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

Effects of Efimov states on quench dynamics in a three-boson trapped system

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

We investigate the effects of Efimov states on the post-quench dynamics of a system of three identical bosons with contact interactions, in a spherically-symmetric three-dimensional harmonic trap. The quench we consider is in the s-wave contact interaction and we focus on quenches from the non-interacting to strongly interacting regimes and vice-versa. The calculations use the hyperspherical solutions of the three-body problem enable us to evaluate the semi-analytical results of the Ramsey and particle separation, post quench. In the case where the interactions are quenched from the non-interacting to strongly interacting regime we find convergent aperiodic solutions for both the Ramsey signal and the particle separation. In contrast for quenches from the strongly interacting regime to the non interacting regime both the Ramsey signal and particle separation are periodic functions. However, in this case we find that the solutions for the particle separation diverges indicating that in such a system large oscillations may be observable.

I Introduction

Efimov states are a unique type of many-body quantum state where short-range interactions create effective long-range forces due to the exchange interaction. They were first predicted by Efimov Efimov (1971) and first observed by Kraemer et al. Kraemer et al. (2006). Efimov states appear in systems of as few as three bodies (Kartavtsev and Malykh, 2007, 2008; Endo et al., 2011; Petrov, 2012; Werner, 2008; Jonsell et al., 2002; Kerin and Martin, 2022a) and while the circumstances under which Efimov states appear is well understood the specifics of those states, e.g. their energies, can vary depending on the specifics of the system. The influence of Efimov states is highly relevant to a number of topics of cold-gas research including their effects on, among other quantities, the two and three-body contacts or three-body decay rates after a quench Fletcher et al. (2017); Colussi et al. (2018); Eigen et al. (2018); D’Incao et al. (2018); Colussi et al. (2019); Musolino et al. (2019, 2022).

In this work we consider the dynamics of a system of three identical bosons, interacting via a contact interaction, in an isotropic harmonic trap. Such systems can be constructed in experiment Serwane et al. (2011); Murmann et al. (2015); Zürn et al. (2013, 2012); Stöferle et al. (2006) in the form of dilute ultracold gases. Specifically, we consider the dynamics of the system after the a quench in the contact s-wave interactions. We consider two quench pathways, from the non-interacting regime to the strongly interacting (unitary) regime and vice-versa. We utilise known solutions of the static case Cui (2012); D’Incao et al. (2018); Jonsell et al. (2002); Blume and Daily (2010); Kerin and Martin (2022a); Kestner and Duan (2007) to calculate the Ramsey signal and particle separation as functions of time following the quench and investigate the effects of different Efimov energy spectra upon the dynamics. Such solutions have been used to calculate thermodynamics quantities such as virial coefficients or Tan contacts Liu et al. (2009, 2010); Stöferle et al. (2006); Cui (2012); Rakshit et al. (2012); Kaplan and Sun (2011); Mulkerin et al. (2012a, b); Nascimbène et al. (2010); Ku et al. (2012); Levinsen et al. (2017); Daily and Blume (2010); Bougas et al. (2021); Colussi (2019); Colussi et al. (2019); Enss et al. (2022). In this work we consider a three-dimensional system complementing previous investigations into quench dynamics in two-dimensional Bougas et al. (2022) and one-dimensional Pecak et al. (2016); Volosniev (2017); Kehrberger et al. (2018); Sowiński and García-March (2019) systems.

We note that the predictions in this paper are experimentally testable with current techniques. Notably Ref. Guan et al. (2019) prepared a harmonically trapped system of two 6Li atoms, quenched in trap geometry and measured the particle separation. It is also possible to experimentally obtain systems of three harmonically trapped atoms Serwane et al. (2011); Murmann et al. (2015); Zürn et al. (2013, 2012); Stöferle et al. (2006) and the quench in s-wave scattering length is possible using tools such as Feshbach resonance Fano (1935); Feshbach (1958); Tiesinga et al. (1993); Chin et al. (2010). Additionally experiments measuring the Ramsey signal of trapped cold gases after a quench have been performed Cetina et al. (2016).

This paper is structured in the following way. Sec. II provides an overview of the hyperspherical solution to the problem of three identical bosons in a spherical harmonic trap interacting via a contact interaction, including a review of Efimov states. In Sec. III we use the static solutions to calculate observables of the post-quench system. We consider the non-interacting to unitary (forwards) and vice versa (backwards) quenches. In these two cases we calculate the Ramsey signal, the overlap of the pre- and post-quench states, and the expectation of the particle separation. For the forwards quench we find that both quantities can be calculated semi-analytically and in the reverse case the Ramsey signal is still calculable but the particle separation diverges.

II Overview of the Three-Body Problem

To begin, the Hamiltonian of three identical non-interacting bodies in an isotropic three-dimensional harmonic trap is

H^=k=13[22mk2+mω2rk22],\displaystyle\hat{H}=\sum_{k=1}^{3}\left[\frac{-\hbar^{2}}{2m}\nabla_{k}^{2}+\frac{m\omega^{2}r_{k}^{2}}{2}\right], (1)

where rk\vec{r}_{k} is the position of the kthk^{\rm th} particle, mm is the particle mass, and ω\omega is the trapping frequency. For convenience we define the length scales

aμ=μω,aM=Mω,\displaystyle a_{\mu}=\sqrt{\frac{\hbar}{\mu\omega}},\quad a_{M}=\sqrt{\frac{\hbar}{M\omega}}, (2)

where μ=m/2\mu=m/2, and M=3mM=3m.

We use the Bethe-Peierls boundary condition to model the contact interactions Bethe and Peierls (1935)

limrij0[d(rijΨ)drij1rijΨ]rij0=1as,\displaystyle\lim_{r_{ij}\rightarrow 0}\left[\frac{d(r_{ij}\Psi)}{dr_{ij}}\frac{1}{r_{ij}\Psi}\right]_{r_{ij}\rightarrow 0}=\frac{-1}{a_{\rm s}}, (3)

where Ψ\Psi 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 wavefunction of three identical harmonically trapped atoms subject to Eq. (3) is known Werner and Castin (2006a); Liu et al. (2009). In particular the hyperspherical formulation Werner and Castin (2006a) is a closed form description of the wavefunction in the strongly interacting (unitary) and non-interacting regimes.

We define the hyperradius RR and hyperangle α\alpha

R2=r2+ρ2,α=arctan((r/ρ)),\displaystyle R^{2}=\sqrt{r^{2}+\rho^{2}},\quad\alpha=\arctan{(r/\rho)}, (4)

where

r\displaystyle\vec{r} =\displaystyle= r2r1,\displaystyle\vec{r}_{2}-\vec{r}_{1}, (5)
ρ\displaystyle\rho =\displaystyle= 23(r3r1+r22),\displaystyle\frac{2}{\sqrt{3}}(\vec{r}_{3}-\frac{\vec{r}_{1}+\vec{r}_{2}}{2}), (6)

and

C\displaystyle\vec{C} =\displaystyle= r1+r2+r33\displaystyle\frac{\vec{r}_{1}+\vec{r}_{2}+\vec{r}_{3}}{3} (7)

is the centre-of-mass (COM) coordiante. The COM Hamiltonian is a simple harmonic oscillator (SHO) Hamiltonian. The COM wavefunction is unaffected by Eq. (3) and is a SHO wavefunction of argument C\vec{C} and lengthscale aMa_{M}. In hyperspherical coordinates the relative Hamiltonian is given

H^rel\displaystyle\hat{H}_{\rm rel} =\displaystyle= 22μ(2R+1R2sin(α)cos(α)2α2(cos(α)sin(α))\displaystyle\frac{-\hbar^{2}}{2\mu}\Bigg{(}\frac{\partial^{2}}{\partial R}+\frac{1}{R^{2}\sin(\alpha)\cos(\alpha)}\frac{\partial^{2}}{\partial\alpha^{2}}\left(\cos(\alpha)\sin(\alpha)\right)
+\displaystyle+ 5RR4R2Λ^r2R2sin(α)Λ^ρ2R2cos(α))+μω2R22.\displaystyle\frac{5}{R}\frac{\partial}{\partial R}-\frac{4}{R^{2}}-\frac{\hat{\Lambda}_{r}^{2}}{R^{2}\sin(\alpha)}-\frac{\hat{\Lambda}_{\rho}^{2}}{R^{2}\cos(\alpha)}\Bigg{)}+\frac{\mu\omega^{2}R^{2}}{2}.

We define an ansatz wavefunction of the form

ψ3brel\displaystyle\psi_{\rm 3brel} =\displaystyle= NqlsFqs(R)R2(1+P^13+P^23)φls(α)sin(2α)Ylm(ρ^),\displaystyle N_{qls}\frac{F_{qs}(R)}{R^{2}}(1+\hat{P}_{13}+\hat{P}_{23})\frac{\varphi_{ls}(\alpha)}{\sin(2\alpha)}Y_{lm}(\hat{\rho}),\qquad (9)

where NqlsN_{qls} is the normalisation constant, FqsF_{qs} is the hyperradial wavefunction and ϕls=(1+P^13+P^23)φls(α)Ylm(ρ^)/sin(2α)\phi_{ls}=(1+\hat{P}_{13}+\hat{P}_{23})\varphi_{ls}(\alpha)Y_{lm}(\hat{\rho})/\sin(2\alpha) is the hyperangular wavefunction. The exchange operators P^13\hat{P}_{13} and P^23\hat{P}_{23} exchange the positions of particles one and three and particles two and three respectively.

Requiring the ansatz to be an eigenfunction of the Hamiltonian leads to the hyperangular and hyperradial equations

s2φls(α)\displaystyle s^{2}\varphi_{ls}(\alpha) =\displaystyle= φls′′(α)+l(l+1)cos2(α)φls(α),\displaystyle-\varphi_{ls}^{\prime\prime}(\alpha)+\frac{l(l+1)}{\cos^{2}(\alpha)}\varphi_{ls}(\alpha), (10)
ErelFqs(R)\displaystyle E_{\rm rel}F_{qs}(R) =\displaystyle= 24μ(Fqs′′(R)+Fqs(R)R)\displaystyle\frac{-\hbar^{2}}{4\mu}\left(F_{qs}^{\prime\prime}(R)+\frac{F_{qs}^{\prime}(R)}{R}\right) (11)
+\displaystyle+ (2s24μR2+μω2R2)Fqs(R),\displaystyle\left(\frac{\hbar^{2}s^{2}}{4\mu R^{2}}+\mu\omega^{2}R^{2}\right)F_{qs}(R),

and noting that a divergence at α=π/2\alpha=\pi/2 is non-physical gives the condition

φls(π2)\displaystyle\varphi_{ls}\left(\frac{\pi}{2}\right) =\displaystyle= 0.\displaystyle 0. (12)

Eqs. (10)-(12) determine the form of Fqs(R)F_{qs}(R) and φls(α)\varphi_{ls}(\alpha) Werner and Castin (2006a, b); Liu et al. (2010)

Fqs(R)\displaystyle F_{qs}(R) =\displaystyle= {(R~)seR~2/2Lqs(R~2),s2>01R~WErel2ω,s2(R~2),s2<0,\displaystyle\begin{cases}\left(\tilde{R}\right)^{s}e^{-\tilde{R}^{2}/2}L_{q}^{s}\left(\tilde{R}^{2}\right),&s^{2}>0\\ &\\ \dfrac{1}{\tilde{R}}W_{\dfrac{E_{\rm rel}}{2\hbar\omega},\dfrac{s}{2}}(\tilde{R}^{2}),&s^{2}<0\end{cases}, (13)
φls(α)\displaystyle\varphi_{ls}(\alpha) =\displaystyle= cosl+1(α)\displaystyle\cos^{l+1}(\alpha)
×\displaystyle\times F12(l+1s2,l+1+s2;l+32;cos2(α)),\displaystyle{}_{2}F_{1}\left(\frac{l+1-s}{2},\frac{l+1+s}{2};l+\frac{3}{2};\cos^{2}(\alpha)\right),

where LqsL_{q}^{s} is the associated Laguerre polynomial, WErel/2ω,s/2W_{E_{\rm rel}/2\hbar\omega,s/2} is the Whittaker function, F12{}_{2}F_{1} is the hypergeometric function, R~=R/aμ\tilde{R}=R/a_{\mu}, q0q\in\mathbb{Z}_{\geq 0}, and s2s^{2}\in\mathbb{R} are the energy quantum numbers and l0l\in\mathbb{Z}_{\geq 0} is the angular momentum quantum number. In the rest of this work we consider only the l=0l=0 case for reasons that are elucidated in the appendix. As such we omit angular momentum indices (ll) in subsequent notation.

In this framework the ss-eigenvalues can only be fully specified in the non-interacting and unitary regime. In the unitary limit applying Eq. (3) to Eq. (9) gives the transcendental equation

0=dφs(α)dα|α=0+83φs(π3),\displaystyle 0=\frac{d\varphi_{s}^{\prime}(\alpha)}{d\alpha}\Big{|}_{\alpha=0}+\frac{8}{\sqrt{3}}\varphi_{s}\left(\frac{\pi}{3}\right), (15)

which determines the ss-eigenvalues, some solutions are presented below in Table 1. In the non-interacting limit applying Eq. (3) to Eq. (9) gives ss as

s={22n+6,\displaystyle s=\begin{cases}2\\ 2n+6\\ \end{cases}, (16)

for l=0l=0, where n0n\in\mathbb{Z}_{\geq 0}.

nn ss
0 ii\cdot 1.006…
1 4.465…
2 6.818…
3 9.324…
Table 1: The three-body ss-eigenvalues at unitarity for the 3 boson case for l=0l=0 to three decimal places.

For s2>0s^{2}>0, the universal case, the energy of the wavefunction is Erel=(2q+s+1)ωE_{\rm rel}=(2q+s+1)\hbar\omega, which is implicitly determined by Eq. (11), recall q0q\in\mathbb{Z}_{\geq 0}. For s2<0s^{2}<0, the Efimov case, the energy is not uniquely determined by requiring the wavefunction, Eq. (9), be an eigenfunction of the Hamiltonian, Eq. (LABEL:eq:HypersphericalHamiltonian), it is instead a free parameter. Hence we require an additional condition to fix the energy. The Efimov hyperradial wavefunction oscillates increasingly rapidly as R0R\rightarrow 0 and we set a condition to fix the phase of the oscillation Werner and Castin (2006a); Jonsell et al. (2002)

argΓ\displaystyle\arg\Gamma [1+sErel/ω2]\displaystyle\left[\frac{1+s-E_{\rm rel}/\hbar\omega}{2}\right] (17)
=|s|ln(Rtaμ)+argΓ(1+s)modπ,\displaystyle=-|s|\ln(\frac{R_{t}}{a_{\mu}})+\arg\Gamma(1+s)\mod\pi,

where RtR_{t} is the three-body parameter, an arbitrary parameter with units of distance. RtR_{t} determines the energies of the Efimov states. Physically speaking RtR_{t} is required because, in the Efimov case, Eq. (11) has an attractive potential term proportional to 1/R21/R^{2} which allows for arbitrarily small interparticle distances. At small distances the contact interaction assumption breaks down and the short range nature of the interaction become significant. The Efimov energies are plotted as a function of RtR_{t} in Fig. 1. The energy spectrum is unbounded from below and above, we label the states with qq\in\mathbb{Z}, defining the q=0q=0 state to be the lowest energy state with Erel>0E_{\rm rel}>0 at Rt=exp(π/|s|)aμR_{t}=\exp(\pi/|s|)a_{\mu}. Note that the energy of the q=Nq=N state evaluated at Rt=aμR_{t}=a_{\mu} is equal to the energy of the q=N1q=N-1 state evaluated at Rt=exp(π/|s|)aμR_{t}=\exp(\pi/|s|)a_{\mu}.

Refer to caption
Figure 1: The energy spectrum for Efimov states as defined by Eq. (17). Calculated using s=i1.006s=i\cdot 1.006\dots. The upper limit on the horizontal axis is Rt/aμ=eπ/|s|22.7R_{t}/a_{\mu}=e^{\pi/|s|}\approx 22.7, and the vertical black line is Rt=aμR_{t}=a_{\mu}.

III Quench Dynamics

In this paper we calculate the Ramsey signal and particle separation after a quench in asa_{s}. As part of this we need to calculate various integrals of the hyperspherical wavefunction and the details of these are presented in the appendix. The COM wavefunction is unaffected by Eq. (3) and so is unaffected by a quench in asa_{\rm s}. As such only the relative motion impacts the system behaviour.

The time-dependent post-quench relative wavefunction is given

|ψ(t)\displaystyle\ket{\psi(t)} =\displaystyle= eiH^relt/|Fqisiϕsi\displaystyle e^{-i\hat{H}_{\rm rel}t/\hbar}\ket{F_{q_{\rm i}s_{\rm i}}\phi_{s_{\rm i}}} (18)
=\displaystyle= q,sFqsϕs|FqisiϕsieiEqst/|Fqsϕs,\displaystyle\sum_{q,s}\bra{F_{qs}\phi_{s}}\ket{F_{q_{\rm i}s_{\rm i}}\phi_{s_{\rm i}}}e^{-iE_{qs}t/\hbar}\ket{F_{qs}\phi_{s}},\quad

where H^rel\hat{H}_{\rm rel} is the post-quench relative Hamiltonian, quantum numbers with subscript i refer to the initial state, quantum numbers with no subscripts are the post-quench eigenvalues and EqsE_{qs} are the post-quench eigenenergies.

III.1 Ramsey signal

The Ramsey signal Cetina et al. (2016) is defined as the wavefunction overlap of the initial and final states,

S(t)\displaystyle S(t) =\displaystyle= Ψi(t)|Ψ(t)=j=0|Ψi(0)|Ψj|2ei(EjEi)t/,\displaystyle\bra{\Psi_{\rm i}(t)}\ket{\Psi^{\prime}(t)}=\sum_{j=0}^{\infty}|\bra{\Psi_{\rm i}(0)}\ket{\Psi^{\prime}_{j}}|^{2}e^{-i(E^{\prime}_{j}-E_{\rm i})t/\hbar},\qquad (19)

where Ψi\Psi_{\rm i} is the initial state with energy EiE_{\rm i}, Ψ\Psi^{\prime} is the post-quench state and the Ψj\Psi^{\prime}_{j}s are the eigenstates of the post-quench system with energy EjE^{\prime}_{j}, jj is summing over all post-quench eigenstates.

As mentioned above the COM wavefunction is unaffected by the quench and integrates to one. Hence the Ramsey signal depends only on the relative wavefunction,

S(t)\displaystyle S(t) =\displaystyle= q,s|Fqisiϕsi|Fqsϕs|2ei(EqsEqisi)t/.\displaystyle\sum_{q,s}|\bra{F_{q_{\rm i}s_{\rm i}}\phi_{s_{\rm i}}}\ket{F_{qs}\phi_{s}}|^{2}e^{-i(E_{qs}-E_{q_{\rm i}s_{\rm i}})t/\hbar}.\quad (20)

To evaluate the Ramsey signal we need to evaluate the hyperradial integral Fqs|Fqisi\bra{F_{qs}}\ket{F_{q_{\rm i}s_{\rm i}}} and the hyperangular integral ϕs|ϕsi\bra{\phi_{s}}\ket{\phi_{s_{\rm i}}}. The appendix contains the details of the evaluation of these integrals. With the integrals known we can then calculate the Ramsey signal for the forwards and backwards quenches. There is a degree of freedom in the choice of the Efimov energy spectrum, determined by the value of RtR_{t}. Whatever the value of RtR_{t} the normalisation is preserved but the post-quench behaviour is nonetheless affected by the choice of RtR_{t}.

Refer to caption
Refer to caption
Refer to caption
Figure 2: Ramsey signal of the system quenched from non-interacting to unitarity. In each panel si=2s_{\rm i}=2 and the solid red line corresponds to qi=0q_{i}=0, the dashed green line to qi=1q_{i}=1, and the dotted blue line to qi=2q_{i}=2. The upper panel uses Rt=aμR_{t}=a_{\mu} to calculate the Efimov energy spectrum, the middle panel uses Rt=5aμR_{t}=5a_{\mu} and the lower Rt=10aμR_{t}=10a_{\mu}. These Ramsey signals are evaluated using Eq. (19) with 40 terms in each of the sums, 1600 terms total. We find that the summation is convergent.

The Ramsey signal is the weighted sum of oscillators, S(t)=Aeiat+Beibt+Ceict+S(t)=Ae^{-iat}+Be^{-ibt}+Ce^{-ict}+\dots, where the weights are the square overlaps between initial state and post-quench eigenstates and the angular frequencies are the differences between the initial energy and post-quench eigenenergies. The magnitude is similarly a weighted sum of oscillators but the angular frequencies of the oscillatory terms are the differences between post-quench eigenenergies, (ab),(bc),(ac),(a-b),(b-c),(a-c),\dots. The phase of the Ramsey signal is dominated by the phase of the most heavily weighted terms.

In Fig. 2 we plot the Ramsey signal of the forwards quench for a number of initial states and values of RtR_{t}. The calculations of the Ramsey signal for the forwards quench are performed including only the q1q\geq-1 Efimov energies except for the Rt=aμR_{t}=a_{\mu} calculation which includes only the q0q\geq 0 Efimov energies. The neglected energies are significantly lower (e.g. for Rt=aμR_{t}=a_{\mu} the q=1q=-1 Efimov energy is 566ω\approx-566\hbar\omega) and do not contribute meaningfully. Unlike with the two-body case Kerin and Martin (2020) the magnitude of the Ramsey signal of the forwards quench is aperiodic. This is because the post-quench eigenenergies are irrational because the unitary ss-eigenvalues are irrational as are the Efimov energies in general. This means the angular frequencies in Eq. (20) (aa, bb, cc\dots from the previous paragraph) are irrational as are the differences between them, hence the magnitude and phase of the Ramsey signal are aperiodic.

In the Rt=aμR_{t}=a_{\mu}, (qi,si)=(0,2)(q_{\rm i},s_{\rm i})=(0,2) case (solid red line in the upper panel of Fig. 2) the post quench states with the largest overlaps are (Eq=1,s)(2.27ω,i1.006)(E_{q=1},s)\approx(2.27\hbar\omega,i\cdot 1.006), with square overlap of 0.666\approx 0.666, (Eq=0,s)(0.85ω,i1.006)(E_{q=0},s)\approx(-0.85\hbar\omega,i\cdot 1.006), with square overlap of 0.14\approx 0.14 and (q,s)=(0,4.465)(q,s)=(0,4.465\dots), with square overlap 0.105\approx 0.105. The two largest modes in the magnitude have periods of 2π/3ω\approx 2\pi/3\omega. The phase is dominated by a period of 2.7π/ω\approx 2.7\pi/\omega.

In the Rt=5aμR_{t}=5a_{\mu}, (qi,si)=(0,2)(q_{\rm i},s_{\rm i})=(0,2) case (solid red line in the middle panel of Fig. 2) the most significant terms are (Eq=0,s)(1.077ω,i1.006)(E_{q=0},s)\approx(1.077\hbar\omega,i\cdot 1.006), with a square overlap of 0.583\approx 0.583, (Eq=1,s)(3.37ω,i1.006)(E_{q=1},s)\approx(3.37\hbar\omega,i\cdot 1.006) with a square overlap of 0.24\approx 0.24 and (q,s)=(0,4.465)(q,s)=(0,4.465\dots) with a square overlap of 0.105\approx 0.105. The two largest modes in the magnitude have periods of 0.87π/ω\approx 0.87\pi/\omega and 0.45π/ω\approx 0.45\pi/\omega. The phase is dominated by the period π/ω\approx\pi/\omega.

In the Rt=10aμR_{t}=10a_{\mu}, (qi,si)=(0,2)(q_{\rm i},s_{\rm i})=(0,2) case (solid red line in the lower panel of Fig. 2) the terms with the largest overlaps are (Eq=0,s)(1.603ω,i1.006)(E_{q=0},s)\approx(1.603\hbar\omega,i\cdot 1.006) with a square overlap of 0.72\approx 0.72, (q,s)=(0,4.465)(q,s)=(0,4.465\dots) with a square overlap of 0.105\approx 0.105 and (Eq=1,s)(3.875ω,i1.006)(E_{q=1},s)\approx(3.875\hbar\omega,i\cdot 1.006) with a square overlap of 0.092\approx 0.092. This leads to two main modes in the magnitude with periods of 0.5π/ω\approx 0.5\pi/\omega and 0.9π/ω\approx 0.9\pi/\omega and the phase is dominated by the period of 1.4π/ω\approx 1.4\pi/\omega.

Refer to caption
Refer to caption
Refer to caption
Figure 3: Ramsey signal of the system quenched from unitarity to non-interacting. In each panel si=i1.006s_{\rm i}=i\cdot 1.006\dots and the solid red line corresponds to qi=0q_{i}=0, the dashed green line to qi=1q_{i}=1, and the dotted blue line to qi=2q_{i}=2. The upper panel uses Rt=aμR_{t}=a_{\mu} to calculate the Efimov energy spectrum, and thus the energy of the initial state. The middle panel uses Rt=5aμR_{t}=5a_{\mu} and the lower Rt=10aμR_{t}=10a_{\mu}. These Ramsey signals are evaluated using Eq. (19) with 40 terms in each of the sums, 1600 terms total. We find that the summation is convergent.

In Fig. 3 we plot the Ramsey signal of the backwards quench for a system initially in an Efimov state for a variety of Efimov energies. Unlike in the forwards quench the magnitude of the Ramsey signal of the backwards quench is periodic. This is because the non-interacting eigenenergies are all odd integer multiples of ω\hbar\omega. The difference between the post-quench eigenenergies are even integers leading to the magnitude having period π/ω\pi/\omega. However the phase is dominated by the largest term in Eq. (20) and the angular frequencies of each term are irrational because the initial state is an Efimov state which, in general, has an irrational energy. This leads to the irregularity in the phase.

For Rt=aμR_{t}=a_{\mu}, (Eqi=0,si)(0.850,i1.006)(E_{q_{\rm i}=0},s_{\rm i})\approx(-0.850,i\cdot 1.006) (solid red line of the upper panel of Fig. 3) the largest terms are the overlaps with (q,s)=(0,2)(q,s)=(0,2) with square overlap 0.14\approx 0.14 and (q,s)=(1,2)(q,s)=(1,2) with square overlap 0.12\approx 0.12. These terms have periods of 0.52π/ω\approx 0.52\pi/\omega and 0.34π/ω\approx 0.34\pi/\omega. For Rt=5aμR_{t}=5a_{\mu} (Eqi=0,si)(1.077,i1.006)(E_{q_{\rm i}=0},s_{\rm i})\approx(1.077,i\cdot 1.006) (solid red line of the middle panel of Fig. 3) the largest term is the overlaps with (q,s)=(0,2)(q,s)=(0,2) with square overlap 0.58\approx 0.58 and period π/ω\approx\pi/\omega. For Rt=10aμR_{t}=10a_{\mu} (Eqi=0,si)(1.602,i1.006)(E_{q_{\rm i}=0},s_{\rm i})\approx(1.602,i\cdot 1.006) (solid red line of the lower panel of Fig. 3) the largest term is the overlaps with (q,s)=(0,2)(q,s)=(0,2) with square overlap 0.72\approx 0.72 and period 1.43π/ω\approx 1.43\pi/\omega.

III.2 Particle separation

We are not limited to calculating only the Ramsey signal. It is also possible to calculate the particle separation, R~(t)\langle\tilde{R}(t)\rangle.

The expectation value of R~(t)\tilde{R}(t) is given

R~(t)=Ψ(t)|R~|Ψ(t)\displaystyle\langle\tilde{R}(t)\rangle=\bra{\Psi^{\prime}(t)}\tilde{R}\ket{\Psi^{\prime}(t)} =\displaystyle= j,jΨi(0)|ΨjΨj|Ψi(0)\displaystyle\sum_{j,j^{\prime}}\bra{\Psi_{\rm i}(0)}\ket{\Psi^{\prime}_{j}}\bra{\Psi^{\prime}_{j^{\prime}}}\ket{\Psi_{\rm i}(0)}
×Ψj|R~|Ψjei(EjEj)t/,\displaystyle\times\bra{\Psi^{\prime}_{j}}\tilde{R}\ket{\Psi^{\prime}_{j^{\prime}}}e^{-i(E_{j^{\prime}}-E_{j})t/\hbar},

where Ψi\Psi_{\rm i} is the initial pre-quench state with energy EiE_{\rm i} and Ψ(t)\Psi^{\prime}(t) is the post-quench state. Ψj\Psi^{\prime}_{j} and Ψj\Psi^{\prime}_{j^{\prime}} are eigenstates of the post-quench system with eigenenergy EjE_{j} and EjE_{j^{\prime}} respectively, with the sums over jj and jj^{\prime} taken over all post-quench eigenstates.

The COM wavefunction is independent of the interparticle interaction and does not impact the post-quench dynamics. Due to the hyperangular wavefunction’s orthogonality in ss, two sums over ss and ss^{\prime} collapse into a single sum over ss. Hence R~(t)\langle\tilde{R}(t)\rangle is given

R~(t)\displaystyle\langle\tilde{R}(t)\rangle =\displaystyle= q,qsFqisiϕsi|FqsϕsFqsϕs|Fqisiϕsi\displaystyle\sum_{q^{\prime},q}\sum_{s}\bra{F_{q_{\rm i}s_{\rm i}}\phi_{s_{\rm i}}}\ket{F_{q^{\prime}s}\phi_{s}}\bra{F_{qs}\phi_{s}}\ket{F_{q_{\rm i}s_{\rm i}}\phi_{s_{\rm i}}} (22)
×Fqsϕs|R~|Fqsϕsei(EqsEqs)t/,\displaystyle\times\bra{F_{q^{\prime}s}\phi_{s}}\tilde{R}\ket{F_{qs}\phi_{s}}e^{-i(E_{qs}-E_{q^{\prime}s})t/\hbar},

As in Eq. (20) indices with subscript i refer to the initial state and indices with no subscript refer to the post-quench eigenstates. As with the Ramsey signal all relevant integrals are presented in the appendix.

Refer to caption
Refer to caption
Refer to caption
Figure 4: R~(t)\langle\tilde{R}(t)\rangle of a system initially in the ground state quenched from non-interacting to unitarity. The upper panel corresponds to Rt=aμR_{t}=a_{\mu}, the middle to Rt=5aμR_{t}=5a_{\mu} and the lower to Rt=10aμR_{t}=10a_{\mu}. Efimov states with q1q\geq-1 are included in the calculation except Rt=aμR_{t}=a_{\mu} where Eq=1566ωE_{q=-1}\approx-566\hbar\omega does not contribute meaningfully. The dot-dashed red line corresponds to Nmax=3N_{\rm max}=3, the dashed green line to Nmax=6N_{\rm max}=6, the solid blue line to Nmax=12N_{\rm max}=12 and the dotted black line to Nmax=24N_{\rm max}=24. We find that the summation is convergent.

In Fig. 4 we plot R~(t)\langle\tilde{R}(t)\rangle for a system initially in the non-interacting ground state quenched to unitarity with the upper, middle and lower panels corresponding to Rt=aμR_{t}=a_{\mu}, Rt=5aμR_{t}=5a_{\mu}, Rt=10aμR_{t}=10a_{\mu} respectively. For Rt5aμR_{t}\geq 5a_{\mu} we include states with q1q\geq-1 in the calculations but for Rt=aμR_{t}=a_{\mu} Eq=1566ωE_{q=-1}\approx-566\hbar\omega and this state does not meaningfully contribute so we only include the q0q\geq 0 Efimov states in the calculation. In Eq. (22) terms with q=qq=q^{\prime} are constants and the ss contributions to the energies cancel out, the angular frequencies depend only on qq and qq^{\prime}. The universal state terms oscillate with an angular frequency that is an even integer multiple of ω\omega because qqq-q^{\prime} is an integer but the Efimov state terms oscillate with irrational angular frequencies because the differences between the Efimov energies, EqE_{q} and EqE_{q^{\prime}}, are irrational in general. For each plot we have calculated R~(t)\langle\tilde{R}(t)\rangle by summing up to Nmax=3,6,12,24N_{\rm max}=3,6,12,24 terms in each of the three sums in Eq. (22), and we find that the sum is convergent.

For Rt=aμR_{t}=a_{\mu} the largest oscillating terms are (q,q,s)=(q,q,s)=(0,1,i1.006)(q^{\prime},q,s)=(q,q^{\prime},s)=(0,1,i\cdot 1.006\dots) with total coefficient 0.17\approx 0.17 and Eq=1Eq=03.12ωE_{q=1}-E_{q=0}\approx 3.12\hbar\omega and (q,q,s)=(1,2,i1.006)(q,q^{\prime},s)=(1,2,i\cdot 1.006\dots) with total coefficient 0.09\approx 0.09 and Eq=2Eq=12.12ωE_{q=2}-E_{q=1}\approx 2.12\hbar\omega . This implies characteristic periods of 2π/3ω\approx 2\pi/3\omega and π/ω\approx\pi/\omega. For Rt=5aμR_{t}=5a_{\mu} the largest oscillating terms are (q,q,s)=(0,1,i1.006)(q,q^{\prime},s)=(0,1,i\cdot 1.006\dots) with total coefficient 0.3\approx 0.3 and Eq=1Eq=02.3ωE_{q=1}-E_{q=0}\approx 2.3\hbar\omega and (q,q,s)=(0,1,4.465)(q,q^{\prime},s)=(0,1,4.465) with total coefficient 0.054\approx 0.054 and associated energy difference 2ω2\hbar\omega. This leads to characteristic periods of 0.9π/ω\approx 0.9\pi/\omega and π/ω\pi/\omega. For Rt=10aμR_{t}=10a_{\mu} the largest oscillating terms are (q,q,s)=(0,1,i1.006)(q,q^{\prime},s)=(0,1,i\cdot 1.006\dots) with total coefficient 0.22\approx 0.22 and Eq=1Eq=02.3ωE_{q=1}-E_{q=0}\approx 2.3\hbar\omega and (q,q,s)=(0,1,4.465)(q,q^{\prime},s)=(0,1,4.465) with total coefficient 0.054\approx 0.054 and associated energy difference 2ω2\hbar\omega. This leads to characteristic periods of 0.9π/ω\approx 0.9\pi/\omega and π/ω\pi/\omega.

Refer to caption
Refer to caption
Refer to caption
Figure 5: R~(t)\langle\tilde{R}(t)\rangle of a system following a quench from unitarity to non-interacting. The upper, middle and lower panels correspond to Rt=aμR_{t}=a_{\mu}, Rt=5aμR_{t}=5a_{\mu} and Rt=10aμR_{t}=10a_{\mu} and the corresponding initial states are Efimov states with q=0q=0. The dot-dashed red line corresponds to Nmax=3N_{\rm max}=3, the dashed green line to Nmax=6N_{\rm max}=6, the solid blue line to Nmax=12N_{\rm max}=12 and the dotted black line to Nmax=24N_{\rm max}=24.

In Fig. 5 we plot R~(t)\langle\tilde{R}(t)\rangle for the backwards quench where the system is initially in a variety of Efimov states. Unlike the forwards quench we find that R~(t)\langle\tilde{R}(t)\rangle is periodic for the backwards quench. This is because in the backwards quench the post-quench states are universal states where the differences between eigenenergies are always even multiples of ω\hbar\omega, leading to a period of π/ω\pi/\omega. However similar to how Ref. Kerin and Martin (2020) found a divergence in r=|r2r1|r=|\vec{r}_{2}-\vec{r}_{1}| in the backwards quench we find that R~(t)\langle\tilde{R}(t)\rangle also diverges for the backwards quench. In particular we find that it is logarithmically divergent with the number of terms in the summation, i.e. R~(tnπ/ω)ln(Nmax)\langle\tilde{R}(t\neq n\pi/\omega)\rangle\propto\ln(N_{\rm max}). This divergence is not exclusively due to the Efimov states as the divergence is present even when there are no Efimov states Kerin and Martin (2022b).

This divergence is quite unusual, it is not obvious why it occurs nor why it occurs only for the reverse quench. To investigate further we look at how the probability distribution of RR, P(R,t)P(R,t), evolves over time for both quenches. P(R,t)P(R,t) is given

P(R,t)=Ψ(t)|δ(RR)|Ψ(t),\displaystyle P(R^{\prime},t)=\bra{\Psi^{\prime}(t)}\delta(R^{\prime}-R)\ket{\Psi^{\prime}(t)}, (23)

where Ψ(t)\Psi^{\prime}(t) is the post-quench state.

Refer to caption
Refer to caption
Refer to caption
Figure 6: The evolution of the probability distribution of the hyperradius, Eq. (23), for the forwards quench. The upper, middle and lower panels correspond to Rt/aμ=1,5R_{t}/a_{\mu}=1,5 and 1010 respectively. The horizontal axis is the hyperradius and the vertical axis is time with dark blue corresponding to low probability density, and yellow to high density. For all plots the initial state is (qi,si)=(0,2)(q_{\rm i},s_{\rm i})=(0,2) and each plot is constructed with Nmax=24N_{\rm max}=24. In this case P(R,t)P(R,t) is convergent with NmaxN_{\rm max}.

In Fig. 6 we plot the evolution of P(R,t)P(R,t) for the forwards quench with Rt/aμ=1,5R_{t}/a_{\mu}=1,5 and 1010 in the upper, middle and lower panels respectively. For all values of RtR_{t} we see a qualitatively similar evolution, the system oscillates between a broad distribution and a tightly peaked one with a smaller mean value. This oscillation is only approximately periodic due to the influence of the irrational Efimov energies. The broad distribution corresponds to the initial universal state and the tightly peaked distribution is dominated by the Efimov states, the system oscillates between these two regimes.

To understand this oscillation it is useful to consider R~\langle\tilde{R}\rangle for the initial and post-quench states. For example for Rt=aμR_{t}=a_{\mu} the states with the largest overlaps with the initial are (Eq=1,s)(2.27ω,i1.006)(E_{q=1},s)\approx(2.27\hbar\omega,i\cdot 1.006) with a square overlap of 0.666\approx 0.666 and (Eq=0,s)(0.85ω,i1.006)(E_{q=0},s)\approx(-0.85\hbar\omega,i\cdot 1.006) with square overlap 0.14\approx 0.14, these states have R~1.5\langle\tilde{R}\rangle\approx 1.5 and 0.57\approx 0.57 respectively. The initial state is (q,s)=(0,2)(q,s)=(0,2) so we have R~(t=0)1.66\langle\tilde{R}(t=0)\rangle\approx 1.66, hence the position of the initial broad distribution is to the right of the tightly peaked Efimov distribution. As RtR_{t} increases R~\langle\tilde{R}\rangle of the strongly overlapping Efimov states increases but they are, on average, still less than R~\langle\tilde{R}\rangle of the initial universal state, hence the peak of the narrow distribution moves rightward with increasing RtR_{t}. Additionally as RtR_{t} increases the narrow Efimov distribution broadens because the higher energy Efimov states are simply broader. Note in the third panel of Fig. 6 one can see small local peaks in probability near R~=0\tilde{R}=0. These come from the (Eq=1,s)(5.6,i1.006)(E_{q=-1},s)\approx(-5.6,i\cdot 1.006) state and have a square overlap with the initial state of 0.01\approx 0.01 and R~0.3\langle\tilde{R}\rangle\approx 0.3, the (Eq=1,s)(E_{q=-1},s) state is also accounted for in the Rt=5aμR_{t}=5a_{\mu} calculation but the overlap is approximately 50 times smaller.

Refer to caption
Refer to caption
Refer to caption
Figure 7: The evolution of the probability distribution of the hyperradius, Eq. (23), for the backward quench plotted at t=0t=0 (dot-dashed red line), t=0.17π/ωt=0.17\pi/\omega (dashed green), t=0.34π/ωt=0.34\pi/\omega (solid blue) and t=π/ωt=\pi/\omega (dotted black). The initial states in the upper, middle and lower panels are the q=0q=0 Efimov states for Rt/aμ=1,5R_{t}/a_{\mu}=1,5 and 1010 respectively. All calculations are performed with Nmax=60N_{\rm max}=60, unlike in the forwards quench we find that P(R,t)P(R,t) is only convergent for t=0t=0.
Refer to caption
Figure 8: The tail of P(R,t=π/2ω)P(R,t=\pi/2\omega) for the reverse quench with the energy of the initial Efimov state given E1.077E\approx 1.077 (q=0q=0 for Rt=5aμR_{t}=5a_{\mu}) for various values of NmaxN_{\rm max}. The dot-dashed red line corresponds to Nmax=10N_{\rm max}=10, the dashed green line to Nmax=20N_{\rm max}=20, the solid blue line to Nmax=30N_{\rm max}=30 and the dotted black line to Nmax=40N_{\rm max}=40.

In Fig, 7 we plot P(R,t)P(R,t) for the backwards quench at t=0,0.17π/ω,0.34π/ωt=0,0.17\pi/\omega,0.34\pi/\omega and π/2ω\pi/2\omega. Unlike the forwards quench the evolution of P(R,t)P(R,t) here is periodic. The mean of P(R,t)P(R,t) increases with time, reaching a maximum at π/2ω\pi/2\omega before returning to its initial shape at t=π/ωt=\pi/\omega, this then repeats with period π/ω\pi/\omega. Initially P(R,t)P(R,t) is tightly peaked but develops a long tail as it evolves and in Fig. 8 we present the long tail in detail for various values of NmaxN_{\rm max}. The tail behaves approximately like R~2\tilde{R}^{-2} until ending in an exponential-like “cut-off”. This cut-off occurs at larger RR for larger NmaxN_{\rm max} and in the limit of NmaxN_{\rm max}\rightarrow\infty the behaviour of the tail of P(R,t)P(R,t) approaches R~2\tilde{R}^{-2} with no cut-off. This means that the integral of P(R,t)P(R,t) over RR from R=0R=0 to RR\rightarrow\infty is finite and properly normalised in the NmaxN_{\rm max}\rightarrow\infty limit, however RP(R,t)RP(R,t) has a R~1\tilde{R}^{-1} tail and so is the integral is not finite, hence the divergence in R~(t)\langle\tilde{R}(t)\rangle for the backwards quench.

Physically speaking there are two likely candidates for the source of the divergence; the zero-range contact interaction and the instantaneous nature of the quench, in reality atoms interact at some finite range and the quench in asa_{\rm s} occurs over some finite time. These are two non-physical inputs into this model and may be responsible for the non-physical outputs.

By considering the finite range of the interaction it is possible to estimate a maximum value of R~(t)\langle\tilde{R}(t)\rangle. The lengthscale of the interaction provides a justification for a maximum energy and thus a cut-off in Eq. (22). Specifically, the range of interaction defines a minimum de Broglie wavelength which defines a maximum energy and thus the cut-off. For sodium in a 1kHz trap and assuming a van der Waals range of one nanometre we obtain an energy of Erel8.7×106ωE_{\rm rel}\approx 8.7\times 10^{6}\hbar\omega and so we predict R~(t)max21\langle\tilde{R}(t)\rangle_{\rm max}\approx 21 for an initial Efimov energy of Eq=00.85E_{q=0}\approx-0.85. This is an order of magnitude larger than the amplitude of oscillations when the system is quenched from the non-interacting to the strongly interacting regime.

In contrast it is difficult to quantify the effects of a finite duration quench. In the formalism used here only quenches between the non-interacting and unitary regimes can be described, meaning a quench to or from the intermediate regime can’t be elucidated. However in the two-body case a quench between any two scattering lengths can be considered Kerin and Martin (2020) so the effects of a finite duration quench can be investigated.

IV Conclusion

In this paper we have examined the effects of different Efimov energy spectra on the time dependent post-quench dynamics of an interacting few-body system. This was done in the context of three interacting bosons in a spherically symmetric trap, where the contact interactions were quenched from the non-interacting regime to the strongly-interacting regime (forwards quench) and vice-versa (backwards quench). In each case we were able to evaluate the post quench dynamics of both the Ramsey signal and the expectation value of the hyperradius.

For the forwards quench we find an irregularly repeating signal for both the Ramsey signal and R~(t)\langle\tilde{R}(t)\rangle. For the Ramsey signal this is due to both the Efimov energies and unitary ss-eigenspectrum being irrational in general. In the case of the particle separation the contributions from ss cancel out and the irregularity is due to the irrationality of the Efimov energies. In both cases the results are convergent and well defined.

For the backwards quench the magnitude of the Ramsey signal and R~(t)\langle\tilde{R}(t)\rangle oscillate with period π/ω\pi/\omega. This is because the non-interacting ss-eigenvalues are even integers and Efimov states are not present when as=0a_{\rm s}=0. The phase of the Ramsey signal is still irregular due to the influence of the initial irrational Efimov energy. However we find, analogous to previous results Kerin and Martin (2020), that the particle separation diverges logarithmically. By enforcing a cut-off on Eq. (22) motivated by a minimum de Broglie wavelngth derived from the van der Waals range we expect a maximum R~(t)21\langle\tilde{R}(t)\rangle\approx 21. This estimate of the size of the oscillations is extremely large compared to the forwards quench case.

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 illuminating discussions regarding the evaluation of the hyperangular integral.

Appendix

In this work we calculate quench observables of systems where the wavefunctions of the pre- and post-quench systems are known. To obtain these observables we need to perform numerous integrals involving these wavefunctions. In this appendix we present those integrals. Firstly, the Jacobian in hyperspherical coordinates is given by

dV=dr1dr2dr3=3332R5sin2(2α)dRdαdΩrdΩρdC,\displaystyle dV=d\vec{r}_{1}d\vec{r}_{2}d\vec{r}_{3}=\frac{3\sqrt{3}}{32}R^{5}\sin^{2}(2\alpha)dRd\alpha d\vec{\Omega}_{r}d\vec{\Omega}_{\rho}d\vec{C}, (24)

and for convenience we define

Fqs(R)|Fqs(R)\displaystyle\bra{F_{qs}(R)}\ket{F_{q^{\prime}s^{\prime}}(R)} =\displaystyle= 0RFqs(R)Fqs(R)𝑑R,\displaystyle\int_{0}^{\infty}RF_{qs}(R)^{*}F_{q^{\prime}s^{\prime}}(R)dR, (25)
ϕs(α)|ϕs(α)\displaystyle\bra{\phi_{s}(\alpha)}\ket{\phi_{s^{\prime}}(\alpha)} =\displaystyle= 0π/2ϕs(α)ϕs(α)2sin2(2α)𝑑α𝑑Ωr𝑑Ωρ.\displaystyle\int\int\int_{0}^{\pi/2}\phi_{s}(\alpha)^{*}\phi_{s^{\prime}}(\alpha)2\sin^{2}(2\alpha)d\alpha d\vec{\Omega}_{r}d\vec{\Omega}_{\rho}. (26)

To calculate the Ramsey signal, Eq. (20), we need the wavefunction overlaps, i.e. Fqs(R)|Fqs(R)\bra{F_{qs}(R)}\ket{F_{q^{\prime}s^{\prime}}(R)} and ϕs(α)|ϕs(α)\bra{\phi_{s}(\alpha)}\ket{\phi_{s^{\prime}}(\alpha)}. Whether ss is imaginary or not does not change the functional form of the hyperangular wavefunction, ϕs(α)\phi_{s}(\alpha), unlike the hyperradial wavefunction, Fqs(R)F_{qs}(R). For the hyperangular integral there is only one case, but for the hyperradial integral there are three; the universal-universal, universal-Efimov and Efiomv-Efimov. We begin by considering the hyperangular integral.

The presence of the permutation operators makes evaluating the hyperangular integral directly difficult. To evaluate we transform the permuted terms into the same Jacobi set as the unpermuted term Nielsen et al. (2001). However this limits us to the l=0l=0 case, if the spherical harmonic term is non-constant then the coordinate transform is more complicated and the integral becomes intractable. The hyperangular integral is given Fedorov and Jensen (1993, 2001); Braaten and Hammer (2006); Thøgersen (2009)

ϕ0s|ϕ0si\displaystyle\bra{\phi_{0s}}\ket{\phi_{0s_{\rm i}}} =\displaystyle= 8π0π/2((1+P^23+P^13)φs(α)sin(2α))((1+P^23+P^13)φsi(α)sin(2α))sin2(2α)𝑑α,\displaystyle 8\pi\int_{0}^{\pi/2}\left(\left(1+\hat{P}_{23}+\hat{P}_{13}\right)\frac{\varphi_{s}(\alpha)}{\sin(2\alpha)}\right)^{*}\left(\left(1+\hat{P}_{23}+\hat{P}_{13}\right)\frac{\varphi_{s_{\rm i}}(\alpha)}{\sin(2\alpha)}\right)\sin^{2}(2\alpha)d\alpha, (27)
=\displaystyle= 24π[0π/2φs(α)φsi(α)𝑑α+430π/2φs(α)[|π/3α|π/2|π/6α|φsi(α)𝑑α]𝑑α].\displaystyle 24\pi\Bigg{[}\int_{0}^{\pi/2}\varphi_{s}^{*}(\alpha)\varphi_{s_{\rm i}}(\alpha)d\alpha+\frac{4}{\sqrt{3}}\int_{0}^{\pi/2}\varphi_{s}^{*}(\alpha)\left[\int_{|\pi/3-\alpha|}^{\pi/2-|\pi/6-\alpha|}\varphi_{s_{\rm i}}(\alpha^{\prime})d\alpha^{\prime}\right]d\alpha\Bigg{]}.

For l=0l=0 we have Werner (2008); Fedorov and Jensen (2001)

φ0ssin(s(π2α)).\displaystyle\varphi_{0s}\propto\sin\left(s\left(\frac{\pi}{2}-\alpha\right)\right). (28)

Note that evaluating Eq. (27) with Eq. (28) does not give the same result as Ref. Werner (2008) in general. This is because Ref. Werner (2008) is firstly concerned with the overlaps with the ground state and secondly combine Eq. (27) with Eq. (15) and so the results presented here and in the latter reference agree when ss is a unitary eigenvalue and (qi,si)=(0,2)(q_{\rm i},s_{\rm i})=(0,2). Note that different hyperangular states of the same regime (i.e. two different unitary values of ss or two different non-interacting values of ss) are orthogonal, but there is non-zero overlap between unitary and non-interacting states.

The hyperradial integrals are integrals of products of well understood functions. The universal-universal integral is given Srivastava et al. (2003)

Fqs(R)|Fqisi(R)=\displaystyle\bra{F_{qs}(R)}\ket{F_{q_{\rm i}s_{\rm i}}(R)}=
aμ22(q+sq)(qi+sis21qi)Γ(s+si2+1)F23(q,s+si2+1,ssi2+1;s+1,ssi2qi+1;1),\displaystyle\frac{a_{\mu}^{2}}{2}\binom{q+s}{q}\binom{q_{\rm i}+\dfrac{s_{\rm i}-s}{2}-1}{q_{\rm i}}\Gamma\left(\frac{s+s_{\rm i}}{2}+1\right){}_{3}F_{2}\bigg{(}-q,\frac{s+s_{\rm i}}{2}+1,\frac{s-s_{\rm i}}{2}+1;s+1,\frac{s-s_{\rm i}}{2}-q_{\rm i}+1;1\bigg{)}, (29)

the Efimov-Efimov Gradshteyn and Ryzhik (2014)

Fqs(R)|Fqis(R)\displaystyle\bra{F_{qs}(R)}\ket{F_{q_{\rm i}s}(R)} =\displaystyle= aμ2Re[Γ(s+1)Γ(s)Γ(1Eqi/ωs2)Γ(3Eq/ω+s2)\displaystyle a_{\mu}^{2}\real[\frac{\Gamma\left(s+1\right)\Gamma(-s)}{\Gamma\left(\dfrac{1-E_{q_{\rm i}}/\hbar\omega-s}{2}\right)\Gamma\left(\dfrac{3-E_{q}/\hbar\omega+s}{2}\right)} (30)
×F23(s+1,1,1Eqi/ω+s2;1+s,3Eq/ω+s2;1)],\displaystyle\times{}_{3}F_{2}\bigg{(}s+1,1,\frac{1-E_{q_{\rm i}}/\hbar\omega+s}{2};1+s,\frac{3-E_{q}/\hbar\omega+s}{2};1\bigg{)}\Bigg{]},

and the universal-Efimov Gradshteyn and Ryzhik (2014)

Fqs(R)|Fqisi(R)=aμ24(1)qiΓ(1+qi)[\displaystyle\bra{F_{qs}(R)}\ket{F_{q_{\rm i}s_{\rm i}}(R)}=\frac{a_{\mu}^{2}}{4}\frac{(-1)^{q_{\rm i}}}{\Gamma(1+q_{\rm i})}\Bigg{[}
Γ(2s+si2)Γ(2+s+si2)Γ(si)Γ(qisi)Γ(3Eq/ω+si2)F23(1+sis2,1+s+si2,qi;1+si,3Eq/ω+si2;1)\displaystyle\frac{\Gamma\left(\dfrac{2-s^{*}+s_{\rm i}}{2}\right)\Gamma\left(\dfrac{2+s^{*}+s_{\rm i}}{2}\right)\Gamma(-s_{\rm i})}{\Gamma(-q_{\rm i}-s_{\rm i})\Gamma\left(\dfrac{3-E_{q}/\hbar\omega+s_{\rm i}}{2}\right)}{}_{3}F_{2}\bigg{(}1+\frac{s_{\rm i}-s^{*}}{2},1+\frac{s^{*}+s_{\rm i}}{2},-q_{\rm i};1+s_{\rm i},\frac{3-E_{q}/\hbar\omega+s_{\rm i}}{2};1\bigg{)}
+Γ(2+ssi2)Γ(2ssi2)Γ(si)Γ(qi)Γ(3Eq/ωsi2)F23(1+ssi2,1s+si2,qisi;1si,3Eq/ωsi2;1)],\displaystyle+\frac{\Gamma\left(\dfrac{2+s^{*}-s_{\rm i}}{2}\right)\Gamma\left(\dfrac{2-s^{*}-s_{\rm i}}{2}\right)\Gamma(s_{\rm i})}{\Gamma(-q_{\rm i})\Gamma\left(\dfrac{3-E_{q}/\hbar\omega-s_{\rm i}}{2}\right)}{}_{3}F_{2}\bigg{(}1+\frac{s^{*}-s_{\rm i}}{2},1-\frac{s^{*}+s_{\rm i}}{2},-q_{\rm i}-s_{\rm i};1-s_{\rm i},\frac{3-E_{q}/\hbar\omega-s_{\rm i}}{2};1\bigg{)}\Bigg{]}, (31)

where we have used the identity

Lnα(z)=(1)nn!ez/2z(n+1)/2W2n+α+12,α2(z).\displaystyle L_{n}^{\alpha}(z)=\frac{(-1)^{n}}{n!}e^{z/2}z^{-(n+1)/2}W_{\dfrac{2n+\alpha+1}{2},\dfrac{\alpha}{2}}(z). (32)

For the hyperradial integral we have that for s=ss=s^{\prime} the integral is 0 for qqq\neq q^{\prime}.

To calculate the particle separation expectation value, Eq. (22), we again need to calculate a number of integrals involving the wavefunction. All the needed integrals except Fqs|R~|Fqs\bra{F_{q^{\prime}s}}\tilde{R}\ket{F_{qs}} are given above. Previously for the hyperradial integral we had three cases, here we do not need to consider the universal-Efimov case as ss is the same in both bra and ket due to the orthogonality in ss of the hyperangular integral. For the universal-universal case we have Srivastava et al. (2003)

Fqs(R)|R~|Fqs(R)=aμ22(q+sq)(q32q)Γ(s+32)F23(q,s+32,32;s+1,32q;1),\displaystyle\bra{F_{qs}(R)}\tilde{R}\ket{F_{q^{\prime}s}(R)}=\frac{a_{\mu}^{2}}{2}\binom{q+s}{q}\binom{q^{\prime}-\frac{3}{2}}{q^{\prime}}\Gamma\left(s+\frac{3}{2}\right){}_{3}F_{2}\left(-q,s+\frac{3}{2},\frac{3}{2};s+1,\frac{3}{2}-q^{\prime};1\right), (33)

and for the Efimov-Efimov case we have Gradshteyn and Ryzhik (2014)

Fqs(R)|R~|Fqs(R)\displaystyle\bra{F_{qs}(R)}\tilde{R}\ket{F_{q^{\prime}s}(R)} =\displaystyle= aμ2Re[Γ(32+s)Γ(32)Γ(s)Γ(1Eq/ωs2)Γ(4Eq/ω+s2)\displaystyle a_{\mu}^{2}\real[\frac{\Gamma\left(\dfrac{3}{2}+s\right)\Gamma\left(\dfrac{3}{2}\right)\Gamma(-s)}{\Gamma\left(\dfrac{1-E_{q^{\prime}}/\hbar\omega-s}{2}\right)\Gamma\left(\dfrac{4-E_{q}/\hbar\omega+s}{2}\right)} (34)
×F23(32+s,32,1Eq/ω+s2;1+s,4Eq/ω+s2;1)].\displaystyle\times{}_{3}F_{2}\bigg{(}\frac{3}{2}+s,\frac{3}{2},\frac{1-E_{q^{\prime}}/\hbar\omega+s}{2};1+s,\frac{4-E_{q}/\hbar\omega+s}{2};1\bigg{)}\Bigg{]}.

References

  • Efimov (1971) V. Efimov, Soviet Journal of Nuclear Physics 12, 589 (1971).
  • 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).
  • 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).
  • 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).
  • Jonsell et al. (2002) S. Jonsell, H. Heiselberg,  and C. Pethick, Physical Review Letters 89, 250401 (2002).
  • Kerin and Martin (2022a) A. Kerin and A. Martin, arXiv preprint arXiv:2204.09205  (2022a).
  • Fletcher et al. (2017) R. J. Fletcher, R. Lopes, J. Man, N. Navon, R. P. Smith, M. W. Zwierlein,  and Z. Hadzibabic, Science 355, 377 (2017).
  • Colussi et al. (2018) V. Colussi, J. P. Corson,  and J. P. D’Incao, Physical review letters 120, 100401 (2018).
  • Eigen et al. (2018) C. Eigen, J. A. Glidden, R. Lopes, E. A. Cornell, R. P. Smith,  and Z. Hadzibabic, Nature 563, 221 (2018).
  • D’Incao et al. (2018) J. P. D’Incao, J. Wang,  and V. E. Colussi, Physical Review Letters 121, 023401 (2018).
  • Colussi et al. (2019) V. Colussi, B. van Zwol, J. D’Incao,  and S. Kokkelmans, Physical Review A 99, 043604 (2019).
  • Musolino et al. (2019) S. Musolino, V. Colussi,  and S. Kokkelmans, Physical Review A 100, 013612 (2019).
  • Musolino et al. (2022) S. Musolino, H. Kurkjian, M. Van Regemortel, M. Wouters, S. Kokkelmans,  and V. Colussi, Physical Review Letters 128, 020401 (2022).
  • Serwane et al. (2011) F. Serwane, G. Zürn, T. Lompe, T. Ottenstein, A. Wenz,  and S. Jochim, Science 332, 336 (2011).
  • Murmann et al. (2015) S. Murmann, A. Bergschneider, V. M. Klinkhamer, G. Zürn, T. Lompe,  and S. Jochim, Physical Review Letters 114, 080402 (2015).
  • Zürn et al. (2013) G. Zürn, A. N. Wenz, S. Murmann, A. Bergschneider, T. Lompe,  and S. Jochim, Physical Review Letters 111, 175302 (2013).
  • Zürn et al. (2012) G. Zürn, F. Serwane, T. Lompe, A. N. Wenz, M. G. Ries, J. E. Bohn,  and S. Jochim, Physical Review Letters 108, 075303 (2012).
  • Stöferle et al. (2006) T. Stöferle, H. Moritz, K. Günter, M. Köhl,  and T. Esslinger, Physical Review Letters 96, 030401 (2006).
  • Cui (2012) X. Cui, Few Body Systems 52, 65 (2012).
  • Blume and Daily (2010) D. Blume and K. M. Daily, Physical review letters 105, 170403 (2010).
  • Kestner and Duan (2007) J. Kestner and L.-M. Duan, Physical Review A 76, 033611 (2007).
  • 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).
  • 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. (2012a) B. Mulkerin, C. Bradly, H. Quiney,  and A. Martin, Physical Review A 85, 053636 (2012a).
  • Mulkerin et al. (2012b) B. C. Mulkerin, C. J. Bradly, H. M. Quiney,  and A. M. Martin, Physical Review A 86, 053631 (2012b).
  • Nascimbène et al. (2010) S. Nascimbène, N. Navon, F. Jiang, K. 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).
  • Daily and Blume (2010) K. Daily and D. Blume, Physical Review A 81, 053615 (2010).
  • Bougas et al. (2021) G. Bougas, S. Mistakidis, P. Giannakeas,  and P. Schmelcher, New Journal of Physics 23, 093022 (2021).
  • Colussi (2019) V. E. Colussi, Atoms 7, 19 (2019).
  • Enss et al. (2022) T. Enss, N. C. Braatz,  and G. Gori, arXiv preprint arXiv:2203.06098  (2022).
  • Bougas et al. (2022) G. Bougas, S. Mistakidis, P. Giannakeas,  and P. Schmelcher, arXiv preprint arXiv:2205.02015  (2022).
  • Pecak et al. (2016) D. Pecak, M. Gajda,  and T. Sowiński, New Journal of Physics 18, 013030 (2016).
  • Volosniev (2017) A. G. Volosniev, Few-Body Systems 58, 1 (2017).
  • Kehrberger et al. (2018) L. M. A. Kehrberger, V. J. Bolsinger,  and P. Schmelcher, Physical Review A 97, 013606 (2018).
  • Sowiński and García-March (2019) T. Sowiński and M. Á. García-March, Reports on Progress in Physics 82, 104401 (2019).
  • Guan et al. (2019) Q. Guan, V. Klinkhamer, R. Klemt, J. H. Becher, A. Bergschneider, P. M. Preiss, S. Jochim,  and D. Blume, Physical Review Letters 122, 083401 (2019).
  • Fano (1935) U. Fano, Il Nuovo Cimento 12, 154 (1935).
  • Feshbach (1958) H. Feshbach, Annals of Physics 5, 357 (1958).
  • Tiesinga et al. (1993) E. Tiesinga, B. J. Verhaar,  and H. T. C. Stoof, Physical Review A 47, 4114 (1993).
  • Chin et al. (2010) C. Chin, R. Grimm, P. Julienne,  and E. Tiesinga, Reviews of Modern Physics 82, 1225 (2010).
  • Cetina et al. (2016) M. Cetina, M. Jag, R. S. Lous, I. Fritsche, J. T. Walraven, R. Grimm, J. Levinsen, M. M. Parish, R. Schmidt, M. Knap, et al., Science 354, 96 (2016).
  • 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).
  • Werner and Castin (2006a) F. Werner and Y. Castin, Physical Review Letters 97, 150401 (2006a).
  • Werner and Castin (2006b) F. Werner and Y. Castin, Physical Review A 74, 053604 (2006b).
  • Kerin and Martin (2020) A. D. Kerin and A. M. Martin, Physical Review A 102, 023311 (2020).
  • Kerin and Martin (2022b) A. Kerin and A. Martin, arXiv preprint arXiv:2207.09091v2  (2022b).
  • Nielsen et al. (2001) E. Nielsen, D. V. Fedorov, A. S. Jensen,  and E. Garrido, Physics Reports 347, 373 (2001).
  • Fedorov and Jensen (1993) D. Fedorov and A. Jensen, Physical review letters 71, 4103 (1993).
  • Fedorov and Jensen (2001) D. V. Fedorov and A. Jensen, Journal of Physics A: Mathematical and General 34, 6003 (2001).
  • Braaten and Hammer (2006) E. Braaten and H.-W. Hammer, Physics Reports 428, 259 (2006).
  • Thøgersen (2009) M. Thøgersen, arXiv preprint arXiv:0908.0852  (2009).
  • Srivastava et al. (2003) H. M. Srivastava, H. A. Mavromatis,  and R. S. Alassar, Applied Mathematics Letters 16, 1131 (2003).
  • Gradshteyn and Ryzhik (2014) I. S. Gradshteyn and I. M. Ryzhik, Table of integrals, series, and products (Academic press, 2014).