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

Energetics and Efimov states of three interacting bosons and mass-imbalanced fermions in a three-dimensional spherical harmonic trap

A. D. Kerin    A. M. Martin School of Physics, University of Melbourne, Parkville, VIC 3010, Australia
Abstract

We consider a system of three particles, either three identical bosons or two identical fermions plus an impurity, within a three-dimensional isotropic trap interacting via a contact interaction. Using two approaches, one using an infinite sum of basis states for the wavefunction and the other a closed form wavefunction, we calculate the allowable energy eigenstates of the system as a function of the interaction strength, including the strongly and weakly interacting limits. For the fermionic case this is done while maintaining generality regarding particle masses. We find that the two methods of calculating the spectrum are in excellent agreement in the strongly interacting limit. However the infinite sum approach is unable to uniquely specify the energy of Efimov states, but in the strongly interacting limit there is, to a high degree of accuracy, a correspondence between the three-body parameter required by the boundary condition of the closed form approach and the summation truncation order required by the summation approach. This specification of the energies and wavefunctions forms the basis with which thermodynamic variables such as the virial coefficients or Tan contacts, or dynamic phenomena like quench dynamics can be calculated.

I Introduction

Ultracold quantum gases provide an ideal testbed for the investigation of many-body quantum systems. The foundation of such studies is the investigation of the properties of few-body systems Busch et al. (1998); Fedorov and Jensen (2001); Werner and Castin (2006a); Kestner and Duan (2007); Daily and Blume (2010); Mulkerin et al. (2012a); Werner and Castin (2006b); Blume (2012). Through solutions of the few-body problem it is possible to calculate the thermodynamics of many-body quantum gases Liu et al. (2009, 2010); Cui (2012); Mulkerin et al. (2012a); Stöferle et al. (2006); Rakshit et al. (2012); Kaplan and Sun (2011); Mulkerin et al. (2012b); Nascimbène et al. (2010); Ku et al. (2012); Levinsen et al. (2017) and the quench dynamics of few-body systems Bougas et al. (2020, 2019); Budewig et al. (2019); Kehrberger et al. (2018). Significant research has been undertaken Cui (2012); D’Incao et al. (2018); Jonsell et al. (2002); Blume and Daily (2010) into the study of the interacting three-body problem for both Bose and Fermi gases trapped in three-dimensional spherically symmetric harmonic traps. Such work has lead to our understanding of Efimov states Efimov (1973, 1970); Jonsell et al. (2002); Kartavtsev and Malykh (2007, 2008); Endo et al. (2011); Wang et al. (2012); Efimov (1971) in Bose and mass-imbalanced Fermi systems.

More specifically, recent theoretical results in quench dynamics allow for the semi-analytic calculation of quench observables in the three-body system for a quench in s-wave scattering length Kerin and Martin (2022a, b). However only the non-interacting to unitary and vice-versa quenches can be considered, unlike the two-body quench where any quench in s-wave scattering length can be considered Kerin and Martin (2020). Generalising the three-body quench quench calculation to arbitrary scattering lengths is then of great interest and the first step towards that is calculating the energy spectrum. This has already been done for equal mass fermions Liu et al. (2010) and here we extend those calculations to the bosonic and mass-imbalanced fermionic cases. In addition to the relevance to quench dynamics this generalisation also allows for the calculation of thermodynamic quantities such as virial coefficients or the Tan contacts Tan (2008a, b, c).

In this work we consider two three-body systems: (i) a 2+1 fermion system, with mass imbalance between two substrate particles and an impurity particle and (ii) three identical bosons. In each case the particles interact via a contact interaction and are confined to a spherically symmetric harmonic trap. We derive for general interaction strength and, for fermions, arbitrary mass imbalance between the substrate and impurity, the eigenspectrum for the system via a matrix approach Liu et al. (2009). For this system we also derive the system eigenstates in the unitary regime, where the s-wave contact interaction dominates. For comparison we revisit the solutions for the three boson system. For the 2+1 Fermi mass imbalanced system, as expected, we find Efimov-like states using the matrix approach. This then leads us to revisit the three-body parameter for investigating Efimov states Jonsell et al. (2002). We find a connection between the numeric matrix approach and the three-body parameter approaches in determining that for a given matrix size there is a value of the three-body parameter which produces, to a high degree of accuracy, the same energy spectrum at unitarity.

II Overview of the Three-Body Problem

We consider a system of three interacting particles with arbitrary masses in a three dimensional spherically symmetric harmonic trap. The positions of the three particles are given r1\vec{r}_{1}, r2\vec{r}_{2}, and r3\vec{r}_{3} with respective masses m1m_{1}, m2m_{2}, and m3m_{3}. The Hamiltonian is given by

H^3b\displaystyle\hat{H}_{\rm 3b} =\displaystyle= k=1322mkk2+mkω2rk22,\displaystyle\sum_{k=1}^{3}-\frac{\hbar^{2}}{2m_{k}}\nabla_{k}^{2}+\frac{m_{k}\omega^{2}r_{k}^{2}}{2}, (1)

and the interparticle interactions are modelled as contact interactions enforced by the Bethe-Peierls boundary condition Bethe and Peierls (1935). For convenience we define the following coordinate transformations,

C\displaystyle\vec{C} =\displaystyle= m1r1+m2r2+m3r3m1+m2+m3,\displaystyle\frac{m_{1}\vec{r}_{1}+m_{2}\vec{r}_{2}+m_{3}\vec{r}_{3}}{m_{1}+m_{2}+m_{3}}, (2)
r\displaystyle\vec{r} =\displaystyle= r2r1,\displaystyle\vec{r}_{2}-\vec{r}_{1}, (3)
ρ\displaystyle\vec{\rho} =\displaystyle= 1γ(r3m1r1+m2r2m1+m2),\displaystyle\frac{1}{\gamma}\left(\vec{r}_{3}-\frac{m_{1}\vec{r}_{1}+m_{2}\vec{r}_{2}}{m_{1}+m_{2}}\right), (4)
γ\displaystyle\gamma =\displaystyle= m1(m1+m2+m3)(m1+m2)(m1+m3).\displaystyle\sqrt{\frac{m_{1}(m_{1}+m_{2}+m_{3})}{(m_{1}+m_{2})(m_{1}+m_{3})}}.

We can then rewrite the Hamiltonian as

H^3b=H^COM+H^rel,\displaystyle\hat{H}_{\rm 3b}=\hat{H}_{\rm COM}+\hat{H}_{\rm rel}, (5)
H^COM=22MC2+Mω2C22,\displaystyle\hat{H}_{\rm COM}=-\frac{\hbar^{2}}{2M}\nabla^{2}_{C}+\frac{M\omega^{2}C^{2}}{2}, (6)
H^rel=22μ12r2+μ12ω2r2222μ13ρ2+μ13ω2ρ22.\displaystyle\hat{H}_{{\rm rel}}=-\frac{\hbar^{2}}{2\mu_{12}}\nabla^{2}_{r}+\frac{\mu_{12}\omega^{2}r^{2}}{2}-\frac{\hbar^{2}}{2\mu_{13}}\nabla^{2}_{\rho}+\frac{\mu_{13}\omega^{2}\rho^{2}}{2}.\quad (7)

where μjk=mjmk/(mj+mk)\mu_{jk}=m_{j}m_{k}/(m_{j}+m_{k}) and M=m1+m2+m3M=m_{1}+m_{2}+m_{3}. H^COM\hat{H}_{\rm COM} is the centre-of-mass (COM) part of the Hamiltonian, the solution to which is the simple harmonic oscillator (SHO) wavefunction, with mass MM. H^rel\hat{H}_{\rm rel} is the relative part of the Hamiltonian.

Throughout the rest of this paper we consider two specific cases. The first is two identical substrate fermions interacting with a third impurity particle. For this case we set m1=mim_{1}=m_{\rm i} and m2=m3=mm_{2}=m_{3}=m, such that μ13=μ12=μ=mim/(mi+m)\mu_{13}=\mu_{12}=\mu=m_{\rm i}m/(m_{\rm i}+m). The second case we consider is three identical interacting bosons. In this case m1=m2=m3=mm_{1}=m_{2}=m_{3}=m and μ13=μ12=m/2\mu_{13}=\mu_{12}=m/2. For convenience we can in general define μ=mim/(mi+m)\mu=m_{\rm i}m/(m_{\rm i}+m) where mi=mm_{\rm i}=m for the bosonic case. Finally we also define the COM harmonic length scale to be aCOM=/(mi+2m)ωa_{\rm COM}=\sqrt{\hbar/(m_{\rm i}+2m)\omega}, and the relative harmonic length scale to be aμ=/μωa_{\mu}=\sqrt{\hbar/\mu\omega}, where for bosons mi=mm_{\rm i}=m.

For the fermionic case we investigate the effects of mass imbalance and so we define κ=m/mi\kappa=m/m_{\rm i}. The effects of changing κ\kappa are investigated in detail in Sections II.1 and II.2. It should be noted that as the value of κ\kappa changes so too does aμa_{\mu}. Due to the dependence of aμa_{\mu} on κ\kappa we use am/asa_{m}/a_{\rm s} for κ1\kappa\leq 1 and ami/asa_{m_{\rm i}}/a_{\rm s} for κ1\kappa\geq 1 in Figs. 1, 2, and 3 where am=/mωa_{m}=\sqrt{\hbar/m\omega} and ami=/miωa_{m_{\rm i}}=\sqrt{\hbar/m_{\rm i}\omega}.

In a non-interacting system Eqs. (6) and (7) can be solved with SHO wavefunctions. However, we impose the Bethe-Peierls boundary condition Bethe and Peierls (1935) to enforce a contact interaction,

limrij0[d(rijΨ3b)drij1rijΨ3b]=1as.\displaystyle\lim_{r_{ij}\rightarrow 0}\left[\frac{d(r_{ij}\Psi_{\rm 3b})}{dr_{ij}}\frac{1}{r_{ij}\Psi_{3\rm b}}\right]=\frac{-1}{a_{\rm s}}. (8)

Here Ψ3b\Psi_{\rm 3b} is the total three-body wavefunction, rij=|rirj|r_{ij}=|\vec{r}_{i}-\vec{r}_{j}|, and asa_{\rm s} is the s-wave scattering length. The COM part of the wavefunction is independent of the boundary condition, and only the relative wavefunction is affected. Enforcing this boundary condition is equivalent to including a Fermi pseudo-potential term in the Hamlitonian Braaten and Hammer (2006).

If the wavefunction is symmetric under the interchange of particles jj and kk and satisfies the Bethe-Peierls boundary condition for rijr_{ij} then it satisfies the Bethe-Peierls boundary condition for rikr_{ik}. This relationship between particle interaction and particle symmetry allows us to consider the 2+1 fermion and three boson cases while specifying only one scattering length. Other cases require additional boundary conditions and are beyond the scope of this work. For example in the 2+1 boson case the scattering length between the two identical bosons is not necessarily the same as between one of those bosons and the third particle, both scattering lengths would need to be specified.

II.1 General interaction strength

By expressing the relative wavefunction as an expansion over basis states it is possible to obtain the allowable energy states for any value of asa_{\rm s} via a matrix eigenvalue problem Liu et al. (2009); Kestner and Duan (2007); Liu et al. (2010). Due to the separability of the centre-of-mass and relative Hamiltonians the total three-body wavefunction may be written

Ψ3b(C,r,ρ)=χCOM(C)ψ3brel(r,ρ).\displaystyle\Psi_{\rm 3b}(\vec{C},\vec{r},\vec{\rho})=\chi_{\rm COM}(\vec{C})\psi_{\rm 3b}^{\rm rel}(\vec{r},\vec{\rho}). (9)

The centre of mass wavefunction, χCOM(C)\chi_{\rm COM}(\vec{C}), is a SHO wavefunction of lengthscale aCOMa_{\rm COM}. The relative wavefunction, ψ3brel(r,ρ)\psi_{\rm 3b}^{\rm rel}(\vec{r},\vec{\rho}), can be written

ψ3brel(r,ρ)\displaystyle\psi_{\rm 3b}^{\rm rel}(\vec{r},\vec{\rho}) =\displaystyle=
(1+Q^)\displaystyle(1+\hat{Q}) n=0Cnψ2brel(νnl,raμ)Rnl(ρaμ)Ylm(ρ^),\displaystyle\sum_{n=0}^{\infty}C_{n}\psi_{\rm 2b}^{\rm rel}\left(\nu_{nl},\frac{r}{a_{\mu}}\right)R_{nl}\left(\frac{\rho}{a_{\mu}}\right)Y_{lm}(\hat{\rho}),

where ψ2brel\psi_{\rm 2b}^{\rm rel} is the relative interacting two-body wavefunctions, first derived by Ref. Busch et al. (1998), and νnl\nu_{nl} are the energy pseudo-quantum numbers. RnlR_{nl} is the normalisation and radial part of the three dimensional SHO wavefunction and YlmY_{lm} is the spherical harmonic. Q^\hat{Q} is the symmetrisation operator, it ensures the correct bosonic or fermionic symmetry of the wavefunction. For three identical bosons Q^=P^13+P^23\hat{Q}=\hat{P}_{13}+\hat{P}_{23}, for two identical fermions and one distinct particle Q^=P^23\hat{Q}=-\hat{P}_{23}, where P^jk\hat{P}_{jk} exchanges the location of particles jj and kk. The CnC_{n} terms are coefficients of expansion.

The energy of ψ3brel\psi_{\rm 3b}^{\rm rel} is given by

Erel=(2νnl+2n+l+3)ω,\displaystyle E_{\rm rel}=(2\nu_{nl}+2n+l+3)\hbar\omega, (11)

where νnl+n\nu_{nl}+n is constant so that the energy of each term in Eq. (LABEL:eq:psi1) is the same.

The explicit form of the relative two-body wavefunction is Busch et al. (1998)

ψ2brel(νnl,r~)=Nνnler~2/2Γ(νnl)U(νnl,32,r~2),\displaystyle\psi_{\rm 2b}^{\rm rel}\left(\nu_{nl},\tilde{r}\right)=N_{\nu_{nl}}e^{-\tilde{r}^{2}/2}\Gamma(-\nu_{nl})U\left(-\nu_{nl},\frac{3}{2},\tilde{r}^{2}\right), (12)
Nνnl=νnlΓ(νnl1/2)2π2aμ3Γ(1νnl)[ψ(0)(νnl1/2)ψ(0)(νnl)],\displaystyle N_{\nu_{nl}}=\sqrt{\frac{\nu_{nl}\Gamma(-\nu_{nl}-1/2)}{2\pi^{2}a_{\mu}^{3}\Gamma(1-\nu_{nl})[\psi^{(0)}(-\nu_{nl}-1/2)-\psi^{(0)}(-\nu_{nl})]}},

where UU is Kummer’s function, r~=r/aμ\tilde{r}=r/a_{\mu} and ψ(0)\psi^{(0)} is the digamma function of degree 0. The explicit form of RnlR_{nl} is given by

Rnl(ρ~)\displaystyle R_{nl}\left(\tilde{\rho}\right) =\displaystyle= Nnl(ρ~)leρ~2/2Lnl+12(ρ~2),\displaystyle N_{nl}\left(\tilde{\rho}\right)^{l}e^{-\tilde{\rho}^{2}/2}L_{n}^{l+\frac{1}{2}}\left(\tilde{\rho}^{2}\right), (13)
Nnl\displaystyle N_{nl} =\displaystyle= 14πaμ62n+l+3n!(2n+2l+1)!!,\displaystyle\sqrt{\sqrt{\frac{1}{4\pi a_{\mu}^{6}}}\frac{2^{n+l+3}n!}{(2n+2l+1)!!}},

where Lnl+12L_{n}^{l+\frac{1}{2}} is the associated Laguerre polynomial, and ρ~=ρ/aμ\tilde{\rho}=\rho/a_{\mu}. The exchange operators are given by

P^23\displaystyle\hat{P}_{23} f(C,r,ρ)\displaystyle f(\vec{C},\vec{r},\vec{\rho})
=f(C,mrm+mi+γρ,(2m+mi)mir(m+mi)2γmρm+mi),\displaystyle=f\left(\vec{C},\frac{m\vec{r}}{m+m_{\rm i}}+\gamma\vec{\rho},\frac{(2m+m_{\rm i})m_{\rm i}\vec{r}}{(m+m_{\rm i})^{2}\gamma}-\frac{m\vec{\rho}}{m+m_{\rm i}}\right),

where m=m2=m3,m1=mim=m_{2}=m_{3},\quad m_{1}=m_{\rm i} and

P^13\displaystyle\hat{P}_{13} f(C,r,ρ)\displaystyle f(\vec{C},\vec{r},\vec{\rho})
=f(C,mirm+miγρ,(2m+mi)mir(m+mi)2γmρm+mi),\displaystyle=f\left(\vec{C},\frac{m_{i}\vec{r}}{m+m_{\rm i}}-\gamma\vec{\rho},\frac{-(2m+m_{\rm i})m_{\rm i}\vec{r}}{(m+m_{\rm i})^{2}\gamma}-\frac{m\vec{\rho}}{m+m_{\rm i}}\right),

where m=m1=m3,mi=m2m=m_{1}=m_{3},\quad m_{\rm i}=m_{2}. In this paper P^13\hat{P}_{13} is only applicable to the bosonic case where m1=m2=m3=mi=mm_{1}=m_{2}=m_{3}=m_{\rm i}=m.

As per Refs. Liu et al. (2009, 2010) the allowable values of the scattering length and expansion coefficients for a chosen energy (and thus chosen νnl\nu_{nl}) can be determined from the following matrix equation,

aμas[C0C1]=[X00lX01lX02lX10lX11lX12l][C0C1],\displaystyle\frac{a_{\mu}}{a_{\rm s}}\begin{bmatrix}C_{0}\\ C_{1}\\ \vdots\end{bmatrix}=\begin{bmatrix}X_{00l}&X_{01l}&X_{02l}&\dots\\ X_{10l}&X_{11l}&X_{12l}&\dots\\ \vdots&\vdots&\vdots&\ddots\\ \end{bmatrix}\begin{bmatrix}C_{0}\\ C_{1}\\ \vdots\end{bmatrix}, (16)

where

Xnnl=2Γ(νnl)Γ(νnl12)δnnη(1)lπAnnl,\displaystyle X_{n^{\prime}nl}=\frac{2\Gamma(-\nu_{n^{\prime}l})}{\Gamma(-\nu_{n^{\prime}l}-\frac{1}{2})}\delta_{n^{\prime}n}-\eta\frac{(-1)^{l}}{\sqrt{\pi}}A_{n^{\prime}nl}, (17)

and

Annl=aμ3Nνnl0ρ~2Rnl(ρ~)Rnl(κρ~1+κ)\displaystyle A_{n^{\prime}nl}=\frac{a_{\mu}^{3}}{N_{\nu_{nl}}}\int_{0}^{\infty}\tilde{\rho}^{2}R_{n^{\prime}l}(\tilde{\rho})R_{nl}\left(\frac{\kappa\tilde{\rho}}{1+\kappa}\right)
×ψ2brel(νnl,2κ+1(1+κ)2ρ~)dρ~,\displaystyle\times\psi_{\rm 2b}^{\rm rel}\left(\nu_{nl},\sqrt{\frac{2\kappa+1}{(1+\kappa)^{2}}}\tilde{\rho}\right)d\tilde{\rho}, (18)

and η=1\eta=-1 or η=2\eta=2 for the fermionic and bosonic cases respectively. With Eqs. (16)–(18) the energy spectrum of the relative three-body wavefunction can be calculated for any desired value of κ\kappa.

Additionally in the κ0\kappa\rightarrow 0 limit (the heavy-impurity limit) the matrix elements can be determined analytically:

Ann,l={2nνΓ(n+3/2)Γ(n+3/2)πΓ(n+1)Γ(n+1)l=00l>0.\displaystyle A_{n^{\prime}n,l}=\begin{cases}\dfrac{2}{n^{\prime}-\nu}\sqrt{\dfrac{\Gamma(n+3/2)\Gamma(n^{\prime}+3/2)}{\pi\Gamma(n+1)\Gamma(n^{\prime}+1)}}&l=0\\ \quad 0&l>0\end{cases}.\quad\quad (19)

However in the heavy substrate limit, κ\kappa\rightarrow\infty, Eq. (18) diverges due to to ψ2brel(ν,r)\psi_{\rm 2b}^{\rm rel}(\nu,r) diverging as r0r\rightarrow 0. It should be noted that the κ\kappa\rightarrow\infty limit can be evaluated using the Born-Oppenheimer approximation Petrov (2012).

The three-body energy spectrum is given in Figs. 1, 2, and 3. In Fig. 1 we present the fermionic energy spectrum for κ0\kappa\rightarrow 0 with l=0l=0, for comparison, we also show our results for κ=1\kappa=1, which reproduce the results of Refs. Liu et al. (2010, 2009). In Fig. 2 we present the bosonic energy spectrum for l=0l=0 and l=1l=1, recall κ=1\kappa=1 for three identical bosons. In Fig. 3 we present the fermionic energy spectrum for κ=1\kappa=1 and κ=13.75\kappa=13.75 with l=1l=1. In all figures the horizontal black lines define the non-interacting energies and the vertical black lines are simply to indicate unitarity (as)(a_{\rm s}\rightarrow\infty).

In Fig 1 we observe for κ=1\kappa=1 and as1>0a_{\rm s}^{-1}>0 sharp and close anticrossings are present which make it difficult to clearly identify which unitary states ultimately diverge and which converge as as1+a_{\rm s}^{-1}\rightarrow+\infty. However in the κ0\kappa\rightarrow 0 limit these anticrossings disappear. This makes the identification of which states are ultimately divergent and convergent much more reliable. Notice that for every unitary energy there is one divergent state and the multiplicity at unitarity increases by one every 4ω4\hbar\omega. This is consistent with Eq. (33) in Section II.2 and implies that the s00s_{00} states (see Section II.2), and only the s00s_{00} states, are divergent as as+0a_{\rm s}\rightarrow+0. This is similar to findings of Ref. Liu et al. (2010) for the κ=1\kappa=1 case where the sn=0,ls_{n=0,l} states are identified as the divergent states.

In Fig. 2, the bosonic case, we observe many of the same features as the fermionic spectra in Fig. 1. Again all states converge to non-interacting energies for as1a_{\rm s}^{-1}\rightarrow-\infty but for as1+a_{\rm s}^{-1}\rightarrow+\infty there are both divergent and convergent states, additionally anticrossings are present. For l=0l=0 we are able to identify the s10s_{10} states as being divergent in the as1+a_{\rm s}^{-1}\rightarrow+\infty, however, unlike the l=0l=0 fermion spectra discussed above, these are not the only divergent states, all Efimov states (discussed in Section II.3) are divergent in the as1+a_{\rm s}^{-1}\rightarrow+\infty limit. For l=1l=1, anticrossings are still present but fewer in number and narrower. The lowest sn=0,ls_{n=0,l} states being divergent no longer appears to be true, the (q,s11)=(0,6.462)(q,s_{11})=(0,6.462\dots) state is divergent and the (q,s01)=(2,2.864)(q,s_{01})=(2,2.864\dots) state is not divergent.

In Fig. 3 for κ=13.75\kappa=13.75 we again observe anticrossings and that the states corresponding to the smallest real value of snls_{nl} diverge as as1+a_{\rm s}^{-1}\rightarrow+\infty. However, unlike the bosonic case some, but curiously not all, Efimov states diverge as as1+a_{\rm s}^{-1}\rightarrow+\infty. Only the lowest three Efimov energies are divergent in this limit.

Refer to caption
Figure 1: Energy spectra of the fermionic three-body relative wavefunction, Eq. (LABEL:eq:psi1), for l=0l=0 calculated using Eq. (16) with a 50×5050\times 50 matrix. Blue dots correspond to κ=1\kappa=1 and red upright triangles to the κ0\kappa\rightarrow 0 limit. The magenta pluses are the unitary energies for κ=1\kappa=1 and the cyan crosses for κ0\kappa\rightarrow 0, as per the calculations in Section II.2. The solid horizontal black lines correspond to the non-interacting energies and the solid vertical black line is simply to indicate unitarity.
Refer to caption
Figure 2: Energy spectra of the bosonic three-body relative wavefunction, Eq. (LABEL:eq:psi1), calculated using Eq. (16) with a 50×5050\times 50 matrix. Blue dots correspond to l=0l=0 and green inverted triangles to l=1l=1. The black pluses are the unitary energies for l=0l=0 and the orange crosses for l=1l=1, as per the calculations in Section II.2. Some l=0l=0 unitary energies are unlabelled by a black plus and these correspond to Efimov states as described in Section II.3. The solid horizontal black lines correspond to the non-interacting energies of the l=0l=0 state and the dashed horizontal black lines to the l=1l=1 state. The solid vertical black line is simply to indicate unitarity.
Refer to caption
Figure 3: Energy spectra of the fermionic three-body relative wavefunction, Eq. (LABEL:eq:psi1), for l=1l=1 calculated using Eq. (16) with a 50×5050\times 50 matrix. Blue dots correspond to κ=1\kappa=1 and red upright triangles to κ=13.75\kappa=13.75. The magenta pluses are the unitary energies for κ=1\kappa=1 and the cyan crosses for κ=13.75\kappa=13.75, as per the calculations in Section II.2. Some κ=13.75\kappa=13.75 unitary energies are unlabelled by a cyan cross and these correspond to Efimov states as described in Section II.3. The solid horizontal black lines correspond to the non-interacting energies and the solid vertical black line is simply to indicate unitarity.

II.2 Hyperspherical approach

In Section II.1 we have derived the three-body relative wavefunction for general s-wave scattering length, asa_{\rm s}, and general mass ratio, κ\kappa, in the form of an infinite sum, Eq. (LABEL:eq:psi1). However, it is possible to derive a closed form eigenfunction of H^rel\hat{H}_{\rm rel} using hyperspherical coordinates Werner and Castin (2006a); Werner (2008). Only in the unitary and non-interacting cases can the wavefunction be fully specified. Enforcing the Bethe-Peierls boundary condition leads to a transcendental equation that can be solved for one of the quantum numbers, snls_{nl}, in the non-interacting and unitary regimes.

We define the hyperradius, RR, and the hyperangle, α\alpha,

R=r2+ρ2,α=arctan(rρ).\displaystyle R=\sqrt{r^{2}+\rho^{2}},\quad\alpha=\arctan(\frac{r}{\rho}). (20)

In turn the relative Hamiltonian is given by

H^rel\displaystyle\hat{H}_{\rm rel} =\displaystyle= 22μ[2R2+5RR+4R2\displaystyle\frac{-\hbar^{2}}{2\mu}\Bigg{[}\frac{\partial^{2}}{\partial R^{2}}+\frac{5}{R}\frac{\partial}{\partial R}+\frac{4}{R^{2}} (21)
+1R2sin2(α)cos2(α)2α2(cos(α)sin(α))\displaystyle+\frac{1}{R^{2}\sin^{2}(\alpha)\cos^{2}(\alpha)}\frac{\partial^{2}}{\partial\alpha^{2}}\Big{(}\cos(\alpha)\sin(\alpha)\mathchoice{\mathbin{\vbox{\hbox{\scalebox{0.5}{$\displaystyle\bullet$}}}}}{\mathbin{\vbox{\hbox{\scalebox{0.5}{$\textstyle\bullet$}}}}}{\mathbin{\vbox{\hbox{\scalebox{0.5}{$\scriptstyle\bullet$}}}}}{\mathbin{\vbox{\hbox{\scalebox{0.5}{$\scriptscriptstyle\bullet$}}}}}\Big{)}
Λ^ρ2R2cos2(α)Λ^r2R2sin2(α)]+μω2R22,\displaystyle-\frac{\hat{\Lambda}^{2}_{\rho}}{R^{2}\cos^{2}(\alpha)}-\frac{\hat{\Lambda}^{2}_{r}}{R^{2}\sin^{2}(\alpha)}\Bigg{]}+\frac{\mu\omega^{2}R^{2}}{2},

where Λ^ρ2\hat{\Lambda}^{2}_{\rho} and Λ^r2\hat{\Lambda}^{2}_{r} are the angular part of the three-dimensional Laplace operators acting on the ρ^\hat{\rho} and r^\hat{r} coordinate spaces respectively. In the unitary and non-interacting limits we can express the wavefunction in closed form. The unnormalised wavefunction is of the form

ψ3brel\displaystyle\psi_{\rm 3b}^{\rm rel} =\displaystyle= aμ2R2Fqsnl(R)(1+Q^)φlsnl(α)sin(2α)Ylm(ρ^).\displaystyle\frac{a_{\mu}^{2}}{R^{2}}F_{qs_{nl}}(R)(1+\hat{Q})\frac{\varphi_{ls_{nl}}(\alpha)}{\sin(2\alpha)}Y_{lm}(\hat{\rho}). (22)

Here qq, ll and snls_{nl} are quantum numbers and give the energy

Erel=(2q+l+snl+1)ω.\displaystyle E_{\rm rel}=(2q+l+s_{nl}+1)\hbar\omega. (23)

While qq and ll are non-negative integers the snls_{nl} quantum number is more complicated. Following the work of Refs. Werner and Castin (2006a); Liu et al. (2010) several conditions can be placed on the wavefunction:

φlsnl(π2)\displaystyle\varphi_{ls_{nl}}\left(\frac{\pi}{2}\right) =\displaystyle= 0,\displaystyle 0, (24)
s2φlsnl(α)\displaystyle s^{2}\varphi_{ls_{nl}}(\alpha) =\displaystyle= φlsnl′′(α)+l(l+1)cos2(α)φlsnl(α),\displaystyle-\varphi_{ls_{nl}}^{\prime\prime}(\alpha)+\frac{l(l+1)}{\cos^{2}(\alpha)}\varphi_{ls_{nl}}(\alpha), (25)
ErelFqsnl(R)\displaystyle E_{\rm rel}F_{qs_{nl}}(R) =\displaystyle= 22μ(Fqsnl′′(R)+Fqsnl(R)R)\displaystyle\frac{-\hbar^{2}}{2\mu}\left(F_{qs_{nl}}^{\prime\prime}(R)+\frac{F_{qs_{nl}}^{\prime}(R)}{R}\right) (26)
+(2snl22μR2+μω2R22)F(R).\displaystyle+\left(\frac{\hbar^{2}s_{nl}^{2}}{2\mu R^{2}}+\frac{\mu\omega^{2}R^{2}}{2}\right)F(R).

The first is enforced because a divergence at α=π/2\alpha=\pi/2 is non-physical, the second and third come from requiring that Eq. (22) is an eigenfunction of Eq. (21). These conditions determine the functional form of the wavefunction. The hyperspherical solutions are given Liu et al. (2010)

φlsnl(α)=\displaystyle\varphi_{ls_{nl}}(\alpha)=
cosl+1(α)F12(l+1snl2,l+1+snl2;l+32;cos2(α)),\displaystyle\cos^{l+1}(\alpha){}_{2}F_{1}\left(\frac{l+1-s_{nl}}{2},\frac{l+1+s_{nl}}{2};l+\frac{3}{2};\cos^{2}(\alpha)\right),
F\displaystyle F (R)=(Raμ)snleR2/2aμ2Lqsnl(R2aμ2),\displaystyle(R)=\left(\frac{R}{a_{\mu}}\right)^{s_{nl}}e^{-R^{2}/2a_{\mu}^{2}}L_{q}^{s_{nl}}\left(\frac{R^{2}}{a_{\mu}^{2}}\right), (28)

where F12{}_{2}F_{1} is the Gaussian hypergeometric function. This solution for F(R)F(R) is valid for snl20s_{nl}^{2}\geq 0. For snl2<0s_{nl}^{2}<0 there is a different solution, see Section II.3. Note that some values of snls_{nl} and ll are forbidden and are not predicted by the method of Section II.1; l=0l=0, s=2s=2 for fermions and, l=1l=1, s=3s=3 and l=0l=0, s=4s=4 for bosons are forbidden because the wavefunctions are 0 Werner (2008).

The interactions are enforced by the Bethe-Peierls boundary condition, Eq. (8), and this allows for the values of snls_{nl} to be calculated, but only in the non-interacting and unitary regimes. In the intermediate regime this formalism breaks down. Note in Eq. (26) snl2s_{nl}^{2} appears not snls_{nl}. Hence there is a degree of arbitrarity in snls_{nl}, convention is to choose snl[0,)s_{nl}\in[0,\infty) for snl20s_{nl}^{2}\geq 0 and snli[0,)s_{nl}\in i\cdot[0,\infty) for snl2<0s_{nl}^{2}<0. The snl2<0s_{nl}^{2}<0 case will be considered in more detail in Section II.3.

In the non-interacting case Eq. (8) implies φlsnl(0)=0\varphi_{ls_{nl}}(0)=0, and so the values of snls_{nl} are given Liu et al. (2010)

NIfermionssnl={2n+4l=02n+l+2l>0,\displaystyle{\rm NI\;fermions\;}s_{nl}=\begin{cases}2n+4&l=0\\ 2n+l+2&l>0\end{cases}, (29)
NIbosonssnl={2l=02n+6l=02n+l+4l=12n+l+2l>1,\displaystyle{\rm NI\;bosons\;}s_{nl}=\begin{cases}2&l=0\\ 2n+6&l=0\\ 2n+l+4&l=1\\ 2n+l+2&l>1\end{cases}, (30)

where n0n\in\mathbb{Z}_{\geq 0}.

In the unitary limit snls_{nl} will solve the following transcendental equation

0=\displaystyle 0=
dφlsnldα|α=0+η(1)l(1+κ)2κ1+2κφlsnl(arctan(1+2κκ)).\displaystyle\frac{d\varphi_{ls_{nl}}}{d\alpha}\Big{|}_{\alpha=0}+\eta(-1)^{l}\frac{(1+\kappa)^{2}}{\kappa\sqrt{1+2\kappa}}\varphi_{ls_{nl}}\left(\arctan(\frac{\sqrt{1+2\kappa}}{\kappa})\right).
(31)

See Table 1 for solutions to Eq. (31).

3 bosons 2+1 fermions κ=1\kappa=1 κ=13.75\kappa=13.75
ll nn snls_{nl}
0 0 i1.006i\cdot 1.006 2.166… 3.538…
1 4.465… 5.127… 4.802…
2 6.818… 7.114… 6.715…
3 9.324… 8.832… 10.912…
1 0 2.863… 1.77… i0.165i\cdot 0.165
1 6.462… 4.358… 3.940…
2 7.852… 5.716… 6.132…
3 9.822… 8.053… 8.211…
2 0 2.823… 3.104… 3.853…
1 5.508… 4.795… 4.965…
2 6.449… 7.238… 6.707…
3 9.272… 8.837… 8.782…
3 0 4.090… 3.959… 3.383…
1 5.771… 6.127… 6.062…
2 8.406… 7.816… 8.196…
3 9.607… 10.172… 10.200…
Table 1: Three-body wavefunction eigenvalues, snls_{nl}, at unitarity for the bosonic case and the fermionic case with κ=1\kappa=1 and κ=13.75\kappa=13.75, to three decimal places.

In the κ0\kappa\rightarrow 0 limit, with η=1\eta=-1 (fermions), Eq. (31) reduces to

0=dφlsnldα|α=0δl,0,\displaystyle 0=\frac{d\varphi_{ls_{nl}}}{d\alpha}\Big{|}_{\alpha=0}-\delta_{l,0}, (32)

in turn this implies, for κ0\kappa\rightarrow 0,

snl={4n+2l=02n+l+1l>0,\displaystyle s_{nl}=\begin{cases}4n+2&l=0\\ 2n+l+1&l>0\end{cases},\quad (33)

where n0n\in\mathbb{Z}_{\geq 0}. Note l=0l=0, s=2s=2 is normally forbidden but because s2s\rightarrow 2 as κ0\kappa\rightarrow 0 this is never violated for finite mass imbalance. In the κ\kappa\rightarrow\infty limit the right-hand term in Eq. (31) diverges unless snls_{nl} takes specific values that cause the φlsnl\varphi_{ls_{nl}} part of the term to be 0. This implies, for κ\kappa\rightarrow\infty

snl={2n+4l=02n+l+2l>0,\displaystyle s_{nl}=\begin{cases}2n+4&l=0\\ 2n+l+2&l>0\end{cases},\quad (34)

where n0n\in\mathbb{Z}_{\geq 0}.

The energies at unitarity are plotted in Figs. 1, 2 and 3. In Fig. 1 the purple pluses indicate unitary energies for the κ=1\kappa=1 case and the cyan crosses for the κ0\kappa\rightarrow 0 case. In Fig. 2 the black pluses indicate the unitary energies for l=0l=0 and the orange crosses for l=1l=1. In Fig. 3 the purple pluses indicate unitary energies for the κ=1\kappa=1 case and the cyan crosses for κ=13.75\kappa=13.75.

The energies calculated using the general method of Section II.1 are in excellent agreement with the hyperspherical approach of this section. However, as noted in the captions of Figs. 2 and 3, some states are not marked at unitarity with an energy from a hyperspherical calculation. These states are Efimov states and they correspond to the case where snls_{nl} is imaginary. In this case Eq. (28) is not the correct wavefunction and so Erel=(2q+snl+1)ωE_{\rm rel}=(2q+s_{nl}+1)\hbar\omega is not applicable.

II.3 Efimov states

The previous two sections have investigated the agreement between two different methods of calculating the unitary energy spectrum of trapped three-body systems including for bosons and the effects of mass imbalance. We find that the two methods are in excellent agreement, however, as noted in Figs. 2 and 3, there are some unitary energies predicted by the method of Section II.1 that are not predicted by Eqs. (23) and (31). These unmarked states, associated with imaginary values of snls_{nl}, are Efimov states, e.g. s00s_{00} for bosons and s01s_{01} for κ=13.75\kappa=13.75 fermions.

Eq. (26) is the Schrödinger equation for a particle moving in a two dimensions with two potential terms, a harmonic potential and a term proportional to snl2/R2s_{nl}^{2}/R^{2}. If snls_{nl} is purely imaginary then this potential is attractive at short distances rather than repulsive. This necessitates a different class of solution compared to the case where snls_{nl} is purely real and these solutions are Efimov states. Physically speaking the Efimov effect is short-range interparticle interactions giving rise to an effective attractive long-range interaction due to a third particle mediating the effective interaction between the other two.

These states were first studied in Ref. Efimov (1971) in the free context and Ref. Jonsell et al. (2002) in a trapped context. Efimov states were first observed in an ultracold gas in Ref. Kraemer et al. (2006).

For imaginary snls_{nl} the hyperradial solution is given by Werner and Castin (2006a)

F(R)=aμRWErel2ω,snl2(R2aμ2),\displaystyle F(R)=\frac{a_{\mu}}{R}W_{\dfrac{E_{\rm rel}}{2\hbar\omega},\dfrac{s_{nl}}{2}}\left(\frac{R^{2}}{a_{\mu}^{2}}\right), (35)

where Wa,b(x)W_{a,b}(x) is the Whittaker function. In the universal (snl20s_{nl}^{2}\geq 0) case Eqs. (26) and (28) imply Erel=(2q+snl+1)ωE_{\rm rel}=(2q+s_{nl}+1)\hbar\omega. However, in the Efimov case the relative energy is left a free parameter. Using the method of Section II.1 the energies of the Efimov states do not converge to a constant value as the matrix size is increased, whereas the universal states do converge. This non-convergence of the Efimov states is shown in Fig. 4, where the change in energy, ΔE=(EEN=10)/EN=10\Delta E=(E-E_{N=10})/E_{N=10} is plotted as a function of the matrix size (N)(N) in Eq. (16). Figs. 4(a)(\rm a) and 4(b)(\rm b) plot ΔE\Delta E for bosons and fermions respectively. From these plots it is clear that as NN is increased the energy for the universal states (snl2>0s_{nl}^{2}>0) states remains constant, whereas the Efimov state energies diverge. This divergence arises from the fact that the Hamiltonian and Bethe-Peierls boundary condition do not contain enough information to uniquely specify the energies of the Efimov states. We find that the Efimov energies diverge logarithmically with NN. As such an additional boundary condition is needed to fix the Efimov energies.

Refer to caption
Figure 4: The relative change in the energies of the lowest energy Efimov (blue circles) and lowest energy universal (red crosses) states calculated as per Section II.1 with increasing matrix size in Eq. (16). NN is the dimension of the matrix, i.e. the matrix is N×NN\times N, and ΔE=(EEN=10)/EN=10\Delta E=(E-E_{N=10})/E_{N=10} is the proportional deviation away from the N=10N=10 energy (universal or Efimov as appropriate). Panel (a) is the bosonic l=0l=0 case (EN=10=0.563ωE_{N=10}=-0.563\hbar\omega) and the panel (b) is the fermionic l=1l=1, κ=13.75\kappa=13.75 case (EN=10=1.507ωE_{N=10}=1.507\hbar\omega). The horizontal black line defines ΔE\Delta E=0.

The Efimov hyperradial function, Eq. (35), oscillates increasingly rapidly as R/aμ0R/a_{\mu}\rightarrow 0 (but remains bounded). To find the Efimov states we require the phase of the oscillation to be fixed, and impose the boundary condition

F(R)=R0Asin[|s0l|ln(Raμ)|s0l|ln(Rtaμ)],\displaystyle F(R)\underset{R\rightarrow 0}{=}A\sin[|s_{0l}|\ln(\frac{R}{a_{\mu}})-|s_{0l}|\ln(\frac{R_{t}}{a_{\mu}})\Big{]}, (36)

where AA is a constant and Rt>0R_{t}>0 is the three-body parameter. The short range behaviour of the Efimov hyperradial wavefunction is given by

aμRWE2ω,s0l2\displaystyle\dfrac{a_{\mu}}{R}W_{\dfrac{E}{2\hbar\omega},\dfrac{s_{0l}}{2}} (R2aμ2)=R0\displaystyle\left(\dfrac{R^{2}}{a_{\mu}^{2}}\right)\underset{R\rightarrow 0}{=}
Γ[s0l]Γ[1s0lE/ω2](Raμ)s0l\displaystyle\dfrac{\Gamma[-s_{0l}]}{\Gamma\left[\dfrac{1-s_{0l}-E/\hbar\omega}{2}\right]}\left(\dfrac{R}{a_{\mu}}\right)^{s_{0l}}
+Γ[s0l]Γ[1+s0lE/ω2](Raμ)s0l.\displaystyle+\dfrac{\Gamma[s_{0l}]}{\Gamma\left[\dfrac{1+s_{0l}-E/\hbar\omega}{2}\right]}\left(\dfrac{R}{a_{\mu}}\right)^{-s_{0l}}.

From this we can derive a transcendental equation that determines the energy spectrum as a function of RtR_{t} Jonsell et al. (2002)

|s0l|ln(Rtaμ)\displaystyle-|s_{0l}|\ln\left(\frac{R_{t}}{a_{\mu}}\right) =\displaystyle= arg{Γ[1+s0lE2]Γ[1+s0l]}modπ.\displaystyle\arg\left\{\frac{\Gamma\left[\dfrac{1+s_{0l}-E}{2}\right]}{\Gamma[1+s_{0l}]}\right\}\quad{\rm mod}\;\pi.\qquad (38)

The energy spectrum is unbounded from above and below. We index the states with qq\in\mathbb{Z}, and define the q=0q=0 state to be the first positive energy state at Rt/aμ=exp(π/|s0l|)R_{t}/a_{\mu}=\exp(\pi/|s_{0l}|) with q>0q>0 referring to higher energy states and q<0q<0 to lower energy states. For example for Rt/aμ=1R_{t}/a_{\mu}=1 and s0l=i1.006s_{0l}=i\cdot 1.006, the energies corresponding to q=3,2,1,0,1,2q=3,2,1,0,-1,-2 are given Erel/ω6.60,4.48,2.27,0.85,566,291649E_{\rm rel}/\hbar\omega\approx 6.60,4.48,2.27,-0.85,-566,-291649 respectively and at Rt/aμ=exp(π/|s0l|)R_{t}/a_{\mu}=\exp(\pi/|s_{0l}|) the energies corresponding to q=3,2,1,0,1,2q=3,2,1,0,-1,-2 are given Erel/ω8.686,6.60,4.48,2.27,0.85,566E_{\rm rel}/\hbar\omega\approx 8.686,6.60,4.48,2.27,-0.85,-566 respectively. In general the energy of the q=Nq=N state at Rt/aμ=exp(π/|s0l|)R_{t}/a_{\mu}=\exp(\pi/|s_{0l}|) is equal to the energy at Rt/aμ=1R_{t}/a_{\mu}=1 of the q=N+1q=N+1 state. For fixed RtR_{t} the states with q0q\geq 0 are have a regular spacing of 2ω\approx 2\hbar\omega whereas the spacing of the q<0q<0 states grows larger for more negative energies.

Refer to caption
Figure 5: The energy spectrum for Efimov states as defined by Eq. (38). Calculated using s=i1.006s=i\cdot 1.006. The upper limit on the horizontal axis is eπ/|s|22.7e^{\pi/|s|}\approx 22.7, and the vertical black line is Rt=aμR_{t}=a_{\mu}.

Physically speaking the additional parameter RtR_{t} is needed because in the Efimov states the specifics of the interparticle interaction become significant, due to the attractive s0l2/R2s_{0l}^{2}/R^{2} potential term, and can no longer accurately be considered a contact interaction. RtR_{t} describes the short-range (but not zero-range) properties of the interparticle interaction and will, in general, differ for different species of atoms although some work suggests that it will vary little in units of the van der Waals length in practice Wang et al. (2012).

Efimov states also exist in the fermionic case for sufficient mass imbalance Efimov (1973); Kartavtsev and Malykh (2007). By solving Eq. (31) for κ\kappa, with snl=0s_{nl}=0 and η=1\eta=-1, we find the mass imbalance for which snls_{nl} becomes imaginary and Efimov states are then allowed. There are no Efimov states for ll is even. They appear for l=1,3,5l=1,3,5\dots for κ13.606,75.994,187.958\kappa\gtrsim 13.606\dots,75.994\dots,187.958\dots respectively. This is in good agreement with previous work (Kartavtsev and Malykh, 2007, 2008; Endo et al., 2011; Petrov, 2012). Using the method of Section II.1, the Efimov states appear (or do not appear as appropriate) at these predicted values of ll and κ\kappa. Additionally while the real values of snls_{nl} converge in the κ\kappa\rightarrow\infty limit the imaginary solutions to Eq. (31) do not converge.

As described above; in the method of Section II.1 the Efimov energies depend on matrix size and are divergent with increasing matrix size, and in the approach of Section II.2 the Efimov energies are not fixed so an additional parameter is required. In fact these two methods, the matrix size and the three-body parameter, are closely linked; for every finite matrix size there is a value of RtR_{t} that produces the same unitary Efimov energy spectrum to a high degree of accuracy. This result implies that the boundary condition Eq. (36) is enforced by the finite matrix size of Eq. (16). The errors between the Efimov energies calculated with the two methods are given in Tables 2 and 3. In Tables 2 and 3 the value of RtR_{t} is calculated by substituting the lowest Efimov energy calculated with the matrix method for a given matrix size NN into Eq. (38), which corresponds to either q=0q=0 or q=1q=-1 depending on NN and the particle symmetry. The error between the two methods of calculating the Efimov energies is 1%\lesssim 1\% for all calculated values and is most likely due to the linear interpolation required to determine the unitary energies using the method of Section II.1.

However away from unitarity this equivalence no longer holds. For a general value of asa_{\rm s}, unlike at unitarity, there is no value of snls_{nl} such that there exists a value of RtR_{t} that matches the non-unitary Efimov energy spectrum of Section II.1. Repeating the above comparison between the Efimov energy spectra (for the same values of snls_{nl}) for bosons at am/as=0.25a_{\rm m}/a_{\rm s}=-0.25 instead of at unitarity lead to errors of 2%\approx 2\% compared to 1%\lesssim 1\% at unitarity. The errors grow as distance from unitarity grows, with an error of 13%\approx 13\% at am/as=5a_{\rm m}/a_{\rm s}=-5. In the fermionic case we have similar results with errors increasing as one moves further from unitarity. The equivalence breaks down in the intermediate asa_{\rm s} regime because snls_{nl} is no longer well defined. Away from the unitary or non-interacting regimes the left-hand side of Eq. (31) depends on ρ\rho and Eq. (22) cannot be used to describe the three-body relative wavefunction. As Eq. (38) depends on snls_{nl} it cannot be used to create an energy spectrum in the intermediate asa_{\rm s} regime.

In drawing an equivalence between these two methods of calculating the unitary Efimov energy spectrum we must note that the spectrum determined by Eq. (38) is unbounded above and below whereas the Efimov energy spectrum defined by Eq. (16) is unbounded above but not below. In Figs. 2 and 3 one can see that the Efimov energies approach non-interacting energies in the as1a_{\rm s}^{-1}\rightarrow-\infty limit. In order to preserve the appropriate multiplicity of the non-interacting eigenstates, the lowest observed energy states must in fact be the lowest energy states. That being said, the lower energies predicted by Eq. (38) are significantly lower, e.g. for Rt=11.555R_{t}=11.555 the q=1q=-1 state has E4.164ωE\approx-4.164\hbar\omega as per Table 2 but for q=2q=-2 we have E566ωE\approx-566\hbar\omega, and this makes direct confirmation of the absence of these states in the formulation of Eq. (16) numerically challenging.

In the hyperspherical formulation Eqs. (LABEL:eq:HyperangularWavefunc) and (28) define an orthogonal set of states in both the unitary and non-interacting regimes. Integrating over the universal hyperradial wavefunction gives an orthogonality in qq and integrating over the hyperangular part gives an orthogonality in snls_{nl} Fedorov and Jensen (1993). The Efimov hyperradial wavefunctions are orthogonal with the energy spectrum defined by Eq. (38).

The orthogonality of the Efimov wavefunctions is not obvious and here we provide a short proof. Consider some energy EE and some other energy EEE^{\prime}\neq E both of which satisfy Eq. (38) for the same values of RtR_{t} and snl=ss_{nl}=s. The overlap of two Efimov hyperradial wavefunctions is given by Gradshteyn and Ryzhik (2014); Kerin and Martin (2022b)

Fqs|Fqs=\displaystyle\bra{F_{q^{\prime}s}}\ket{F_{qs}}= aμ22Γ(s+1)Γ(s)Γ(1E/ωs2)Γ(3E/ω+s2)F12(1,1E/ω+s2;3E/ω+s2;1)\displaystyle\frac{a_{\mu}^{2}}{2}\frac{\Gamma\left(s+1\right)\Gamma(-s)}{\Gamma\left(\dfrac{1-E/\hbar\omega-s}{2}\right)\Gamma\left(\dfrac{3-E^{\prime}/\hbar\omega+s}{2}\right)}{}_{2}F_{1}\bigg{(}1,\frac{1-E/\hbar\omega+s}{2};\frac{3-E^{\prime}/\hbar\omega+s}{2};1\bigg{)}
+\displaystyle+ aμ22Γ(1s)Γ(s)Γ(1E/ω+s2)Γ(3E/ωs2)F12(1,1E/ωs2;3E/ωs2;1),\displaystyle\frac{a_{\mu}^{2}}{2}\frac{\Gamma\left(1-s\right)\Gamma(s)}{\Gamma\left(\dfrac{1-E/\hbar\omega+s}{2}\right)\Gamma\left(\dfrac{3-E^{\prime}/\hbar\omega-s}{2}\right)}{}_{2}F_{1}\bigg{(}1,\frac{1-E/\hbar\omega-s}{2};\frac{3-E^{\prime}/\hbar\omega-s}{2};1\bigg{)}, (39)

where FqsF_{q^{\prime}s} has energy EE^{\prime} and FqsF_{qs} has energy EE. Note the identity

F12(a,b;c;1)=Γ(c)Γ(cab)Γ(ca)Γ(cb),if(c)>(a+b),\displaystyle{}_{2}F_{1}(a,b;c;1)=\frac{\Gamma\left(c\right)\Gamma\left(c-a-b\right)}{\Gamma\left(c-a\right)\Gamma\left(c-b\right)},\;{\rm if}\;\mathbb{R}(c)>\mathbb{R}(a+b),

which applies if E<EE^{\prime}<E which we can always set to be true given that Fqs|Fqs=Fqs|Fqs\bra{F_{q^{\prime}s}}\ket{F_{qs}}=\bra{F_{qs}}\ket{F_{q^{\prime}s}}. Additionally note Γ(1+z)=zΓ(z)\Gamma(1+z)=z\Gamma(z) and that the two terms in Eq. (39) are complex conjugates. We then have

Fqs(R)|Fqs(R)=aμ2×\displaystyle\bra{F_{q^{\prime}s}(R)}\ket{F_{qs}(R)}=a_{\mu}^{2}\times
Re[2ωs(EE)Γ(1s)Γ(1E/ωs2)Γ(1+s)Γ(1E/ω+s2)].\displaystyle\real\left[\frac{2\hbar\omega}{s(E^{\prime}-E)}\frac{\Gamma(1-s)}{\Gamma\left(\dfrac{1-E/\hbar\omega-s}{2}\right)}\frac{\Gamma(1+s)}{\Gamma\left(\dfrac{1-E^{\prime}/\hbar\omega+s}{2}\right)}\right].
(41)

Note that the second and third terms in Eq. (41) (or their complex conjugates) appear in the right-hand side of Eq. (38) and therefore have opposite phase (up to modulo π\pi). We can then write

BeiA\displaystyle Be^{-iA} =Γ(1s)Γ(1E/ωs2),\displaystyle=\dfrac{\Gamma(1-s)}{\Gamma\left(\dfrac{1-E/\hbar\omega-s}{2}\right)}, (42)
Cei(A+nπ)\displaystyle Ce^{i(A+n\pi)} =Γ(1+s)Γ(1E/ω+s2),\displaystyle=\dfrac{\Gamma(1+s)}{\Gamma\left(\dfrac{1-E^{\prime}/\hbar\omega+s}{2}\right)}, (43)

where A,B,CA,B,C\in\mathbb{R} and nn\in\mathbb{Z}. As such in Eq. (41) the first term in the square brackets is purely imaginary and the product of the second and third terms is purely real, hence Fqs(R)|Fqs(R)=0\bra{F_{q^{\prime}s}(R)}\ket{F_{qs}(R)}=0. Since EE and EE^{\prime} are only required to be not equal and satisfy Eq. (38) for the same RtR_{t} and snls_{nl} we have that Eq. (38) produces an orthogonal spectrum of energies.

The wavefunctions given by Eq. (LABEL:eq:psi1) and (22) both solve the three-body Hamiltonian and satisfy the Bethe-Peierls condition and reproduce the same energy spectrum in the non-interacting and unitary regimes. These are different representations of the same wavefunction and as selecting a specific RtR_{t} defines an orthogonal subset of Efimov energies a specific matrix size also selects an orthogonal subset.

Bosons q=0q=0 q=1q=1 q=2q=2 q=3q=3 q=4q=4
N=10N=10 -0.563 2.393 4.612 6.747 8.849
Rt/aμ=1.131R_{t}/a_{\mu}=1.131 -0.563 2.365 4.566 6.682 8.763
Error % 0 1.191 1.000 0.963 0.967
q=1q=-1 q=0q=0 q=1q=1 q=2q=2 q=3q=3
N=20N=20 -1.531 2.112 4.353 6.490 8.588
Rt/aμ=18.216R_{t}/a_{\mu}=18.216 -1.531 2.094 4.326 6.455 8.544
Error % 0 0.870 0.621 0.543 0.511
N=30N=30 -2.428 1.943 4.198 6.341 8.441
Rt/aμ=14.897R_{t}/a_{\mu}=14.897 -2.428 1.928 4.178 6.315 8.409
Error % 0 0.767 0.494 0.410 0.375
N=40N=40 -3.301 1.823 4.087 6.234 8.336
Rt/aμ=12.912R_{t}/a_{\mu}=12.912 -3.301 1.810 4.070 6.212 8.310
Error % 0 0.709 0.423 0.344 0.310
N=50N=50 -4.164 1.731 4.001 6.150 8.253
Rt/aμ=11.555R_{t}/a_{\mu}=11.555 -4.164 1.719 3.990 6.131 8.231
Error % 0 0.681 0.386 0.303 0.266
Table 2: Comparison of the matrix and three-body parameter Efimov energies (in units of ω\hbar\omega) for bosons. The three-body parameter is found by substituting the lowest unitary Efimov energy predicted by the matrix method into Eq. (38). For matrix size N=10N=10 this means that the q=0q=0 state is the corresponding lowest energy state, and for N20N\geq 20 q=1q=-1 is the corresponding state. All values are stated to 3 decimal places. Errors are 1%\lesssim 1\%, with the primary source of error being the accurate calculation of the matrix energies, this also affects the accuracy of RtR_{t} as RtR_{t} is chosen depending on the lowest Efimov state energy.
Fermions q=0q=0 q=1q=1 q=2q=2 q=3q=3 q=4q=4
N=10N=10 1.507 3.626 5.713 7.787 9.857
Rt/aμ=2.59×107R_{t}/a_{\mu}=2.59\times 10^{7} 1.507 3.611 5.680 7.731 9.773
Error % 0 0.380 0.573 0.721 0.8544
N=20N=20 1.433 3.525 5.588 7.640 9.685
Rt/aμ=1.84×107R_{t}/a_{\mu}=1.84\times 10^{7} 1.43 3.51 5.57 7.61 9.65
Error % 0 0.150 0.221 0.272 0.312
N=30N=30 1.397 3.477 5.531 7.574 9.611
Rt/aμ=1.50×107R_{t}/a_{\mu}=1.50\times 10^{7} 1.397 3.474 5.524 7.563 9.594
Error % 0 0.085 0.127 0.153 0.178
N=40N=40 1.374 3.447 5.496 7.535 9.567
Rt/aμ=1.30×107R_{t}/a_{\mu}=1.30\times 10^{7} 1.374 3.445 5.491 7.526 9.556
Error % 0 0.061 0.087 0.106 0.121
N=50N=50 1.355 3.425 5.463 7.491 9.515
Rt/aμ=1.14×107R_{t}/a_{\mu}=1.14\times 10^{7} 1.355 3.421 5.463 7.496 9.523
Error % 0 0.113 0.005 0.063 0.082
Table 3: Comparison of the matrix and three-body parameter Efimov energies (in units of ω\hbar\omega) for fermions with κ=13.75\kappa=13.75 and l=1l=1. The three-body parameter is found by substituting the lowest unitary Efimov energy predicted by the matrix method into Eq. (38). All values are stated to 3 decimal places. Errors are 1%\lesssim 1\%, with the primary source of error being the accurate calculation of the matrix energies, this also affects the accuracy of RtR_{t} as RtR_{t} is chosen depending on the lowest Efimov state energy.

III Conclusion

In this work we have considered exact solutions for three particles in a spherically symmetric trap. We have provided results for both fermionic systems with mass imbalance and bosonic systems where by definition there is no mass imbalance. In each case these results can be used to investigate physical phenomena such as quench dynamics Mulkerin et al. (2012a); Liu et al. (2009, 2010); Cui (2012); Stöferle et al. (2006); Rakshit et al. (2012); Kaplan and Sun (2011); Mulkerin et al. (2012b); Nascimbène et al. (2010); Ku et al. (2012); Levinsen et al. (2017) and the derivation of the equation of state of such a gas Bougas et al. (2020, 2019); Budewig et al. (2019); Kehrberger et al. (2018).

We have also revisited, for fermions with mass imbalance, and bosons, the Efimov states in such systems. Specifically we have compared solutions with the matrix approach, Section II.1, and the hyperspherical approach, Section II.2. We find in each case very good agreement between the two approaches, at unitarity, for fermions with and without mass imbalance and for mass-balanced bosons. Finally, we have investigated, in Section II.3 the emergence of Efimov states for the mass-imbalanced fermion case and the mass-balanced boson case. Remarkably we find a connection between the matrix and three-body parameter approaches in determining the energy of Efimov states at unitarity. Specifically, for the matrix approach we find the energy of an Efimov state diverges with increasing matrix size, see Fig. 4. However for every matrix size there is a value of RtR_{t} that produces the same energy spectrum, see Tables 2 and 3.

This parameterisation of the energy spectrum in the intermediate regime lays the groundwork for future calculations regarding three-body quench dynamics, virial coefficients and Tan contacts. Three-body quench dynamics calculations can currently be performed for the non-interacting to unitary (and vice-versa) quenches using the hyperspherical formalism Kerin and Martin (2022a, b), but calculations for intermediate quenches are lacking. In this work we have uniquely specified the energies and wavefunctions in the intermediate regime, the remaining barrier to calculating for an arbitrary quench is the difficulty of calculating the wavefunction overlaps using Eq. (LABEL:eq:psi1), correctly accounting for the permutation operator is quite difficult. However in the hyperspherical formalism, Eq. (22), integrating over the terms acted on by the permutation operator can be done by utilising a “kinematic rotation”, a coordinate transform on the hyperangle α\alpha which is a function of rr and ρ\rho Nielsen et al. (2001); Braaten and Hammer (2006); Thøgersen (2009); Fedorov and Jensen (2001, 1993). It should be possible to modify the “kinematic rotation” technique and, with the wavefunctions and energies now specified for general asa_{\rm s}, calculate quench observables for a general quench between intermediate scattering lengths.

IV Data Availability Statement

All data will be provided upon request to A. D. Kerin.

V Acknowledgements

A.D.K. is supported by an Australian Government Research Training Program Scholarship and by the University of Melbourne.

With thanks to Victor Colussi for enlightening conversations.

References

  • Busch et al. (1998) T. Busch, B.-G. Englert, K. Rzażewski,  and M. Wilkens, Foundations of Physics 28, 549 (1998).
  • Fedorov and Jensen (2001) D. V. Fedorov and A. Jensen, Journal of Physics A: Mathematical and General 34, 6003 (2001).
  • Werner and Castin (2006a) F. Werner and Y. Castin, Physical Review Letters 97, 150401 (2006a).
  • Kestner and Duan (2007) J. Kestner and L.-M. Duan, Physical Review A 76, 033611 (2007).
  • Daily and Blume (2010) K. M. Daily and D. Blume, Physical Review A 81, 053615 (2010).
  • Mulkerin et al. (2012a) B. C. Mulkerin, C. J. Bradly, H. M. Quiney,  and A. M. Martin, Physical Review A 86, 053631 (2012a).
  • Werner and Castin (2006b) F. Werner and Y. Castin, Physical Review A 74, 053604 (2006b).
  • Blume (2012) D. Blume, Reports on Progress in Physics 75, 046401 (2012).
  • Liu et al. (2009) X.-J. Liu, H. Hu,  and P. D. Drummond, Physical Review Letters 102, 160401 (2009).
  • Liu et al. (2010) X.-J. Liu, H. Hu,  and P. D. Drummond, Physical Review A 82, 023619 (2010).
  • Cui (2012) X. Cui, Few Body Systems 52, 65 (2012).
  • Stöferle et al. (2006) T. Stöferle, H. Moritz, K. Günter, M. Köhl,  and T. Esslinger, Physical Review Letters 96, 030401 (2006).
  • Rakshit et al. (2012) D. Rakshit, K. M. Daily,  and D. Blume, Physical Review A 85, 033634 (2012).
  • Kaplan and Sun (2011) D. B. Kaplan and S. Sun, Physical Review Letters 107, 030601 (2011).
  • Mulkerin et al. (2012b) B. Mulkerin, C. Bradly, H. Quiney,  and A. Martin, Physical Review A 85, 053636 (2012b).
  • Nascimbène et al. (2010) S. Nascimbène, N. Navon, K. Jiang, F. Chevy,  and C. Salomon, Nature 463, 1057 (2010).
  • Ku et al. (2012) M. Ku, A. Sommer, L. Cheuk,  and M. Zwierlein, Science 335, 563 (2012).
  • Levinsen et al. (2017) J. Levinsen, P. Massignan, S. Endo,  and M. M. Parish, Journal of Physics B: Atomic, Molecular and Optical Physics 50, 072001 (2017).
  • Bougas et al. (2020) G. Bougas, S. Mistakidis, G. Alshalan,  and P. Schmelcher, Physical Review A 102, 013314 (2020).
  • Bougas et al. (2019) G. Bougas, S. I. Mistakidis,  and P. Schmelcher, Physical Review A 100, 053602 (2019).
  • Budewig et al. (2019) L. Budewig, S. Mistakidis,  and P. Schmelcher, Molecular Physics 117, 2043 (2019).
  • Kehrberger et al. (2018) L. M. A. Kehrberger, V. J. Bolsinger,  and P. Schmelcher, Physical Review A 97, 013606 (2018).
  • D’Incao et al. (2018) J. P. D’Incao, J. Wang,  and V. E. Colussi, Physical Review Letters 121, 023401 (2018).
  • Jonsell et al. (2002) S. Jonsell, H. Heiselberg,  and C. Pethick, Physical Review Letters 89, 250401 (2002).
  • Blume and Daily (2010) D. Blume and K. M. Daily, Physical review letters 105, 170403 (2010).
  • Efimov (1973) V. Efimov, Nuclear Physics A 210, 157 (1973).
  • Efimov (1970) V. N. Efimov, WEAKLY BOUND STATES OF THREE RESONANTLY INTERACTING PARTICLES., Tech. Rep. (Ioffe Inst. of Physics and Tech., Leningrad, 1970).
  • Kartavtsev and Malykh (2007) O. Kartavtsev and A. Malykh, Journal of Physics B: Atomic, Molecular and Optical Physics 40, 1429 (2007).
  • Kartavtsev and Malykh (2008) O. Kartavtsev and A. Malykh, JETP letters 86, 625 (2008).
  • Endo et al. (2011) S. Endo, P. Naidon,  and M. Ueda, Few-Body Systems 51, 207 (2011).
  • Wang et al. (2012) J. Wang, J. D’Incao, B. Esry,  and C. H. Greene, Physical review letters 108, 263001 (2012).
  • Efimov (1971) V. Efimov, Soviet Journal of Nuclear Physics 12, 589 (1971).
  • Kerin and Martin (2022a) A. Kerin and A. Martin, arXiv preprint arXiv:2207.09091v2  (2022a).
  • Kerin and Martin (2022b) A. Kerin and A. Martin, arXiv preprint arXiv:2208.05666  (2022b).
  • Kerin and Martin (2020) A. D. Kerin and A. M. Martin, Physical Review A 102, 023311 (2020).
  • Tan (2008a) S. Tan, Annals of Physics 323, 2952 (2008a).
  • Tan (2008b) S. Tan, Annals of Physics 323, 2971 (2008b).
  • Tan (2008c) S. Tan, Annals of Physics 323, 2987 (2008c).
  • Bethe and Peierls (1935) H. Bethe and R. Peierls, Proceedings of the Royal Society of London. Series A-Mathematical and Physical Sciences 148, 146 (1935).
  • Braaten and Hammer (2006) E. Braaten and H.-W. Hammer, Physics Reports 428, 259 (2006).
  • Petrov (2012) D. S. Petrov, Many-Body Physics With Ultracold Gases (Les Houches 2010) Lecture Notes of the Les Houches Summer School 94, 109 (2012).
  • Werner (2008) F. Werner, Theses, Université Pierre et Marie Curieâ Paris VI, Paris France  (2008).
  • Kraemer et al. (2006) T. Kraemer, M. Mark, P. Waldburger, J. G. Danzl, C. Chin, B. Engeser, A. D. Lange, K. Pilch, A. Jaakkola, H.-C. Nägerl, et al., Nature 440, 315 (2006).
  • Fedorov and Jensen (1993) D. Fedorov and A. Jensen, Physical review letters 71, 4103 (1993).
  • Gradshteyn and Ryzhik (2014) I. S. Gradshteyn and I. M. Ryzhik, Table of integrals, series, and products (Academic press, 2014).
  • Nielsen et al. (2001) E. Nielsen, D. V. Fedorov, A. S. Jensen,  and E. Garrido, Physics Reports 347, 373 (2001).
  • Thøgersen (2009) M. Thøgersen, arXiv preprint arXiv:0908.0852  (2009).