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

thanks: ychen@gscaep.ac.cn

A Hydrodynamical Description of Bose Liquid with Fractional Exclusion Statistics

Zhaoyu Fei1    Yu Chen1 1Graduate School of China Academy of Engineering Physics, Beijing 100193, China
Abstract

Hydrodynamical systems are usually taken as chaotic systems with fast relaxations. It is counter intuitive for “ideal” gas to have a hydrodynamical description. We find that a hydrodynamical model of one-dimensional |Φ|6|\Phi|^{6} theory shares the same ground state density profile, density-wave excitation, as well as the similar dynamical and statistical properties with the Calogero-Sutherland model in thermodynamic limit when their interaction strengths matches each other. The interaction strength g0g_{0} in |Φ|6|\Phi|^{6} theory is then the index of fractional statistics. Although the model is interacting in Bose liquid sense, but it shows integrability with periodical coherent evolution. We also discussed the fractional statistics emerges from the |Φ|6|\Phi|^{6} theory.

I Introduction

In conventional wisdom of hydrodynamics, the existence of the hydrodynamical equations are based on the rapid relaxation of the system Landau . Usually this assumption is well satisfied in chaotic systems such like the interacting Bose gases. However, there are some interesting exceptions, such as the SGZ solutions which shows integrable motions in hydrodynamics shi2021 , where it shows hydrodynamics equations can also emerge for “ideal” systems. These integrability are shown in the examples of two-dimensional superfluids paris ; qi20 . On the other hand, we also notice that the ideal anyons satisfying fractional exclusion statistics which can be described by interacting bosons or fermions  Haldane ; wu1994 . Naturally, one question is raised on whether the dynamics of ideal anyons can be treated by specific integrable hydrodynamical theory. In this article, we are going to explore this problem.

The Calogero-Sutherland (CS) model csm1969 ; sutherland246 describes trapped particles with 1/r21/r^{2} long-range two-body interactions,

H^CS=j=1N(12xj2+12ωcs2xj2)+i<jgcs(gcs1)|xixj|2,\displaystyle\hat{H}_{\rm CS}=\sum_{j=1}^{N}\left(\frac{-1}{2}\partial_{x_{j}}^{2}+\frac{1}{2}\omega_{\rm cs}^{2}x_{j}^{2}\right)+\sum_{i<j}\frac{g_{\rm cs}(g_{\rm cs}-1)}{|x_{i}-x_{j}|^{2}}, (1)

where xj=/xj\partial_{x_{j}}=\partial/\partial x_{j}, NN is the total particle number, ωcs\omega_{\rm cs} the trap frequency and gcsg_{\rm cs} a dimensionless interaction strength. The particle mass mm and \hbar are taken as 1. This model is exactly solvable and one can prove that the excitations of CS model obey fractional exclusion statistics murthy1994 . Such statistics was first introduced by Haldane, and Yongshi Wu Haldane ; wu1994 . Here, we stress that the parameter gcsg_{\rm cs} is the index labelling the fractional statistics. When gcs=0g_{\rm cs}=0, the system is ideal Bosons. And when gcs=1g_{\rm cs}=1, the system is impenetrable Bosons, which can be mapped to spinless ideal Fermions in terms of the theorems in girardeau1960 .

We find the one-dimensional |Φ|6|\Phi|^{6} Bose liquid theory shares the same ground state density profile, density-wave excitation as well as the similar dynamical and statistical properties as them of the CS model. The trapped |Φ|6|\Phi|^{6} theory can be given as

H=12+dx[|Φx|2+ω02x2|Φ|2+g02π23|Φ|6],\displaystyle H=\frac{1}{2}\int_{-\infty}^{+\infty}\!\!\!\!\!\!\!\mathrm{d}x\left[\left|\frac{\partial\Phi}{\partial x}\right|^{2}\!\!\!+\omega_{0}^{2}x^{2}|\Phi|^{2}+\frac{g_{0}^{2}\pi^{2}}{3}|\Phi|^{6}\right]\!\!, (2)

where Φ(x)\Phi(x) is a one-dimensional complex field. ω0\omega_{0} is the trap frequency and g0g_{0} is the three-body contact interaction strength. Remarkably, we find g0g_{0} and ω0\omega_{0} matches gcsg_{\rm cs} and ωcs\omega_{\rm cs} exactly. More precisely, it means that both the ground state density profile of the CS model and it of the |Φ|6|\Phi|^{6} model share the same Wigner semicircle distribution in the thermodynamic limit wigner1955 . Here, the thermodynamic limit means that N,ω0N\to\infty,\omega\to 0 while keep NωN\omega as a constant dalfovo1999 . The excitation frequencies are equal spacing. The period of the density wave excitation is interaction strength independent and trap frequency dependent only. The direct consequence of this property is the coherence in time evolution. We also find that the |Φ|6|\Phi|^{6} theory satisfies the fractional exclusion statistics whose statistical parameter is g0g_{0} corresponding to gcsg_{\rm cs} in the CS model. It is worth mentioning that the connection between the two models has been implicitly revealed by using the collective-field method Andric1988 ; Sen1997 .

In the following, we will present the correspondence between the CS model and trapped |Φ|6|\Phi|^{6} theory from the dynamical symmetry, ground state profile, excitation spectra, dynamical behavior, and statistical properties respectively. These correspondence are verified analytically in the thermodynamic limit and numerically checked for a finite-NN case.

II The Correspondence Between |Φ|6|\Phi|^{6} theory and the Calogero-Sutherland model

II.1 SU(1,1)\mathrm{SU}(1,1) dynamical symmetry

Previously, it is discovered that for hamiltonian of the form as H^=12ii2+V(x1,,xN)+12iω02xi2\hat{H}=-\frac{1}{2}\sum_{i}\partial_{i}^{2}+V(x_{1},\cdots,x_{N})+\frac{1}{2}\sum_{i}\omega_{0}^{2}x_{i}^{2}, the system has SU(1,1)SU(1,1) dynamical symmetry as long as V(x1,,xN)V(x_{1},\cdots,x_{N}) satisfying V(λx1,,λxN)=λ2V(x1,,xN)V(\lambda x_{1},\cdots,\lambda x_{N})=\lambda^{-2}V(x_{1},\cdots,x_{N}). A simple review of this symmetry can be checked by the commutative relation among H^0=12ii2+V(x1,,xN)\hat{H}_{0}=-\frac{1}{2}\sum_{i}\partial_{i}^{2}+V(x_{1},\cdots,x_{N}), I^=12iω02xi2\hat{I}=\frac{1}{2}\sum_{i}\omega^{2}_{0}x_{i}^{2}, and Q^=i2i(ixi+xii)\hat{Q}=-\frac{i}{2}\sum_{i}(\partial_{i}x_{i}+x_{i}\partial_{i}), i. e.,

[Q^,H^0]=2iH^0,[Q^,I^]=2iI^,\displaystyle[\hat{Q},\hat{H}_{0}]=2i\hat{H}_{0},\hskip 17.22217pt[\hat{Q},\hat{I}]=-2i\hat{I},
[I^,H^0]=iω02Q^.\displaystyle[\hat{I},\hat{H}_{0}]=i\omega_{0}^{2}\hat{Q}.

By introducing L1=(H^0I^)/2ω0L_{1}=(\hat{H}_{0}-\hat{I})/2\omega_{0}, L3=H^/2ω0L_{3}=\hat{H}/2\omega_{0}, and L2=Q^/2L_{2}=\hat{Q}/2, we have

[L1,L2]=iL3,[L2,L3]=iL1,[L3,L1]=iL2.\displaystyle[L_{1},L_{2}]=-iL_{3},\hskip 12.91663pt[L_{2},L_{3}]=iL_{1},\hskip 12.91663pt[L_{3},L_{1}]=iL_{2}.

Clearly, these three operators form \mathfraksu(1,1)\mathfrak{su}(1,1) algebra. We can check that both of the CS model and the trapped |Φ|6|\Phi|^{6} theory have this scale property. For the |Φ|6|\Phi|^{6} theory, its microscopic first quantization interaction term is a three-body contact interaction

V(x1,,xN)=g02π2i<j<kδ(xixj)δ(xjxk).\displaystyle V(x_{1},\cdots,x_{N})=g_{0}^{2}\pi^{2}\sum_{i<j<k}\delta(x_{i}-x_{j})\delta(x_{j}-x_{k}). (3)

The dynamical symmetry is easily checked by noticing δ(λx)=δ(x)/|λ|\delta(\lambda x)=\delta(x)/|\lambda|, which is also inherited by the trapped |Φ|6|\Phi|^{6} theory.

Refer to caption
Figure 1: In (a), (b) and (c) we show the ground state density profile of CS model (solid line, short for CS), |Φ|6|\Phi|^{6} theory (dashed line, short for |Φ|6|\Phi|^{6}) and Wigner semicricle distribution (Gray line, short for W) for N=6N=6. g=gcs=g0=0.5g=g_{\rm cs}=g_{0}=0.5, g=gcs=g0=1.0g=g_{\rm cs}=g_{0}=1.0 and g=gcs=g0=1.5g=g_{\rm cs}=g_{0}=1.5 in (a), (b) and (c) respectively. In (d), (e) and (f), we show how the ground state density profile of the |Φ|6|\Phi|^{6} theory approaches Wigner semicircle distribution in the thermodynamic limit. The interaction strength is taken as 0.50.5 in (d), 1.01.0 in (e) and 1.51.5 in (f). N=6,12,18N=6,12,18 respectively. For comparison, the ground state density profile for N=12N=12 and N=18N=18 are rescaled such that their radius R=2g0N/ω0R=\sqrt{2g_{0}N/\omega_{0}} and the area under the curve are the same as them in the N=6N=6 case.

The direct result for this dynamical symmetry is a 2ω02\omega_{0} frequency breathing mode in the trap pitaev1997 . We also stress that the symmetry property is correct for any particle number NN. But the symmetry alone can not ensure the correspondence between the two systems. Actually we can see the deviation between these two theories at large g0=gcsg_{0}=g_{\rm cs} with a finite NN, where the long-range interaction in the CS model induce a Wigner crystal wigner while this density order is clearly absent in the |Φ|6|\Phi|^{6} theory.

II.2 ground state density profile

In this section, we first verify that the ground state density profile in the CS model and trapped |Φ|6|\Phi|^{6} theory are identical in the thermodynamic limit. Then, for a finite NN, we present the numerical results of the ground state density ρgCS(x)=|Ψg(x)|2\rho^{\rm CS}_{\rm g}(x)=|\Psi_{\rm g}(x)|^{2}, ρg(x)\rho_{\rm g}(x) for the CS model and the |Φ|6|\Phi|^{6} theory respectively.

In the |Φ|6|\Phi|^{6} theory, the ground state density profile can be obtained by an equivalent hydrodynamical description, where the density and velocity field obey the following equations,

tρ+x(ρv)=0,\displaystyle\frac{\partial}{\partial t}\rho+\frac{\partial}{\partial x}(\rho v)=0, (4)
vt+x(12v2+12ω02x2+g02π2ρ22+PQ)=0.\displaystyle\frac{\partial v}{\partial t}\!+\!\frac{\partial}{\partial x}\left(\frac{1}{2}v^{2}\!+\!\frac{1}{2}\omega_{0}^{2}x^{2}\!+\!\frac{g_{0}^{2}\pi^{2}\rho^{2}}{2}\!+\!P_{Q}\!\right)\!=0. (5)

Here, ρ=ρ(x,t)|Φ(x,t)|2\rho=\rho(x,t)\equiv|\Phi(x,t)|^{2} is the density field, and v=v(x,t)(ΦxΦxΦΦ)/2iρ(x,t)v=v(x,t)\equiv\left(\Phi^{*}\partial_{x}\Phi-\partial_{x}\Phi^{*}\Phi\right)/2i\rho(x,t) is the velocity field. PQ=ρ(x,t)x2ρ(x,t)/2P_{Q}=-\sqrt{\rho(x,t)}\partial^{2}_{x}\sqrt{\rho(x,t)}/2 is the so-called quantum pressure. When the trap varies slowly on the scale of the inter-particle spacing, the density profile ρ\rho becomes smooth and the quantum pressure PQP_{Q} can be neglected, which is called the Thomas-Fermi (TF) approximation. Because in ground state, the velocity field is zero, tρ=0\partial_{t}\rho=0, then it requires

x(ω02x2+g02π2ρ2)=0.\displaystyle\frac{\partial}{\partial x}\left(\omega_{0}^{2}x^{2}+g_{0}^{2}\pi^{2}\rho^{2}\right)=0. (6)

Therefore, the ground state density profile is

ρg(x)=ω0πg02g0Nω0x2θ(2g0Nω0|x|),\displaystyle\rho_{\rm g}(x)=\frac{\omega_{0}}{\pi g_{0}}\sqrt{\frac{2g_{0}N}{\omega_{0}}-x^{2}}\theta\left(\sqrt{\frac{2g_{0}N}{\omega_{0}}}-|x|\right), (7)

where N=dxρ(x)N=\int{\mathrm{d}x}\rho(x) is the total particle number, θ(x)\theta(x) denotes the Heaviside step function. When g0=1g_{0}=1, such an equation is the ground state profile of spinless ideal fermions in a harmonic trap, which has been studied in Ref. eugene2000 Although this result is obtained under TF approximation, we see the semicircle law is quickly reached for finite NN with our numerical simulations.

Next, we solve the ground state density profile for the CS model. The expression of the ground state of the CS model is given as

Ψg({xj})=C1i<jN|xixj|gcseωcs2jxj2,\displaystyle\Psi_{\rm g}(\{x_{j}\})=\sqrt{C}\prod_{1\leq i<j\leq N}|x_{i}-x_{j}|^{g_{\rm cs}}e^{-\frac{\omega_{\rm cs}}{2}\sum_{j}x_{j}^{2}}, (8)

with the ground state energy Eg=ωcs[N/2+gcsN(N1)/2]E_{g}=\omega_{\rm cs}[N/2+g_{\rm cs}N(N-1)/2], where xjx_{j} are the positions of particles, and CC is the normalization factor, whose inverse is

C1=(ωcs2)Eg/ωcs(π2)N/2j=1NΓ(1+jgcs)Γ(1+gcs).\displaystyle C^{-1}=\left(\frac{\omega_{\rm cs}}{2}\right)^{-E_{g}/\omega_{\rm cs}}\left(\frac{\pi}{2}\right)^{N/2}\prod_{j=1}^{N}\frac{\Gamma(1+jg_{\rm cs})}{\Gamma(1+g_{\rm cs})}. (9)

The density profile is obtained by ρgCS(x)=j=2Ndxj|Ψg(x,x2,,xN)|2\rho^{\rm CS}_{\rm g}(x)=\int\prod_{j=2}^{N}\mathrm{d}x_{j}|\Psi_{g}(x,x_{2},\cdots,x_{N})|^{2}. As the wave function itself is just the joint probability of the eigenvalues for random Gaussian matrices ensemble, the density function can be viewed as the charge distribution of log-potential Coulomb gas forrester2010 . In the thermodynamic limit, the solution of the density profile is obtained

ρgCS(x)=ωcsπgcs2gcsNωcsx2θ(2gcsNωcs|x|).\displaystyle\rho^{\rm CS}_{\rm g}(x)=\frac{\omega_{\rm cs}}{\pi g_{\rm cs}}\sqrt{\frac{2g_{\rm cs}N}{\omega_{\rm cs}}-x^{2}}\theta\left(\sqrt{\frac{2g_{\rm cs}N}{\omega_{\rm cs}}}-|x|\right). (10)

By comparing Eq. (7) and Eq. (10), we find that these two density profiles are exactly the same when we match the frequencies and the interaction strength. It is worth mentioning that ρg(x)\rho_{\rm g}(x) also obtained from the zero-temperature distribution of the occupation number n(ϵ)n(\epsilon) for anyons satisfying fractional exclusion statistics murthy1994 :

n(ϵ)=g01θ(g0Nω0ϵ)\displaystyle\begin{split}n(\epsilon)=g_{0}^{-1}\theta(g_{0}N\omega_{0}-\epsilon)\end{split} (11)

with a semiclassical single-particle energy ϵ=p2/2+ω02x2/2\epsilon=p^{2}/2+\omega_{0}^{2}x^{2}/2. Then, ρg(x)\rho_{\rm g}(x) is given by ρg(x)=(2π)1dpn(ϵ)\rho_{\rm g}(x)=(2\pi)^{-1}\int\mathrm{d}pn(\epsilon).

In the following, we are going to show the numerical results for ground state density profile for the CS model and for the |Φ|6|\Phi|^{6} theory. Here and in the following figures, we set ωcs=ω0=1\omega_{\rm cs}=\omega_{0}=1. For the CS model, we use the Monte-Carlo integration. Here we keep x1=xx_{1}=x, and we reparametrize xi=2x_{i=2} to xi=Nx_{i=N} on a (N1)(N-1)-sphere i=2Nxi2=r2\sum_{i=2}^{N}x_{i}^{2}=r^{2}. The measure can be then written as j=2Ndxi=rN2drdΩN1\prod_{j=2}^{N}\mathrm{d}x_{i}=r^{N-2}\mathrm{d}r\mathrm{d}\Omega_{N-1} with ΩN1\Omega_{N-1} denoting the solid angle of the (N1)(N-1)-sphere. The ground state density profile is then expressed as

ρgCS(x)\displaystyle\rho_{\rm g}^{\rm CS}(x) =\displaystyle= rN2+gcs(N1)(N2)eωcsr2drdΩN1\displaystyle\int r^{N-2+g_{\rm cs}(N-1)(N-2)}e^{-\omega_{\rm cs}r^{2}}\mathrm{d}r\int\mathrm{d}\Omega_{N-1}
j=2N|xxj(r,Θ)|2gcs2i<jN|xi(r,Θ)xj(r,Θ)r|2gcs.\displaystyle\prod_{j=2}^{N}|x\!-\!x_{j}(r,\Theta)|^{2g_{\rm cs}}\!\!\!\!\!\!\!\prod_{2\leq i<j\leq N}\!\!\!\!\!\!\!\left|\frac{x_{i}(r,\Theta)\!-\!x_{j}(r,\Theta)}{r}\right|^{2g_{\rm cs}}.

Here Θ\Theta represent all the angle parameters of the (N1)(N-1)-sphere. From the expression, the integrant is semi-positive definite, thus the multi-dimensional integral is very suitable for Monte-Carlo integration. Meanwhile, because of the measure and the Gaussian distribution in wave-function, the major contribution of above integration comes from very narrow region of r=0r=0. The concentration of measure in angular part due to Lévy’s Lemma Ledoux can further reduce the complexity in integration. With all these helps, the ground state density profile of CS model for N=3N=3, N=6N=6 can be obtained.

On the other hand, we carry out a numerical calculation of |Φ|6|\Phi|^{6} theory as a non-linear Schrödinger equation. A cutoff in real space is taken as XmaxX_{\rm max}, and a boundary condition is taken as Φ(x=Xmax)=Φ(x=Xmax)\Phi(x=X_{\rm max})=\Phi(x=-X_{\rm max}). To solve the ground state of the |Φ|6|\Phi|^{6} theory, we turn on an evolution in imaginary time, where the ground state is the fixed point of this flow.

The numerical results are summarized in Fig. 1. In Fig. 1(a), (b) and (c), we have shown the ground state density profile of CS model, the |Φ|6|\Phi|^{6} theory and the semicircle law with different interaction strength gcs=g0=0.5g_{\rm cs}=g_{0}=0.5, 11 and 1.51.5. For gcs>1g_{\rm cs}>1, the density wave order becomes prominent. This is the tendency for Wigner crystal because of the long-range nature of r2r^{-2} interaction. For gcs=g0<1g_{\rm cs}=g_{0}<1, the ground state density profile of CS model and the |Φ|6|\Phi|^{6} model is very close to each other even for N=6N=6. At the same time, in Fig. 1(d), (e) and (f), the ground state density profile of the |Φ|6|\Phi|^{6} theory approaches the semicircle distribution as NN increases. Actually, the semicircle law is already a very good approximation for N=6N=6.

II.3 The Correspondence in density-wave excitations and the coherent evolution

In the following, we are going to connect the density-wave excitation of the |Φ|6|\Phi|^{6} theory and the CS model in the thermodynamic limit. By comparison, we find that the density-wave excitations of the two models share the same expression. Then, we check the finite-NN effect by full numerical calculations of the time evolution of the excitations. We analytically prove the correspondence between the density-wave excitations for any excited modes where gcs=g0=1g_{\rm cs}=g_{0}=1, or for the second excited mode where gcs=g0g_{\rm cs}=g_{0} is arbitrary. A guess for the correspondence between the density-wave excitations of the two models for any excited modes and for arbitrary gcs=g0g_{\rm cs}=g_{0} is made and is checked with finite NN. From the numerical results we obtained, the guess seems to be hold in the thermodynamic limit.

Refer to caption
Figure 2: In (a) and (c), we show the time evolution of δρn(x=0,t)\delta\rho_{n}(x=0,t) for n=2n=2 and n=4n=4. The particle number N=40N=40. In (b) and (d), we show the time evolution of δρn(x=0,t)\delta\rho_{n}(x=0,t) for n=2n=2 and n=4n=4. The particle number N=125N=125. All interaction strength are fixed at g0=2g_{0}=\sqrt{2}. The dotdashed lines are nonlinear fittings of the curves. The ansatz is taken as 0.1n=2,4ancos(nω¯n+φn)0.1\sum_{n=2,4}a_{n}\cos(n\bar{\omega}_{n}+\varphi_{n}), a2a_{2}, a4a_{4}, ω¯2\bar{\omega}_{2}, ω¯4\bar{\omega}_{4}, φ2\varphi_{2} and φ4\varphi_{4} are fitting parameters. The fitting parameters are (a) a2=0.587a_{2}=0.587, a4=0.245a_{4}=0.245, ω¯2=1.006\bar{\omega}_{2}=1.006, ω¯4=1.003\bar{\omega}_{4}=1.003, φ2=1.35\varphi_{2}=1.35, φ4=0.465\varphi_{4}=-0.465; (b) a2=0.584a_{2}=-0.584, a4=0.221a_{4}=-0.221, ω¯2=1.00439\bar{\omega}_{2}=1.00439, ω¯4=1.00415\bar{\omega}_{4}=1.00415, φ2=0.099\varphi_{2}=-0.099, φ4=0.202\varphi_{4}=-0.202; (c) a2=0.584a_{2}=-0.584, a4=0.2208a_{4}=-0.2208, ω¯2=1.00439\bar{\omega}_{2}=1.00439, ω¯4=1.00415\bar{\omega}_{4}=1.00415; (d) a2=0.2358a_{2}=0.2358, a4=0.5907a_{4}=0.5907, ω¯2=1.0007\bar{\omega}_{2}=1.0007, ω¯4=1.0006\bar{\omega}_{4}=1.0006.

First of all, let us focus on the |Φ|6|\Phi|^{6} theory. Here our starting point is the hydrodynamical equation Eq. (4) and Eq. (5). Following the procedure in Refs. string1996 ; menotti2002 , we linearize the density ρ=ρ0+δρ\rho=\rho_{0}+\delta\rho (correspondingly v=δvv=\delta v) with a perturbation δρ\delta\rho obeying the linear expansion of Eqs. (4), (5)

2δρt2ω02x[R2x2x(R2x2δρ)]=0.\displaystyle\frac{\partial^{2}\delta\rho}{\partial t^{2}}-\omega_{0}^{2}\frac{\partial}{\partial x}\left[\sqrt{R^{2}-x^{2}}\frac{\partial}{\partial x}(\sqrt{R^{2}-x^{2}}\delta\rho)\right]=0. (13)

where R=2g0N/ω0R=\sqrt{2g_{0}N/\omega_{0}} is the radius of the semicircle distribution in the thermodynamic limit. The boundary conditions come from the absence of the particle flow at x=±Rx=\pm R imam2006

ρ0δv|x=±R=0.\displaystyle\rho_{0}\delta v|_{x=\pm R}=0. (14)

Then, by performing the transformation ϑ=arccos(x/R)\vartheta=\arccos(x/R) (0ϑπ0\leq\vartheta\leq\pi), δρ~=R2x2δρ,δv~=ρ0δv\delta\tilde{\rho}=\sqrt{R^{2}-x^{2}}\delta\rho,\delta\tilde{v}=\rho_{0}\delta v, we obtain a standard wave function

2δρ~t2ω22δρ~ϑ2=0,\displaystyle\frac{\partial^{2}\delta\tilde{\rho}}{\partial t^{2}}-\omega^{2}\frac{\partial^{2}\delta\tilde{\rho}}{\partial\vartheta^{2}}=0, (15)

with the Neumann boundary condition

δρ~ϑ|ϑ=0,π=0.\displaystyle\left.\frac{\partial\delta\tilde{\rho}}{\partial\vartheta}\right|_{\vartheta=0,\pi}=0. (16)

Thus, the solution of δρ~\delta\tilde{\rho} and δv~\delta\tilde{v} consists of Fourier series

δρ~n(ϑ,t)\displaystyle\delta\tilde{\rho}_{n}(\vartheta,t) =\displaystyle= A0+n=1Ancos(ωnt+φn)cos(nϑ),\displaystyle A_{0}\!+\!\!\sum_{n=1}^{\infty}\!A_{n}\cos(\omega_{n}t+\varphi_{n})\cos(n\vartheta), (17)
δv~(ϑ,t)=δv0ω0n=1Ansin(ωnt+φn)sin(nϑ),\displaystyle\delta\tilde{v}(\vartheta,t)=\delta v_{0}\!-\!\omega_{0}\!\!\!\sum_{n=1}^{\infty}\!A_{n}\sin(\omega_{n}t+\varphi_{n})\sin(n\vartheta), (18)

where ωn=nω0\omega_{n}=n\omega_{0} denotes the quantized excitation energy. This equidistant excitation energy is also consistent with the SU(1,1)\mathrm{SU}(1,1) dynamic symmetry. The constant A0A_{0} as a perturbation actually implies a small derivation from the chemical potential.The constant δv0\delta v_{0} is set to be 0 with a Galilean transformation.

In the above calculation, we assumed that the spatial derivative part of Gross-Pitaevskii (GP) equation can be neglected. The validity of this assumption has to be checked. By GP equation, we mean Φ(x,t)\Phi(x,t) satisfy a non-linear Schrödinger equation as follows,

itΦ=(x22μ+ω02x22+g02π22|Φ|4)Φ,\displaystyle i\partial_{t}\Phi=\!\left(\!-\frac{\partial_{x}^{2}}{2}\!-\!\mu\!+\!\frac{\omega_{0}^{2}x^{2}}{2}\!+\!\frac{g_{0}^{2}\pi^{2}}{2}|\Phi|^{4}\!\right)\Phi, (19)

where μ\mu denotes the chemical potential, Φ\Phi is short for Φ(x,t)\Phi(x,t). A direct numerical calculation for the excited states from the GP equation is difficult. Here we carry out the check in a reversed way. We set the initial wave function as a superposition of the ground state and the nn-th excited mode. Then, we check the time evolution of the density wave with this initial condition, where the information of the excited modes can be extracted. Approximately, we take the initial state as

Φn(x,0)=ρg(x)+0.1cos[narccos(x/R)].\displaystyle\Phi_{n}(x,0)=\sqrt{\rho_{g}(x)+0.1\cos[n\arccos(x/R)]}. (20)

The dynamics of ρn(x,t)|Φn(x,t)|2\rho_{n}(x,t)\equiv|\Phi_{n}(x,t)|^{2} can be calculated by the GP equation. We find the dynamics of ρn(x,t)\rho_{n}(x,t) can be factorized into several harmonic oscillations,

δρn(x,t)\displaystyle\delta\rho_{n}(x,t) \displaystyle\equiv ρn(x,t)ρn(x,t)¯\displaystyle\rho_{n}(x,t)-\overline{\rho_{n}(x,t)} (21)
=\displaystyle= mAm(x)cos[ωm(x)t+φm(x)],\displaystyle\sum_{m}A_{m}(x)\cos[\omega_{m}(x)t+\varphi_{m}(x)],

where ρn(x,t)¯=0Tmaxdtρn(x,t)/Tmax\overline{\rho_{n}(x,t)}=\int_{0}^{T_{\text{max}}}\mathrm{d}t\rho_{n}(x,t)/T_{\text{max}} is the average value of ρn(x,t)\rho_{n}(x,t). TmaxT_{\text{max}} is the time cutoff, which is taken as 50π50\pi here. These data are shown in Fig. 2. By a nonlinear fitting of the dynamics, we obtain the values of Am(x)A_{m}(x), ωm(x)\omega_{m}(x) and φm(x)\varphi_{m}(x). These parameters are xx dependent. If our result matches the theoretical predictions, ωm(x)=ωm\omega_{m}(x)=\omega_{m} is xx independent. Furthermore, we check that ωm\omega_{m} is interaction g0g_{0} independent. This property is exact in NN\rightarrow\infty limit, but it is approximate for finite NN. Also, we check ωm=mω1\omega_{m}=m\omega_{1}. If this equation fails to be true, there are soft modes in the system. The existence of these soft modes will lead to decoherence for the long-time evolution. As we find ω~m(x)=ωm(x)/m\tilde{\omega}_{m}(x)=\omega_{m}(x)/m are all very close to 11, therefore we will give the mean value of ω(x)\omega(x) as ω¯m=XmaxXmaxdxω~m(x)/(2Xmax)\overline{\omega}_{m}=\int_{-X_{\rm max}}^{X_{\rm max}}\mathrm{d}x\tilde{\omega}_{m}(x)/(2X_{\rm max}) and its variance Δωn2\Delta\omega_{n}^{2} as a function of g0g_{0} and NN in Tables 1, 2. In Fig. 2, one can observe that the dynamics in (b) is more stable than in (a), this is manifested in the frequency difference between ω¯2\bar{\omega}_{2} and ω¯4\bar{\omega}_{4}. The difference between ω¯2\bar{\omega}_{2} and ω¯4\bar{\omega}_{4} shrinks from 0.0030.003 to 0.00020.0002 when the particle increase from N=40N=40 to N=125N=125.

Table 1: In this table, we show the ω¯n=2,4,6\bar{\omega}_{n=2,4,6} for different excited modes. The particle number NN is fixed at 4040. The average is over different positions. Δωn2\Delta\omega_{n}^{2} are the variance of ω¯n(x)\bar{\omega}_{n}(x).
N=40N=40 ω¯2\ \ \ \bar{\omega}_{2} Δω22\ \ \Delta\omega_{2}^{2}    ω¯4\bar{\omega}_{4}\ \ \   Δω42\Delta\omega_{4}^{2}    ω¯6\bar{\omega}_{6}   Δω62\Delta\omega_{6}^{2}
g0=0.40g_{0}=0.40 1.010031.01003 0.00060.0006 0.997520.99752 0.00050.0005 1.001141.00114 0.00650.0065
g0=0.60g_{0}=0.60 1.019581.01958 0.02000.0200 1.003791.00379 0.00050.0005 0.988120.98812 0.01840.0184
g0=0.80g_{0}=0.80 1.011091.01109 0.00050.0005 1.005581.00558 0.00020.0002 0.984960.98496 0.02330.0233
g0=1.20g_{0}=1.20 1.013071.01307 0.00170.0017 1.006751.00675 0.00650.0065 0.998350.99835 0.00020.0002
g0=1.40g_{0}=1.40 1.014681.01468 0.00110.0011 1.010451.01045 0.00400.0040 1.000211.00021 0.00100.0010
Table 2: In this table, we show the ω¯n=2,4,6\bar{\omega}_{n=2,4,6} for different excited modes. The particle number NN is fixed at 125125. The average is over different positions. Δωn2\Delta\omega_{n}^{2} are the variance of ω¯n(x)\bar{\omega}_{n}(x).
N=125N=125 ω¯2\ \ \ \bar{\omega}_{2} Δω22\ \ \Delta\omega_{2}^{2}    ω¯4\bar{\omega}_{4}\ \ \   Δω42\Delta\omega_{4}^{2}    ω¯6\bar{\omega}_{6}   Δω62\Delta\omega_{6}^{2}
g0=0.40g_{0}=0.40 1.005681.00568 0.00460.0046 1.005041.00504 0.00010.0001 0.996430.99643 0.00250.0025
g0=0.60g_{0}=0.60 1.000311.00031 0.02280.0228 1.010041.01004 0.01900.0190 1.001221.00122 0.00150.0015
g0=0.80g_{0}=0.80 1.003811.00381 0.00940.0094 1.007431.00743 0.00100.0010 0.999530.99953 0.00660.0066
g0=1.20g_{0}=1.20 1.004421.00442 0.00320.0032 1.003711.00371 0.00100.0010 1.001611.00161 0.00100.0010
g0=1.40g_{0}=1.40 1.003831.00383 0.00250.0025 1.003711.00371 0.00120.0012 0.997020.99702 0.00200.0020

We also study the xx dependence in Am(x)A_{m}(x), which gives us the information about the eigenfunctions of the excited modes,

Am(x)δρ~m(x)1(x/R)2.\displaystyle A_{m}(x)\propto\frac{\delta\tilde{\rho}_{m}(x)}{\sqrt{1-(x/R)^{2}}}. (22)

We find these relations are satisfied both for N=40N=40 and N=125N=125. In Fig. 3(a) and (b), we show two typical δρ2(x,t)\delta\rho_{2}(x,t) at two different positions x=0x=0 and x=3.711x=3.711 (x/R=0.3x/R=0.3). In Fig. 3 (c), we show A2(x)A_{2}(x), A4(x)A_{4}(x) and A6(x)A_{6}(x) for g0=1.4g_{0}=1.4 and g0=0.6g_{0}=0.6 with the reference of theoretical prediction of δρ~2(x)/1(x/R)2\delta\tilde{\rho}_{2}(x)/\sqrt{1-(x/R)^{2}} and δρ~4(x)/1(x/R)2\delta\tilde{\rho}_{4}(x)/\sqrt{1-(x/R)^{2}} (Am(x)A_{m}(x) are even function of xx). One finds precise agreement between the numerical results and the theoretical predictions when |x|<R|x|<R. When |x|R|x|\approx R, the derivation indicates the finite-NN effects. The eigenfunctions show g0g_{0} independence up to a rescaling of RR (RR is in general g0g_{0} dependent).

Refer to caption
Figure 3: In (a) and (b), we show two typical time evolution of the excited modes for δρ2(x,t)\delta\rho_{2}(x,t) at different positions x=0x=0 and x=3.711x=3.711. N=125N=125 is fixed. By fitting the curves, we obtain A2(x)A_{2}(x), A4(x)A_{4}(x) and A6(x)A_{6}(x), which are even function of xx.By fixing the value of An=2,4,6(x=0)A_{n=2,4,6}(x=0) to be the same, we give the An=2,4,6(x/R)A_{n=2,4,6}(x/R) in (c).
Refer to caption
Figure 4: In (a), (b), and (c), we show the density-wave excitation δρ2CS(x)\delta\rho_{2}^{\rm CS}(x), δρ4CS(x)\delta\rho_{4}^{\rm CS}(x) and δρ6CS(x)\delta\rho_{6}^{\rm CS}(x) for fixed gcs=0.5g_{\rm cs}=0.5, respectively. The system size is taken as N=3N=3 ( the dot-dashed line ) and N=6N=6 ( the purple solid line ). The reference N=N=\infty line is given δρn(x)=cos[narccos(x/R)]/1(x/R)2\delta\rho_{n}^{\infty}(x)=\cos[n\arccos(x/R)]/\sqrt{1-(x/R)^{2}}. In (d), (e), and (f), the density wave excitation δρ2CS(x)\delta\rho_{2}^{\rm CS}(x), δρ4CS(x)\delta\rho_{4}^{\rm CS}(x) and δρ6CS(x)\delta\rho_{6}^{\rm CS}(x) are given for gcs=1.5g_{\rm cs}=1.5, respectively. The purple dashed lines in (d), (e), and (f) are the coarse graining of N=6N=6 curves.

Secondly, we study the density-wave excitation of the CS model. Because the CS model can be mapped identically to a set of free harmonic oscillators with EgE_{g} as the ground-state energy gurappa1999 , the excited energy is also equidistant. We consider the first excited state in the irreducible subspace involving the ground state, that is csm1969 ; gambar1975

Ψ1({xi})=(Egωcs)1/2(Egωcsωcsi=1Nxi2)Ψg.\displaystyle\Psi_{1}(\{x_{i}\})=\left(\frac{E_{g}}{\omega_{\rm cs}}\right)^{-1/2}\left(\frac{E_{g}}{\omega_{\rm cs}}-\omega_{\rm cs}\sum_{i=1}^{N}x_{i}^{2}\right)\Psi_{\rm g}. (23)

More generally, we have

Ψm({xi})=mB(Egωcs,m)LmEgωcs1(ωcsi=1Nxi2)×Ψg({xi}),\displaystyle\begin{split}\Psi_{m}(\{x_{i}\})=&\!\sqrt{mB\left(\frac{E_{g}}{\omega_{\rm cs}},m\right)}L_{m}^{\frac{E_{g}}{\omega_{\rm cs}}\!-\!1}\!\left(\omega_{\rm cs}\sum_{i=1}^{N}x_{i}^{2}\right)\\ &\times\Psi_{\rm g}(\{x_{i}\}),\end{split} (24)

where the index mm in Ψm\Psi_{m} refers to the mm-th excited state in the subspace. Here B(p,q)B(p,q) denotes the Beta-function, and Lmα(x)L_{m}^{\alpha}(x) denotes the generalized Laguerre polynomials. Notice that the jj-th excited state’s energy is Eg+2jωcsE_{g}+2j\omega_{\rm cs}.

Let Ψ(t=0)=c0Ψg+c1Ψ1\Psi(t=0)=c_{0}\Psi_{\rm g}+c_{1}\Psi_{1} with |c0|2+|c1|2=1|c_{0}|^{2}+|c_{1}|^{2}=1. We extract the information from the time evolution of the density ρ1CS(x,t)=j=2Ndxj|Ψ(t)|2=j=2Ndxj[|c0|2Ψg2+|c1|2Ψ12+(c0c1e2iωcst+c0c1e2iωcst)ΨgΨ1]\rho_{1}^{\rm CS}(x,t)=\int\prod_{j=2}^{N}\mathrm{d}x_{j}|\Psi(t)|^{2}=\int\prod_{j=2}^{N}\mathrm{d}x_{j}[|c_{0}|^{2}\Psi_{\rm g}^{2}+|c_{1}|^{2}\Psi_{1}^{2}+(c_{0}c_{1}^{*}e^{2i\omega_{\rm cs}t}+c_{0}^{*}c_{1}e^{-2i\omega_{\rm cs}t})\Psi_{\rm g}\Psi_{1}]. Since δρ\delta\rho is time dependent, we have δρ1CS(x,t)=(c0c1e2iωcst+c0c1e2iωcst)δρ1CS(x)\delta\rho_{1}^{\rm CS}(x,t)=(c_{0}c_{1}^{*}e^{2i\omega_{\rm cs}t}+c_{0}^{*}c_{1}e^{-2i\omega_{\rm cs}t})\delta\rho_{1}^{\rm CS}(x), where

δρ1CS(x)=j=2NdxjΨg(x,{xj})Ψ1(x,{xj})\displaystyle\delta\rho_{1}^{\rm CS}(x)=\int_{-\infty}^{\infty}\prod_{j=2}^{N}\mathrm{d}x_{j}\Psi_{\rm g}(x,\{x_{j}\})\Psi_{1}(x,\{x_{j}\}) (25)
=\displaystyle= j=2NdxjC(Egωcsωcsi=1Nxi2)C1Ψg(x,{xj})2\displaystyle\int_{-\infty}^{\infty}\prod_{j=2}^{N}\mathrm{d}x_{j}C\left(\frac{E_{g}}{\omega_{\rm cs}}-\omega_{\rm cs}\sum_{i=1}^{N}x_{i}^{2}\right)C^{-1}\Psi_{\rm g}(x,\{x_{j}\})^{2}
=\displaystyle= j=2NdxjC(Egωcs+ωcsωcs)[C1Ψg(x,{xj})2]\displaystyle\int_{-\infty}^{\infty}\prod_{j=2}^{N}\mathrm{d}x_{j}C\left(\frac{E_{g}}{\omega_{\rm cs}}+\omega_{\rm cs}\partial_{\omega_{\rm cs}}\right)\left[C^{-1}\Psi_{\rm g}(x,\{x_{j}\})^{2}\right]
=\displaystyle= C(Egωcsωcsωcs)[C1ρgCS(x)]\displaystyle C\left(\frac{E_{g}}{\omega_{\rm cs}}-\omega_{\rm cs}\frac{\partial}{\partial\omega_{\rm cs}}\right)\left[C^{-1}\rho_{\rm g}^{\rm CS}(x)\right]
\displaystyle\propto cos[2arccos(x/R)]1(x/R)2,\displaystyle\frac{\cos[2\arccos(x/R)]}{\sqrt{1-(x/R)^{2}}},

which is consistent with Eq. (17) for n=2n=2. In the derivation, we have assumed that the partial derivative ωcs\partial_{\omega_{\rm cs}} and the thermodynamic limit are interchangeable. However, for higher excited states which involves higher partial derivatives, this assumption is invalid due to the divergent derivative of ρgCS(x)\rho_{\rm g}^{\rm CS}(x) at the boundary x=±Rx=\pm R. Although we can not prove the general form of δρm(x)\delta\rho_{m}(x), we can guess its form is consistent with Eq. (17), i. e.,

δρmCS(x)\displaystyle\delta\rho_{m}^{\rm CS}(x) =\displaystyle= j=2NdxjΨg({xj})Ψm({xj})\displaystyle\int_{-\infty}^{\infty}\prod_{j=2}^{N}\mathrm{d}x_{j}\Psi_{\rm g}(\{x_{j}\})\Psi_{m}(\{x_{j}\}) (26)
\displaystyle\propto cos[2marccos(x/R)]1(x/R)2\displaystyle\frac{\cos[2m\arccos(x/R)]}{\sqrt{1-(x/R)^{2}}}

For an arbitrary gcsg_{\rm cs}, it is too difficult to solve the overlap integral for an any excited state, but Eq. (26) can be proved when gcs=1g_{\rm cs}=1. Here, one of the excited states can be represented by the single-particle energy eigenstates of the harmonic oscillator {fn(x)}n=0,1,2,.\{f_{n}(x)\}_{n=0,1,2,\cdots.}. Let |0,1,,N1|0,1,\cdots,N-1\rangle denote the occupied single-particle energy eigenstates in the ground state of NN fermions. Then, one of the mm-th excited states |ψm=A^b^Nl+mb^Nl|0,1,,N1|\psi_{m}\rangle=\hat{A}\hat{b}^{\dagger}_{N-l+m}\hat{b}_{N-l}|0,1,\cdots,N-1\rangle, for some 1lm1\leq l\leq m, where A^\hat{A} denotes the unit antisymmetric function girardeau1960 , b^l\hat{b}^{\dagger}_{l} (b^l\hat{b}_{l}) denotes the creation (annihilation) operator of ll single-particle energy eigenstates. The corresponding perturbation of the density is given by (notice A^2=1\hat{A}^{2}=1)

δρmCS(x)\displaystyle\delta\rho_{m}^{\rm CS}(x) =\displaystyle= j=2Ndxiψ0(x,{xj})ψm(x,{xj})\displaystyle\int_{-\infty}^{\infty}\prod_{j=2}^{N}\mathrm{d}x_{i}\psi_{0}(x,\{x_{j}\})\psi_{m}(x,\{x_{j}\}) (27)
=\displaystyle= fNl(x)fNl+m(x).\displaystyle f_{N-l}(x)f_{N-l+m}(x).

Using the asymptotic expansion of the product in the thermodynamic limit (appendix A):

fNl(2Nωcsx)fNl+m(2Nωcsx)ωcs2π2Ncos(marccosx)1x2.\displaystyle\begin{split}&f_{N-l}\left(\sqrt{\frac{2N}{\omega_{\rm cs}}}x\right)f_{N-l+m}\left(\sqrt{\frac{2N}{\omega_{\rm cs}}}x\right)\\ &\sim\sqrt{\frac{\omega_{\rm cs}}{2\pi^{2}N}}\frac{\cos(m\arccos x)}{\sqrt{1-x^{2}}}.\end{split} (28)

Comparing this result with Eq. (17), we prove that the density-wave excitations in the trapped |Φ|6|\Phi|^{6} theory and the CS model are the same when gcs=1g_{\rm cs}=1.

Now, we are going to check Eq. (26) for finite-NN systems by numerics. The overlap integral is numerically done by using Monte-Carlo method because the integrant is semi-positive definite. Here we introduce a normalization condition dxδρm(x)2ρgCS(x)=π/2\int\mathrm{d}x\delta\rho_{m}(x)^{2}\rho_{\rm g}^{\rm CS}(x)=\pi/2 for the comparison between different results, where δρm(x)\delta\rho_{m}(x) and ρgCS(x)\rho_{\rm g}^{\rm CS}(x) are numerically obtained with a finite NN. In Fig. 4, we present δρm=1,2,3(x)\delta\rho_{m=1,2,3}(x) for gcs=0.5g_{\rm cs}=0.5. There are datas for two sizes: N=3N=3 and N=6N=6. One can see the tendency for δρm(x)\delta\rho_{m}(x) approaching the guessed function in the thermodynamic limit. The results for gcs=1.5g_{\rm cs}=1.5 are given in (d), (e) and (f). There are clear deviations from the guessed results due to the high-frequency oscillation. The coarse-graining line is taken by replacing δρm(x)\delta\rho_{m}(x) as xΔxx+Δxdxδρm(x)/(2Δx)\int_{x-\Delta x}^{x+\Delta x}\mathrm{d}x^{\prime}\delta\rho_{m}(x^{\prime})/(2\Delta x). Here Δx\Delta x is taken as 0.250.25. The purple dashed lines are results after two times of coarse graining, which can be viewed as the average behavior of δρm(x)\delta\rho_{m}(x). One can see their resemblance with the guessed result. From the numerical result we find when gcs1g_{\rm cs}\gg 1, then the deviation from the theoretical prediction is more serious. In the CS model, when gcsg_{\rm cs} is large, the particles are inclined to be aligned in equal space in one-dimension due to the long-range interaction. On the other hand, when gcs<1g_{\rm cs}<1, ϑ=arccos(x/R)\vartheta=\arccos(x/R) characterizes the oscillation of the density wave. But the oscillation in ϑ\vartheta coordinate is not consistent with the equal spacing in xx coordinate. Thus these two requirements are conflicting with each other.For this reason, when gcs/N0g_{\rm cs}/N\rightarrow 0 is a suitable requirement for our guess.

II.4 Fractional exclusion statistics in |Φ|6|\Phi|^{6} theory

The wave function Eq. (15) implies the canonical quantization of the density-wave excitation of the trapped |Φ|6|\Phi|^{6} theory. Following the procedure of the phenomenological bosonization in Ref. giamarchi2003 , we introduce a scalar field ϕ\phi and its conjugate momentum Π\Pi through δρ~=ϕ/ϑ,δv~=ω(gπ)1Π\delta\tilde{\rho}=-\partial\phi/\partial\vartheta,\delta\tilde{v}=\omega(g\pi)^{-1}\Pi. Then, using the solution (Eqs. 17, 18) and to the second order of δρ\delta\rho, the energy functional (Eq. 2) is rewritten in terms of ϕ\phi and Π\Pi as

H=12N2g0ω0+120πdϑ[ω0g0πΠ2+πg0ω0(ϕϑ)2],\displaystyle H=\frac{1}{2}N^{2}g_{0}\omega_{0}+\frac{1}{2}\int_{0}^{\pi}\mathrm{d}\vartheta\left[\frac{\omega_{0}}{g_{0}\pi}\Pi^{2}+\pi g_{0}\omega_{0}\left(\frac{\partial\phi}{\partial\vartheta}\right)^{2}\right], (29)

which describes a Luttinger-liquid system. In the derivation, we neglect the term dx(ρ0/x)2\int\mathrm{d}x(\partial\rho_{0}/\partial x)^{2} within the TF approximation.

We quantize the scalar field with the commutation relation [ϕ^(t,y),Π^(t,y)]=iδ(yy)[\hat{\phi}(t,y),\hat{\Pi}(t,y^{\prime})]=i\delta(y-y^{\prime}). Thus, the density wave (Eq. 17, 18) is of bosonic-excitation type. It follows from Eqs. (17), (18) that

H^=12g0N2ω0+n=1nω0a^na^n,\displaystyle\hat{H}=\frac{1}{2}g_{0}N^{2}\omega_{0}+\sum_{n=1}^{\infty}n\omega_{0}\hat{a}_{n}^{\dagger}\hat{a}_{n}, (30)

where ϕ^/ϑ=2/πn(2g0πn)1/2(a^n+a^n)cos(nϑ)\partial\hat{\phi}/\partial\vartheta=\sqrt{2/\pi}\sum_{n}(2g_{0}\pi n)^{-1/2}(\hat{a}_{n}^{\dagger}+\hat{a}_{n})\cos(n\vartheta), Π^=ni(g0πn/2)1/2(a^na^n)2/πsin(nϑ)\hat{\Pi}=\sum_{n}i(g_{0}\pi n/2)^{1/2}(\hat{a}_{n}^{\dagger}-\hat{a}_{n})\sqrt{2/\pi}\sin(n\vartheta) with {a^n,a^n}\{\hat{a}_{n}^{\dagger},\hat{a}_{n}\} are the creation and annihilation operators of the density-wave modes. Compared with the energy spectrum of the CS model gurappa1999 , Eq. (30) is also a set of free harmonic oscillators with 12g0N2ω0\frac{1}{2}g_{0}N^{2}\omega_{0} as the ground-state energy. Because, the only difference between the two models is just a negligible chemical potential deviation (1g0)ω0/2(1-g_{0})\omega_{0}/2 for ω00\omega_{0}\to 0 in the thermodynamic limit, we conclude that the trapped |Φ|6|\Phi|^{6} theory has the CS model-like energy spectra and thus obeys the fractional exclusion statistics with g0g_{0} as the statistical parameter. About the connection between the Luttinger-liquid and the fractional exclusion statistics, we introduce Refs. Pham1998 ; Wu2001 for further reading .

Refer to caption
Refer to caption
Figure 5: The density of the |Φ|6|\Phi|^{6} theory at finite temperature. For x<0x<0, the density is an even function. (a) The black solid line denotes the numerical result from Eq. (33). The red and blue solid lines denote the approximate results from Eq. (35) and Eq. (36) respectively. (b) The black solid line denotes the numerical result from Eq. (33). The blue dashed line denotes the density of a ideal fermionic system with the same temperature.

II.5 Finite Temperature Results

The linearized density-wave excitations capture the low-energy excitations of the |Φ|6|\Phi|^{6} theory. When the temperature kBTk_{\text{B}}T (kBk_{\text{B}} denotes the Boltzmann constant) of the field grows up to NωN\hbar\omega, the thermal-excitation energy in Eq. (30) is comparable to the ground-state energy. At this moment, the linearized theory fails and we use the semiclassical approximation to study the finite-temperature property of the |Φ|6|\Phi|^{6} theory in the following.

Using the Hatree-Fock approximation and the WKB approximation huse1982 ; giorgini1997 ; bhaduri2000 , the density ρ(x)\rho(x) is determined self-consistently. First there should no condensate part in ρ(x)\rho(x) at a high temperature. The system consists of thermal particles and the density ρ(x)\rho(x) at finite temperature β=1/kBT\beta=1/k_{\text{B}}T reads

ρ(x)=dp/(2π)eβε(p,x)1,\displaystyle\rho(x)=\int_{-\infty}^{\infty}\frac{\mathrm{d}p/(2\pi\hbar)}{e^{\beta\varepsilon(p,x)}-1}, (31)

where

ε(p,x)=p22+12ω02x2+g02π222ρ(x)2μ\displaystyle\varepsilon(p,x)=\frac{p^{2}}{2}+\frac{1}{2}\omega_{0}^{2}x^{2}+\frac{g_{0}^{2}\pi^{2}\hbar^{2}}{2}\rho(x)^{2}-\mu (32)

denotes the effective single-particle spectrum, and μ\mu denotes the chemical potential. After performing the pp-integration of Eq. (31), we find the self-consistent equation of ρ(x)\rho(x):

u(x)=Li1/2[eβμβω02x2/2πg02u(x)2/4],\displaystyle u(x)=\mathrm{Li}_{1/2}[e^{\beta\mu-\beta\omega_{0}^{2}x^{2}/2-\pi g_{0}^{2}u(x)^{2}/4}], (33)

where u(x)=2πβρ(x)u(x)=\sqrt{2\pi\beta}\rho(x) denotes the dimensionless density, and Lis(z)=n=1zn/ns\mathrm{Li}_{s}(z)=\sum_{n=1}^{\infty}z^{n}/n^{s} denotes the polylogarithm function. Also, the constraint is given by

N2πβ=dxu(x).\displaystyle N\sqrt{2\pi\beta}=\int_{-\infty}^{\infty}\mathrm{d}xu(x). (34)

Specially, we find the solution of ρ(x)\rho(x) can always be obtained self-consistently for any temperature, which indicates that there are no condensate in ρ(x)\rho(x) at any temperature.

Refer to caption
Figure 6: The μT\mu-T curve of the |Φ|6|\Phi|^{6} theory at finite temperature. The black solid line denotes the numerical result from Eq. (33). The blue dashed denotes the μT\mu-T curve of the CS model.

In Fig. 5, we show the density ρ(x)\rho(x) numerically solved from Eq. (33). Fig. 5a indicates three typical regions: (1) semicircle distribution: |x|<R|x|<R; (2) Gaussian tail: |x|>R|x|>R; (3) near the boundary |x|R|x|\approx R. Actually, the region 1 and 3 respectively corresponds to the two limits of the polylogarithm function: Li1/2(z)π/(1z)\mathrm{Li}_{1/2}(z)\to\sqrt{\pi/(1-z)} for z1z\to 1^{-}, and Li1/2(z)z\mathrm{Li}_{1/2}(z)\to z for z0z\to 0. Accordingly, the density in the two regions are

u1(x)=(2μω02x2)πg02kBT1+1+(2πg0kBT2μω02x2)2,\displaystyle u^{1}(x)=\sqrt{\frac{(2\mu-\omega_{0}^{2}x^{2})}{\pi g_{0}^{2}k_{\text{B}}T}}\sqrt{1+\sqrt{1+\left(\frac{2\pi g_{0}k_{\text{B}}T}{2\mu-\omega_{0}^{2}x^{2}}\right)^{2}}}, (35)
u3(x)=eβ(μω02x2/2).\displaystyle u^{3}(x)=e^{\beta(\mu-\omega_{0}^{2}x^{2}/2)}. (36)

Substituting Eq. (35), or Eq. (36) into Eq. (32), we have

ε1(p,x)=p22+(πg0kBT)24μ2ω02x2\displaystyle\varepsilon^{1}(p,x)=\frac{p^{2}}{2}+\frac{(\pi g_{0}k_{\text{B}}T)^{2}}{4\mu-2\omega_{0}^{2}x^{2}} (37)

which results in the semicircle distribution of the density and implies the existence of the Fermi surface, or

ε3(p,x)=p22+12ω02x2μ\displaystyle\varepsilon^{3}(p,x)=\frac{p^{2}}{2}+\frac{1}{2}\omega_{0}^{2}x^{2}-\mu (38)

which describes the ideal Bosons, but now the chemical potential contains the effects of interaction. In addition, near the boundary |x|R|x|\approx R with a length scale RkBT/μRk_{\text{B}}T/\mu, the analytical solutions (Eqs. 35, 36) are invalid. Fig. 5b compares the density of the |Φ|6|\Phi|^{6} theory (g0=1g_{0}=1) with it of a free-fermion system for fixed NN and TT. The narrower distribution of the density from the |Φ|6|\Phi|^{6} theory indicates that more particles occupy the low-energy states, which lead to a lower chemical potential.

In Fig. 6, we show the μT\mu-T figure of the |Φ|6|\Phi|^{6} theory (Eqs. 33, 34) and the CS model wu1994

μCS=g0Nω0+kBTln[1eNω0/kBT].\displaystyle\mu_{\text{CS}}=g_{0}N\omega_{0}+k_{B}T\ln[1-e^{-N\omega_{0}/k_{B}T}]. (39)

Compared with the CS model’s result, the chemical potential of the |Φ|6|\Phi|^{6} theory drops rapidly as the temperature grows, which is consistent with Fig. 5b. The deviation of the two curves indicates that the |Φ|6|\Phi|^{6} theory does not satisfy the fractional exclusion statistics at the temperature scale kBTNω0k_{\text{B}}T\sim N\omega_{0}, in contrast to the result that the |Φ|6|\Phi|^{6} theory satisfies the fractional exclusion statistics within the linearized theory at the temperature scale kBTNω0k_{\text{B}}T\ll N\omega_{0}. This deviation is a result of the expansion of the semicircle distribution at finite temperature, which decreases the density and weakens the three-body interaction of the field (see also Fig. 5b).

III Conclusion

In this paper, we find a hydrodynamical description of ideal anyons satisfying fractional exclusion statistics in the one-dimensional |Φ|6|\Phi|^{6} theory. The |Φ|6|\Phi|^{6} theory shares many similar features as the Calogero-Sutherland model. Specifically, when the trap frequency and the interaction strength in |Φ|6|\Phi|^{6} theory and them in the CS model are consistent the two models share the same ground state density profile, excitation spectrum, density-wave excitation , as well as dynamical and statistical properties in the thermodynamic limit and at low temperature. The hydrodynamical version of ideal anyons satisfying fractional exclusion statistics is more accessible in experiments because it only relates local contact interactions, while r2r^{-2} interaction in the CS model is, on the other hand, very difficult to realize experimentally.

IV Acknowledgement

Z. Y. Fei is supported by the NSFC (Grants No. 12088101), NSAF (Grants No. U1930403, No. U1930402), and the China Postdoctoral Science Foundation (No. 2021M700359). Y. Chen is supported by the National Key Research and Development Program of China (Grant No. 2022YFA1405300), and NSFC under Grant No. 12174358 and No. 11734010.

Appendix A: asymptotic expansion of fnf_{n}

The eigenstates of the harmonic oscillator is given by

fn(x)=1n!2n(ωcsπ)1/4Hn(ωcsx)eωcs2x2.\displaystyle f_{n}(x)=\frac{1}{\sqrt{n!2^{n}}}\left(\frac{\omega_{\rm cs}}{\pi}\right)^{1/4}H_{n}(\sqrt{\omega_{\rm cs}}x)e^{-\frac{\omega_{\rm cs}}{2}x^{2}}. (A.1)

In the thermodynamic limit, according to Ref. Dominici2005 , we have (|x|<1|x|<1)

fN(2Nωcsx)[2ωcsπ2N(1x2)]1/4cosΘN(x),\displaystyle\begin{split}f_{N}\left(\sqrt{\frac{2N}{\omega_{\rm cs}}}x\right)\sim\left[\frac{2\omega_{\rm cs}}{\pi^{2}N(1-x^{2})}\right]^{1/4}\cos\Theta_{N}(x),\end{split} (A.2)

where ΘN(x)=(N+1/2)arccosxNx1x2π/4\Theta_{N}(x)=(N+1/2)\arccos x-Nx\sqrt{1-x^{2}}-\pi/4. Then, for the product fNlfNl+mf_{N-l}f_{N-l+m}, in the limit m/N0m/N\to 0, we have

2π2N(1x2)ωcsfNl(2Nωcsx)fNl+m(2Nωcsx)12cosΘNl(NNlx)cosΘNl+m(NNl+mx)=cos[ΘNl(NNlx)ΘNl+m(NNl+mx)]+cos[ΘNl(NNlx)+ΘNl+m(NNl+mx)]cos(marccosx)+cos[ΘNl(NNlx)+ΘNl+m(NNl+mx)].\displaystyle\begin{split}&\sqrt{\frac{2\pi^{2}N(1-x^{2})}{\omega_{\rm cs}}}f_{N-l}\left(\sqrt{\frac{2N}{\omega_{\rm cs}}}x\right)f_{N-l+m}\left(\sqrt{\frac{2N}{\omega_{\rm cs}}}x\right)\\ &\sim\frac{1}{2}\cos\Theta_{N-l}\left(\sqrt{\frac{N}{N-l}}x\right)\cos\Theta_{N-l+m}\left(\sqrt{\frac{N}{N-l+m}}x\right)\\ &=\cos\left[\Theta_{N-l}\left(\sqrt{\frac{N}{N-l}}x\right)-\Theta_{N-l+m}\left(\sqrt{\frac{N}{N-l+m}}x\right)\right]+\cos\left[\Theta_{N-l}\left(\sqrt{\frac{N}{N-l}}x\right)+\Theta_{N-l+m}\left(\sqrt{\frac{N}{N-l+m}}x\right)\right]\\ &\sim\cos(m\arccos x)+\cos\left[\Theta_{N-l}\left(\sqrt{\frac{N}{N-l}}x\right)+\Theta_{N-l+m}\left(\sqrt{\frac{N}{N-l+m}}x\right)\right].\end{split} (A.3)

Ignoring the high frequency oscillating term (the second term), we obtain Eq. (LABEL:ea).

References

  • (1) L. Landau and E. Lifshitz, Fluid Mechanics: Volume 6 (Course of Theoretical Physics S) (Butterworth-Heinemann, London, 1987).
  • (2) Z.-Y. Shi, C. Gao, and H. Zhai, Phys. Rev. X 11, 041031 (2021).
  • (3) R. Saint-Jalm, P. C. M Castilho, É. Le Cerf, B. Bakkali- Hassani, J.-L. Ville, S. Nascimbene, J. Beugnon, and J. Dalibard, Phys. Rev. X 9, 021035 (2019).
  • (4) C. Lv, R. Zhang, and Q. Zhou, Phys. Rev. Lett. 125, 253002 (2020).
  • (5) F. D. M. Haldane, Phys. Rev. Lett. 67, 937 (1991).
  • (6) Y.-S. Wu, Phys. Rev. Lett. 73, 922 (1994).
  • (7) F. Calogero, J. Math. Phys. 10, 2191 (1969); J. Math. Phys. 10, 2197 (1969); J. Math. Phys. 12, 419 (1971)
  • (8) B. Sutherland, J. Math. Phys. 12, 246 (1971).
  • (9) M. V. N. Murthy, and R. Shankar, Phys. Rev. Lett. 73(25), 3331 (1994).
  • (10) M. Girardeau, J. Math. Phys. 1, 516 (1960); M. D. Girardeau, and E. M. Wright, Phys. Rev. Lett. 84, 5691, (2000).
  • (11) E. P. Wigner, Ann. Math. 62, 548 (1955).
  • (12) F. Dalfovo, S. Giorgini, L. P. Pitaevskii, and S. Stringari, Rev. Mod. Phys. 71, 463 (1999).
  • (13) I. Andrić, and A. Jevicki, and H. Levine, Nuclear Physics B 215 [FS7] 307-315 (1983); I AndriĆ and V Bardek, J. Phys. A: Math. Gen. 21 2847-2853 (1988).
  • (14) D. Sen and R. K. Bhaduri, Annals of Physics 260, 203-223 (1997).
  • (15) L. P. Pitaevskii and A. Rosch, Phys. Rev. A 55, R853 (1997).
  • (16) E. Wigner, Phys. Rev. 46, 1002 (1934).
  • (17) E. B. Kolomeisky, T. J. Newman, J. P. Straley, and X. Y. Qi, Phys. Rev. Lett. 85, 1146 (2000).
  • (18) P. J. Forrester, Log-gases and random matrices. (Princeton university press, 2010).
  • (19) M. Ledoux, The Concentration of Measure Phenomenon (American Mathematical Soc., August 2001).
  • (20) S. Stringari, Phys. Rev. Lett. 77, 2360 (1996).
  • (21) C. Menotti, and S. Stringari, Phys. Rev. A 66, 043610 (2002).
  • (22) A. Imambekov, and E. Demler, Annals of Physics, 321(10), 2390-2437 (2006).
  • (23) N. Gurappa and P. K. Panigrahi, Phys. Rev. B 59, R2490(R) (1999).
  • (24) P. J. Gambardella, J. Math. Phys, 16 (5), 1172 (1975).
  • (25) T. Giamarchi, Quantum physics in one dimension (Vol. 121). (Clarendon press, 2003).
  • (26) K.-V. Pham, M. Gabay, P. Lederer, arXiv:cond-mat/9808221 (1998).
  • (27) Y.-S. Wu, Y. Yu, H.-X. Yang, Nuclear Physics B 604[FS], 551–579 (2001).
  • (28) D. A. Huse, and E. D. Siggia, J. Low Temp. Phys. 46, 137 (1982).
  • (29) S. Giorgini, L. P. Pitaevskii, and S. Stringari, J. Low Temp. Phys. 109(1), 309-355 (1997).
  • (30) R. K. Bhaduri, S. M. Reimann, S. Viefers, A. G. Choudhury, and M. K. Srivastava. Phys. B: At. Mol. Opt. Phys. 33 (2000) 3895–3903.
  • (31) D. Dominici, arXiv:math.CA/0508264 (2005); Journal of Difference Equations and Applications 13(12), 1115–1128 (2007).