 72.0562 1145.97
  0.9 -0.17169 0.69306 6.30443 84.1746 1443.87
  1.0 -0.16955 0.73705 7.10848 101.516 1866.95

From Eq.(A) and TABLEII\rm{\,II\,}, when the value of the anisotropic massless lattice Green’s function 𝒱0(𝟎){\cal{V}}_{0}\left(\mathbf{0}\right) at the origin x=0x\!=\!0 is evaluated as the sum of the power-series up to n=10n\!=\!10, it becomes as shown in TABLEIII\rm{\,III\,} and Figure 6:

Table 3: Asymptotic massless lattice Green’s function 𝒱0(𝟎){\cal{V}}_{0}\left(\mathbf{0}\right) at the origin x=0x\!=\!0 for values of nn up to 10 for values of the anisotropic parameter γ~=0\tilde{\gamma}=0 to 1 in the case of 2<d32<d\leq 3, where d~2+γ~\tilde{d}\equiv 2+\tilde{\gamma}.
γ~\tilde{\gamma} 𝒱0(𝟎){\cal{V}}_{0}\left(\mathbf{0}\right)
   0.1  0.444983243
   0.2  0.389718363
   0.3  0.355795798
   0.4  0.331151594
   0.5  0.311864992
   0.6  0.29608701
   0.7  0.282786758
   0.8  0.271328982
   0.9  0.261294231
   1.0  0.252390927
Refer to caption
Figure 8: Asymptotic massless lattice Green’s function 𝒱0(𝟎){\cal{V}}_{0}\left(\mathbf{0}\right) at the origin x=0x=0 versus the anisotropic parameter γ~\tilde{\gamma}.

Appendix B Appendix B. Effective energy approach for the QPSJQ\!P\!S\!J model

Consider the order parameter representation of the super insulator by the partition function of Eq.(6). To simplify the problem, we deal with the case where there is no coupling of gauge fields. First, to derive the one-loop effective field theory, the free correlation function in the presence of a non-vanishing background field is shown below:

=ESδ2Fδψ~^a(x1)δψ~^b(x2)|ψ~^=α,ψ~^l(x)α~l,\displaystyle\scalebox{0.9}{$\displaystyle\!=\!{{\left.{{E^{\prime}}_{S}}\frac{{{\delta}^{2}}{F^{\prime}}}{\delta{{\hat{\tilde{\psi}}}_{a}}\left(x_{1}\right)\delta{{\hat{\tilde{\psi}}}_{b}}\left(x_{2}\right)}\right|}_{\hat{\tilde{\psi}}\!=\!\alpha}},\;\left\langle{{{\hat{\tilde{\psi}}}}_{l}}\left(x\right)\right\rangle\equiv{{\tilde{\alpha}}_{l}}$}, (99)

where αl{\alpha}_{l} is an expectation of ψ~^l(x){\hat{\tilde{\psi}}}_{l}\left(x\right) , and, in general, all nn-th order one-particle irreducible graphs involving the vertex functions can be computed as follows:

Vnψ~(x1,x2,xn)ab|a1,a2,an=ESδnFδψ~^a1(x1)δψ~^a2(x2)δψ~^an(xn)|ψ=α,\displaystyle\scalebox{0.75}{$\displaystyle{{\left.V_{n}^{{\tilde{\psi}}}{{\left(x_{1},x_{2}\cdot\cdot\cdot,x_{n}\right)}_{ab}}\right|}_{a_{1},a_{2}\cdot\cdot\cdot,a_{n}}}\!\!\!\!\!\!\!\!\!\!\!=\!\!{{\left.{{E^{\prime}}\!_{S}}\frac{{{\delta}^{n}}{F^{\prime}}}{\delta{{{\hat{\tilde{\psi}}}}_{a_{1}}}\!\left(x_{1}\right)\delta{{{\hat{\tilde{\psi}}}}_{a_{2}}}\!\left(x_{2}\right)\cdot\cdot\cdot\delta{{{\hat{\tilde{\psi}}}}_{a_{n}}}\!\left(x_{n}\right)}\right|}_{\psi=\alpha}}\!\!\!\!\!\!\!\!\!$}, (100)

The 2×22\!\times\!2 matrix Gψ~1(x1,x2)abG_{{\tilde{\psi}}}^{-1}{\!\left(x_{1},x_{2}\right)}_{ab} defined in Eq.(96) is divided into longitudinal Gψ1(x,y)LG_{\psi}^{-1}{\!\left(x,y\right)}^{L} and transverse Gψ1(x,y)TG_{\psi}^{-1}{\!\left(x,y\right)}^{T} parts, which are respectively parallel and orthogonal to the expected value of the field αl{\alpha}_{l} , as follows:

=Gψ~1(x1,x2)LPabL+Gψ~1(x1,x2)TPabT,\displaystyle\scalebox{0.78}{$\displaystyle=G_{{\tilde{\psi}}}^{-1}{{\left(x_{1},x_{2}\right)}^{L}}P_{ab}^{L}+G_{{\tilde{\psi}}}^{-1}{{\left(x_{1},x_{2}\right)}^{T}}P_{ab}^{T}$},\quad\quad\quad
Gψ1(x,y)L12ESd~{δxy2ESd~ηL(1+12d~gμν¯μν)(x,y)},\displaystyle\scalebox{0.78}{$\displaystyle G_{\psi}^{-1}{{\!\left(x,y\right)}^{L}}\!\equiv\!\frac{1}{2{E^{\prime}}\!_{S}\tilde{d}}\!\left\{\!{\delta}_{xy}\!-\!2{E^{\prime}}\!_{S}\tilde{d}{{\eta}_{L}}\!\left(\!1\!+\!\frac{1}{2\tilde{d}}{g^{\mu\nu}}{{{\bar{\nabla}}}_{\mu}}{{\nabla}_{\nu}}\right)\!\!\left(x,y\right)\!\right\}$},
Gψ1(x,y)T12ESd~{δxy2ESd~ηT(1+12d~gμν¯μν)(x,y)},\displaystyle\scalebox{0.78}{$\displaystyle G_{\psi}^{-1}{{\!\left(x,y\right)}^{T}}\!\equiv\!\frac{1}{2{E^{\prime}}\!_{S}\tilde{d}}\!\left\{\!{\delta}_{xy}\!-\!2{E^{\prime}}\!_{S}\tilde{d}{{\eta}_{T}}\!\left(\!1\!+\!\frac{1}{2\tilde{d}}{g^{\mu\nu}}{{{\bar{\nabla}}}_{\mu}}{{\nabla}_{\nu}}\right)\!\!\left(x,y\right)\!\right\}$}, (101)

where ηab{\eta}_{ab} is a 2×22\!\times\!2 matrix; and PabLP_{ab}^{L} and PabTP_{ab}^{T} are longitudinal and transverse projection matrices, respectively:

(102)

By integrating over all quadratic fluctuations in the partition function of Eq.(6), the one-loop effective energy is as follows:

+Trlog[m22d~(1m22d~)12d~gμν¯μν]},\displaystyle\scalebox{0.78}{$\displaystyle\left.+Tr\log\left[\frac{m^{2}}{2\tilde{d}}\!-\!\left(1-\frac{m^{2}}{2\tilde{d}}\right)\frac{1}{2\tilde{d}}{g^{\mu\nu}}{{\bar{\nabla}}_{\mu}}{{\nabla}_{\nu}}\right]\right\}$},
m24d~(2d~)2ES{1[I1(ES)I0(ES)]2},\displaystyle\scalebox{0.78}{$\displaystyle m^{2}\equiv 4\tilde{d}\!-\!{\left(2\tilde{d}\right)}^{2}\!{E^{\prime}}\!_{S}\!\left\{1\!-\!\left[\frac{I_{1}\left({E^{\prime}}_{S}\right)}{I_{0}\left({E^{\prime}}_{S}\right)}\right]^{2}\right\}$}, (103)

where I0(E)I_{0}\left(E\right) and I1(E)I_{1}\left(E\right) are order zero and order one the modified Bessel functions, respectively. The first trace means zero mass fluctuations (Goldstone modes) and the second trace represents massive fluctuations and can be calculated by hopping expansion as follows:

=n=21n(1m22d~)nhn(2d~)n,\displaystyle\scalebox{0.78}{$\displaystyle=-\sum\limits_{n=2}^{\infty}{\frac{1}{n}}{{\left(1-\frac{m^{2}}{2\tilde{d}}\right)}^{n}}\frac{h_{n}}{(2\tilde{d})^{n}}$}, (104)

where hnh_{n} is the anisotropic hopping coefficient of the anisotropic massive lattice Green’s function Vm(𝟎)V_{m}\left(\mathbf{0}\right) at the source x=0x=0 shown in Appendix A. The one-loop free energy can be expressed by the hopping expansion of Eq.(B) as follows:

F1loop=12n=0,2,4,hnn[(b~ηLd~)n+(b~ηTd~)n],\displaystyle\scalebox{0.78}{$\displaystyle{F^{\prime}}^{1loop}\!=\!\frac{-1}{2}\!\!\!\sum\limits_{n=0,2,4,...}^{\infty}\!\!{\frac{h_{n}}{n}}\left[{{\left(\frac{\tilde{b}{\eta}_{L}}{\tilde{d}}\right)}^{n}}+{{\left(\frac{\tilde{b}{\eta}_{T}}{\tilde{d}}\right)}^{n}}\right]$}, (105)

where b~ESd~\tilde{b}\equiv{E^{\prime}}\!_{S}\tilde{d}. Next, from the one-particle irreducible diagrams \bigcirc\!\bigcirc and \ominus , the free energy of the two loop corrections is derived as follows, respectively:

×(12G^¯2(0)+6[TrG^(0)]G^¯(0))+16Q˙˙˙˙(|α~|)3[G^¯(0)]2},\displaystyle\scalebox{0.78}{$\displaystyle\left.\times\left(12{\bar{\hat{G}}}^{2}\!\!\left(0\right)+6\left[\mathrm{Tr}\hat{G}\left(0\right)\right]\bar{\hat{G}}\left(0\right)\right)+16\ddddot{Q}\!\left(|\tilde{\alpha}|\right)3{{\left[\bar{\hat{G}}\left(0\right)\right]}^{2}}\right\}$}, (106)
+36(4Q¨)(8Q˙˙˙)[G^¯2(x)]G^¯(x)+6(8Q˙˙˙)2[G^¯(x)]3},\displaystyle\scalebox{0.78}{$\displaystyle\left.\!+36\!\left(4\ddot{Q}\right)\!\!\Bigl{(}8\dddot{Q}\Bigr{)}\!\!\left[{\bar{\hat{G}}}^{2}\!\!\!\left(x\right)\right]\!\bar{\hat{G}}\!\left(x\right)\!+\!6\left(8\dddot{Q}\right)^{2}\!{{\left[\bar{\hat{G}}\!\left(x\right)\right]}^{3}}\right\}$}, (107)

where the dotted accent for Q˙(|α~|)\dot{Q}\left(\left|\tilde{\alpha}\right|\right) is defined by the modified derivative i.e., Q˙(1/|2α~|)dQ/d|α~|\dot{Q}\!\equiv\!(\!1/|2\tilde{\alpha}|)dQ/{d\left|\tilde{\alpha}\right|}, and G^(x)ab\hat{G}{{\left(x\right)}_{ab}} is defined as:

G^¯(x)ψaG^(x)abψb,TrG^(𝟎)lG^(𝟎)ll,\displaystyle\scalebox{0.76}{$\displaystyle\bar{\hat{G}}\left(x\right)\equiv{\psi}_{a}\hat{G}{{\left(x\right)}_{ab}}{\psi}_{b},\;\mathrm{Tr}\hat{G}\left(\mathbf{0}\right)\equiv\sum\limits_{l}{\hat{G}{{\left(\mathbf{0}\right)}_{ll}}}$},\quad\quad (108)

where the trace refers only to the index of the 2×22\times 2 matrix G^(x)ab\hat{G}{{\left(x\right)}_{ab}}. To calculate the free energy of the two-loop correction, introduce the hopping expansion of G^(x)ab\hat{G}{{\left(x\right)}_{ab}}, as follows:

h(x,y)2d~δxy+gμν¯μν,,\displaystyle\scalebox{0.82}{$\displaystyle h\left(x,y\right)\equiv 2\tilde{d}{{\delta}_{xy}}+{g^{\mu\nu}}{{\bar{\nabla}}_{\mu}}{{\nabla}_{\nu}},$},\quad\quad (109)

For b~d~\tilde{b}\ll\tilde{d}, i.e., at the limit of small ES{{E^{\prime}}\!_{S}}, the free energy due to the mean field approximation with up to the two loop corrections for the up to order b~4{{\tilde{b}}^{4}} is as follows:

+1ψ~(3η˙TηT24ηLηTη˙T+2ηLη˙LηT)+ηL2η¨L}+0(b~5),\displaystyle\scalebox{0.7}{$\displaystyle\left.+\frac{1}{{\tilde{\psi}}}\left(3{{{\dot{\eta}}}_{T}}\eta_{T}^{2}-4{{\eta}_{L}}{{\eta}_{T}}{{{\dot{\eta}}}_{T}}+2{{\eta}_{L}}{{\dot{\eta}}_{L}}{{\eta}_{T}}\right)+\eta_{L}^{2}{{{\ddot{\eta}}}_{L}}\right\}+0\left({{{\tilde{b}}}^{5}}\right)$},\quad\quad
h^22(γ~2γ~),h^46γ~4+(24D12)γ~2+(624D)γ~,\displaystyle\scalebox{0.7}{$\displaystyle{{\hat{h}}_{2}}\equiv 2\left({{{\tilde{\gamma}}}^{2}}-\tilde{\gamma}\right),\;{{\hat{h}}_{4}}\equiv 6{{\tilde{\gamma}}^{4}}+\left(24D-12\right){{\tilde{\gamma}}^{2}}+\left(6-24D\right)\tilde{\gamma}$},\quad (110)

The result of finding the minimum value ψ~0{{\tilde{\psi}}_{0}} of ψ~\tilde{\psi} according to Eq.(B) is as follows:

ψ~0=8(11/b~)+32Δ2164Δ1,\displaystyle\scalebox{0.78}{$\displaystyle{{\tilde{\psi}}_{0}}=\sqrt{\frac{8\left(1-{1}/{{\tilde{b}}}\;\right)+32{{\Delta}_{2}}}{1-64{{\Delta}_{1}}}}$},\quad\quad\quad\quad\quad\quad
Δ1b~4(580d~2612d~+51h^4)+68b~2d~2(2d~+h^2)4096d~4,\displaystyle\scalebox{0.78}{$\displaystyle{{\Delta}_{1}}\equiv\frac{{{{\tilde{b}}}^{4}}\left(580{{{\tilde{d}}}^{2}}-612\tilde{d}+51{{{\hat{h}}}_{4}}\right)+68{{{\tilde{b}}}^{2}{{{\tilde{d}}}^{2}}}\left(2\tilde{d}+{{{\hat{h}}}_{2}}\right)}{4096{{{\tilde{d}}}^{4}}}$},\quad
Δ238b~4d~260b~3d~2(2d~+h^2)3b~5(34d~260d~+5h^4)768d~4b~,\displaystyle\scalebox{0.78}{$\displaystyle{{\Delta}_{2}}\equiv\frac{38{{{\tilde{b}}}^{4}}{{{\tilde{d}}}^{2}}-60{{{\tilde{b}}}^{3}}{{{\tilde{d}}}^{2}}\left(2\tilde{d}+{{{\hat{h}}}_{2}}\right)-3{{{\tilde{b}}}^{5}}\left(34{{{\tilde{d}}}^{2}}-60\tilde{d}+5{{\hat{h}}_{4}}\right)}{768{{\tilde{d}}^{4}}\tilde{b}}$},\; (111)

Therefore, for the mean field approximation with up to the two loop corrections of ES{E^{\prime}}\!_{S}, the critical point (ES)c2loop\left({E^{\prime}}\!_{S}\right)_{c}^{2\text{loop}} and the triple critical point (ES)tri{{\left({E^{\prime}}\!_{S}\right)}^{t\!r\!i}} are respectively as follows:

..+4D(60γ~3+120γ~258γ~45)+2γ(105γ~358γ~45)]/192(D+γ~)4}1,\displaystyle\scalebox{0.62}{$\displaystyle\Bigl{.}\Bigl{.}+4D\left(60{{{\tilde{\gamma}}}^{3}}+120{{{\tilde{\gamma}}}^{2}}-58\tilde{\gamma}-45\right)+2\gamma\left(105{{{\tilde{\gamma}}}^{3}}-58\tilde{\gamma}-45\right)\Bigr{]}/{192{{\left(D+\tilde{\gamma}\right)}^{4}}}\Bigr{\}}^{-1}$}, (112)
(ES)tri=1156(2d~+h^2)2+64(580d~2612d~+51h^4)34(2d~+h^2)580d~2612d~+51h^4,\displaystyle\scalebox{0.65}{$\displaystyle{\left({E^{\prime}}\!_{S}\right)}^{t\!r\!i}\!=\!\sqrt{\frac{\sqrt{1156{{\left(2\tilde{d}+{{\hat{h}}_{2}}\right)}^{2}}\!+\!64\left(580{{\tilde{d}}^{2}}\!-\!612\tilde{d}\!+\!51{{\hat{h}}_{4}}\right)}\!-\!34\left(2\tilde{d}\!+\!{\hat{h}}_{2}\right)}{580{{\tilde{d}}^{2}}\!-\!612\tilde{d}\!+\!51{{\hat{h}}_{4}}}}$}, (113)

References

  • Mooij et al. (2005) J. E. Mooij and C. J. P. M. Harmans, New Journal of Physics 7, 219 (2005).
  • Mooij et al. (2006) J. E. Mooij and Yu. V. Harmans, Nature Physics 2, 169–172 (2006).
  • Mooij (2015) J. E. Mooij, New J. Phys 17, 033006 (2015).
  • Astafiev (2012) O. V. Astafiev, et al, Nature 484, 355–358 (2012).
  • Hooft (1978) G.’ t. Hooft, Nucl. Phys B 138, 1-25 (1978).
  • Diamantini et al. (1996) M. C. Diamantini, P. Sodano, and C. A. Trugenberger, Nucl. Phys. B 474, 641-677 (1996).
  • Doniach et al. (1998) A. Krämer and S. Doniach, Phys. Rev. Lett 81, 3523-3527 (1998).
  • Yoneda et al. (2012) M. Yoneda, M. Niwa, and M. Motohashi, Physica Scripta 2012, T151 (2012).
  • Vinokur et al. (2008) V. M. Vinokur, et al, Nature 452, 613-615 (2008).
  • Ovadia et al. (2015) M. Ovadia, et al, Scientific Reports 5, 13503 (2015).
  • Mironov et al. (2018) M. Mironov, et al, Scientific Reports 8, 4082 (2018).
  • Diamantini et al. (2018) M. C. Diamantini, et al, Nature Comm. Phys 1, 77 (2018).
  • Diamantini et al. (2019) M. C. Diamantini, et al, Journal of Superconductivity and Novel Magnetism 51, 32-47 (2019).
  • Yoneda et al (2019) M. Yoneda, et al., Int. J. Mod. Phys. B Vol. 33, No. 25, 1950291 (2019).
  • Diamantini et al (2017) M. C. Diamantini, et al, Gauge Topological Nature of the Superconductor-Insulator Transition. arXiv:1710.10575.
  • Diamantini et al. (2019) M. C. Diamantini, et al, Bosonic topological insulator intermediate state in the superconductor-insulator transition. arXiv:1906.07969.
  • Hollen et al. (2014) S. M Hollen, et al, Phys. Rev. B 90, 140506(R) (2014).
  • Brayden Ware et al (2015) Brayden Ware, et al, Phys. Rev. B 92, 195105 (2015).
  • Zhang and Schilling (2018) X. Zhang, and A. Schilling, Phys. Rev. B 97, 214524 (2018).
  • Das and Doniach (1999) D. Das,and S. Doniach, Phys. Rev. B 60, 1261 (1999).
  • Das and Doniach (2001) D. Das,and S. Doniach, Phys. Rev. B 64, 134511 (2001).
  • Phillips, and Dalidovich (2003) P. Phillips,and D. Dalidovich, SCIENCE 302, 243-247 (2003).
  • Wu and Phillips (2006) J. Wu,and P. Phillips, Phys. Rev. B 73, 214507 (2006).
  • Phillips (2019) P. Phillips, SCIENCE 366, 1450-1451 (2019).
  • Kleinert (1989) H. Kleinert, Gauge fields in condensed matter. Vol. 1: Superflow and vortex lines. Disorder fields, phase transitions (World Scientific, Singapore, 1989).
  • Villain (1975) J. Villain, J. de Phys 36, 581 (1975).
  • Janke and Kleinert (1986) W. Janke and H. Kleinert, Nuclear Physics B 270, 135-153 (1986).
  • Kleinert (1985) H. Kleinert, Int. J. Engng. Sci. 23, 927-935 (1985).
  • Herbut (2007) I. Herbut, A Modern Approach to Critical Phenomena. :7.1 frozen lattice superconductor (Cambridge University Press, 2007).
  • Neuhaus et al. (2003) T. Neuhaus, A. Rajantie, and K. Rummukainen, Nuclear Physics B 119, 915-917 (2003).
  • Neuhaus et al. (2003) T. Neuhaus, A. Rajantie, and K. Rummukainen, Phys. Rev. B 67, 014525 (2003)