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

Quench dynamics of mass-imbalanced three-body fermionic systems in a spherical trap

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

We consider a system of two identical fermions of general mass interacting with a third distinguishable particle via a contact interaction within an isotropic three-dimensional harmonic trap. We calculate time-dependent observables of the system after it is quenched in s-wave scattering length. To do this we use exact closed form mass-imbalanced hyperspherical solutions to the static three-body problem. These exact solutions enable us to calculate two time-dependent observables, the Ramsey signal and particle separation, after the system undergoes a quench from non-interacting to the unitary regime or vice-versa.

I Introduction

Investigating the dynamics of non-equilibrium quantum systems is relevant to many areas in condensed matter physics. Areas of interest include the study of superfluid turbulence, topological states, mesoscopic circuits and can even be extended to neutron star dynamics. In these systems there are many interacting entities, making exact calculations problematic. However, it is possible to theoretically and experimentally examine quantum dynamics when there are only a few quantum bodies in the system of interest. An idealised example of such a system, where few-body quantum dynamics can be studied, is harmonically trapped quantum gases, which can be constructed at the single to few atom limit Serwane et al. (2011); Murmann et al. (2015); Zürn et al. (2013, 2012); Stöferle et al. (2006). In this work we focus on the interaction quench dynamics of three fermionic atoms trapped in a spherically symmetric harmonic trap.

We consider a system of two identical fermions which interact via a contact interaction with a third distinguishable particle, which can have a different mass. We consider the scenario where the contact interaction, between the third particle and the two identical fermions, is quenched either from the non-interacting regime to the unitary regime or vice-versa. To do this we utilise the exact solutions for the system Cui (2012); D’Incao et al. (2018); Jonsell et al. (2002); Blume and Daily (2010); Kerin and Martin (2022); Kestner and Duan (2007). These exact solutions have previously been used to elucidate the thermodynamic properties of quantum gases 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); Colussi et al. (2018, 2019); Enss et al. (2022). In the context of quench dynamics the results presented in this paper complement previous studies in two-dimensional Bougas et al. (2022) and one-dimensional systems Pecak et al. (2016); Volosniev (2017); Kehrberger et al. (2018); Sowiński and García-March (2019).

The paper is structured as follows. In Sec. II we briefly review the stationary three-body problem of two identical fermions interacting with a third distinguishable particle in a spherically symmetric harmonic trap. In Sec. III we use the eigenstates of the system to investigate the quench dynamics. In particular we focus on quenches from the non-interacting regime to the strongly interacting (unitary) regime, a forwards quench, and vice-versa, a backwards or reverse quench. For these two quenches we evaluate the Ramsey signal, i.e. the overlap of the time evolving state with the initial state, and post-quench evolution of the particle separation, as defined by the hyperradius. For the Ramsey signal we find that it can be calculated semi-analytically for any initial state for both quenches whilst for the particle separation we find that it can also be calculated semi-analytically for any initial state for the forwards quench but the reverse quench leads to non-physical divergences, as is the case for two-body quench dynamics Kerin and Martin (2020).

II Overview of the Three-Body Problem

Our starting point is the Hamiltonian of three non-interacting bodies in a three-dimensional spherical harmonic trap:

H^=k=13[22mkk2+mkω2rk22],\displaystyle\hat{H}=\sum_{k=1}^{3}\left[\frac{-\hbar^{2}}{2m_{k}}\nabla_{k}^{2}+\frac{m_{k}\omega^{2}r_{k}^{2}}{2}\right], (1)

where rk\vec{r}_{k} is the position of the kthk^{\rm th} particle, mkm_{k} is its mass, and ω\omega is the trapping frequency.

In this paper we consider the case of two identical fermions interacting with a distinct third particle. We define particle one to be the impurity (m1=mi)(m_{1}=m_{\rm i}) and particles two and three to be identical (m=m2=m3)(m=m_{2}=m_{3}). For convenience we define the reduced mass μ=mim/(mi+m)\mu=m_{\rm i}m/(m_{\rm i}+m), the lengthscale, aμ=/μωa_{\mu}=\sqrt{\hbar/\mu\omega}, and the mass imbalance κ=m/mi\kappa=m/m_{\rm i}.

For such a system the hyperspherical formulation Werner and Castin (2006a) gives a closed form solution for the wavefunction. However, because the interactions are enforced with the Bethe-Peierls boundary condition, the wavefunction can only be fully specified in the non-interacting and strongly interacting (unitary) 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)}, (2)

where

r\displaystyle\vec{r} =\displaystyle= r2r1,\displaystyle\vec{r}_{2}-\vec{r}_{1}, (3)
ρ\displaystyle\vec{\rho} =\displaystyle= 1γ(r3mir1+mr2mi+m),\displaystyle\frac{1}{\gamma}\left(\vec{r}_{3}-\frac{m_{\rm i}\vec{r}_{1}+m\vec{r}_{2}}{m_{\rm i}+m}\right), (4)
γ\displaystyle\gamma =\displaystyle= mi(mi+2m)m+mi,\displaystyle\frac{\sqrt{m_{\rm i}(m_{\rm i}+2m)}}{m+m_{\rm i}}, (5)

and the centre-of-mass (COM) coordinate is

C\displaystyle\vec{C} =\displaystyle= mir1+mr2+mr3mi+2m.\displaystyle\frac{m_{\rm i}\vec{r}_{1}+m\vec{r}_{2}+m\vec{r}_{3}}{m_{\rm i}+2m}. (6)

The centre-of-mass Hamiltonian is a simple harmonic oscillator (SHO) Hamiltonian of a single particle of mass M=2m+miM=2m+m_{\rm i} and position C\vec{C}. As such the centre-of-mass wavefunction is a SHO wavefunction. 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)\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}{$\scriptscriptstyle\bullet$}}}}}\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},

where Λ^r2\hat{\Lambda}_{r}^{2} and Λ^ρ2\hat{\Lambda}_{\rho}^{2} are the angular momentum operators in the r^\hat{r} and ρ^\hat{\rho} coordinate systems.

We define the trial wavefunction Werner (2008)

ψ3brel\displaystyle\psi_{\rm 3b}^{\rm rel} =\displaystyle= NqlsFqs(R)R2(1P^23)φls(α)sin(2α)Ylm(ρ^),\displaystyle N_{qls}\frac{F_{qs}(R)}{R^{2}}(1-\hat{P}_{23})\frac{\varphi_{ls}(\alpha)}{\sin(2\alpha)}Y_{lm}(\hat{\rho}), (8)

where NqlsN_{qls} is the normalisation constant, Fqs(R)F_{qs}(R) is the hyperradial wavefunction, ϕls(α)=(1P^23)φls(α)Ylm(ρ^)/sin(2α)\phi_{ls}(\alpha)=(1-\hat{P}_{23})\varphi_{ls}(\alpha)Y_{lm}(\hat{\rho})/\sin(2\alpha) is the hyperangular wavefunction, and P^23\hat{P}_{23} is the particle exchange operator which swaps the positions of particles two and three.

Three conditions determine the functional forms of Fqs(R)F_{qs}(R) and φls(α)\varphi_{ls}(\alpha),

φls(π2)\displaystyle\varphi_{ls}\left(\frac{\pi}{2}\right) =\displaystyle= 0,\displaystyle 0, (9)
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)
Erel\displaystyle E_{\rm rel} =\displaystyle= 24μ(F′′(R)+F(R)R)\displaystyle\frac{-\hbar^{2}}{4\mu}\left(F^{\prime\prime}(R)+\frac{F^{\prime}(R)}{R}\right) (11)
+\displaystyle+ (2s24μR2+μω2R2)F(R).\displaystyle\left(\frac{\hbar^{2}s^{2}}{4\mu R^{2}}+\mu\omega^{2}R^{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 the Schrödinger equation. l0l\in\mathbb{Z}_{\geq 0} is the angular momentum quantum number, q0q\in\mathbb{Z}_{\geq 0} and s>0s\in\mathbb{R}_{>0} are the energy eigenvalues and the energy is given Erel=(2q+l+s+1)ωE_{\rm rel}=(2q+l+s+1)\hbar\omega. Fqs(R)F_{qs}(R) and φls(α)\varphi_{ls}(\alpha) are given Werner and Castin (2006a, b); Liu et al. (2010)

Fqs(R)=R~seR~2/2Lqs(R~2),\displaystyle F_{qs}(R)=\tilde{R}^{s}e^{-\tilde{R}^{2}/2}L_{q}^{s}\left(\tilde{R}^{2}\right), (12)
φls(α)=cosl+1(α)\displaystyle\varphi_{ls}(\alpha)=\cos^{l+1}(\alpha)
×F12(l+1s2,l+1+s2;l+32;cos2(α)),\displaystyle\qquad\qquad\times{}_{2}F_{1}\left(\frac{l+1-s}{2},\frac{l+1+s}{2};l+\frac{3}{2};\cos^{2}(\alpha)\right),
(13)

where R~=R/aμ\tilde{R}=R/a_{\mu}, LqsL_{q}^{s} is a Laguerre polynomial, and F12{}_{2}F_{1} is the Gaussian hypergeometric function.

The contact interactions are enforced by the Bethe-Peierls condition Bethe and Peierls (1935)

limrij0[d(rijΨ)drij1rijΨ]=1as,\displaystyle\lim_{r_{ij}\rightarrow 0}\left[\frac{d(r_{ij}\Psi)}{dr_{ij}}\frac{1}{r_{ij}\Psi}\right]=\frac{-1}{a_{\rm s}}, (14)

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.

In the non-interacting limit (as0a_{\rm s}\rightarrow 0) Eq. (14) implies, for all values of κ=m/mi\kappa=m/m_{\rm i},

s={2n+4l=02n+l+2l>0,\displaystyle s=\begin{cases}2n+4&l=0\\ 2n+l+2&l>0\end{cases}, (15)

where n0n\in\mathbb{Z}_{\geq 0}. In the unitary limit (asa_{\rm s}\rightarrow\infty) the Bethe-Peierls boundary condition gives the transcendental equation

0=\displaystyle 0=
dφlsdα|α=0(1)l(1+κ)2κ1+2κφls(arctan(1+2κκ)).\displaystyle\frac{d\varphi_{ls}}{d\alpha}\bigg{|}_{\alpha=0}-(-1)^{l}\frac{(1+\kappa)^{2}}{\kappa\sqrt{1+2\kappa}}\varphi_{ls}\left(\arctan(\frac{\sqrt{1+2\kappa}}{\kappa})\right).
(16)

The values of ss at unitarity for l=0l=0 and a variety of κ\kappa are given in Table 1.

l=0l=0 κ=0.1\kappa=0.1 κ=1\kappa=1 κ=10\kappa=10
nn snls_{nl}
0 2.004… 2.166… 3.316…
1 5.817… 5.127… 4.707…
2 6.195… 7.114… 6.747…
3 9.685… 8.832… 8.876…
Table 1: Some three-body ss-eigenvalues at unitarity with l=0l=0 and κ=0.1,1\kappa=0.1,1, and 1010, (heavy impurity, equal mass and light impurity)

to three decimal places.

From this it is possible, in the non-interacting and unitary regimes, to evaluate the eigenenergies and eigenstates of the three-particle system. From this foundation in the following section we utilise these states to evaluate the quench dynamics of this system.

III Quench Dynamics

Below we investigate the behaviour of the system after a quench in the s-wave scattering length, asa_{\rm s}. We are interested in the forwards quench (non-interacting to unitary) and the backwards quench (unitary to non-interacting). Specifically, we calculate the Ramsey signal, S(t)S(t), and the particle separation, R~(t)\langle\tilde{R}(t)\rangle. In order to do this we need to calculate various integrals involving the wavefunction. First we have the Jacobian

dV=dr1dr2dr3=R54sin2(2α)γ3dRdαdΩrdΩρdC.\displaystyle dV=d\vec{r}_{1}d\vec{r}_{2}d\vec{r}_{3}=\frac{R^{5}}{4}\sin^{2}(2\alpha)\gamma^{3}dRd\alpha d\vec{\Omega}_{r}d\vec{\Omega}_{\rho}d\vec{C}.\quad (17)

We make the definitions

Fqs(R)|Fqs(R)=0RFqs(R)Fqs(R)𝑑R,\displaystyle\bra{F_{qs}(R)}\ket{F_{qs}(R)}=\int_{0}^{\infty}RF_{qs}(R)^{*}F_{qs}(R)dR, (18)
ϕls(α)|ϕls(α)=\displaystyle\bra{\phi_{ls}(\alpha)}\ket{\phi_{ls}(\alpha)}=
0π/2ϕls(α)ϕls(α)2sin2(2α)𝑑α𝑑Ωr𝑑Ωρ.\displaystyle\qquad\int\int\int_{0}^{\pi/2}\phi_{ls}(\alpha)^{*}\phi_{ls}(\alpha)2\sin^{2}(2\alpha)d\alpha d\vec{\Omega}_{r}d\vec{\Omega}_{\rho}.\quad (19)

In this work we wish to calculate time-varying observables of a post-quench system. To do this we need the time-dependent post-quench wavefunction. The COM wavefunction is independent of asa_{\rm s} and so is unaffected by the quench. As such we only need the time-dependent post-quench relative wavefunction, it is given

|ψ3brel(t)\displaystyle\ket{\psi_{\rm 3b}^{\rm rel}(t)} =\displaystyle= eiH^relt/|Fqisiϕlisi,\displaystyle e^{-i\hat{H}_{\rm rel}t/\hbar}\ket{F_{q_{\rm i}s_{\rm i}}\phi_{l_{\rm i}s_{\rm i}}},
=\displaystyle= q,sFqsϕlis|FqisiϕlisieiEqlist/|Fqsϕlis,\displaystyle\sum_{q,s}\bra{F_{qs}\phi_{l_{\rm i}s}}\ket{F_{q_{\rm i}s_{\rm i}}\phi_{l_{\rm i}s_{\rm i}}}e^{-iE_{ql_{\rm i}s}t/\hbar}\ket{F_{qs}\phi_{l_{\rm i}s}},

where quantum numbers with an i subscript are the initial quantum numbers and the summation is over all eigenstates of the post-quench system.

III.1 Ramsey signal

The Ramsey signal is defined as the wavefunction overlap of the pre- and post-quench wavefunctions, it is given

S(t)\displaystyle S(t) =\displaystyle= Ψi(t)|Ψ(t),\displaystyle\bra{\Psi_{\rm i}(t)}\ket{\Psi^{\prime}(t)},
S(t)\displaystyle S(t) =\displaystyle= j|Ψi(0)|Ψj|2ei(EjEi)t/,\displaystyle\sum_{j}|\bra{\Psi_{\rm i}(0)}\ket{\Psi^{\prime}_{j}}|^{2}e^{-i(E_{j}-E_{\rm i})t/\hbar}, (21)

where Ψi\Psi_{\rm i} is the pre-quench wavefunction with energy EiE_{\rm i} and Ψ\Psi^{\prime} is the post-quench wavefunction. To obtain Eq. (21) we have inserted a complete set of post-quench eigenstates, Ψj\Psi^{\prime}_{j}, with energies EjE_{j}, where the sum over jj is a sum over all post-quench eigenstates Kerin and Martin (2020).

Since the COM wavefunction is unaffected by the quench the Ramsey signal is given

S(t)\displaystyle S(t) =\displaystyle= q,s|Fqisiϕlisi|Fqsϕlis|2ei(EqlisEqilisi)t/,\displaystyle\sum_{q,s}|\bra{F_{q_{\rm i}s_{\rm i}}\phi_{l_{\rm i}s_{\rm i}}}\ket{F_{qs}\phi_{l_{\rm i}s}}|^{2}e^{-i(E_{ql_{\rm i}s}-E_{q_{\rm i}l_{\rm i}s_{\rm i}})t/\hbar},\quad (22)

where indices with subscript i are the eigenvalues of the initial state and the unlabelled indices correspond to the post-quench eigenvalues. Note that ϕls|ϕlls=0\bra{\phi_{ls}}\ket{\phi_{l^{\prime}\neq ls^{\prime}}}=0, i.e. hyperangular states of different angular momenta are orthogonal.

To evaluate the Ramsey signal we need to evaluate the hyperradial integral Fqs|Fqs\bra{F_{qs}}\ket{F_{q^{\prime}s^{\prime}}} and the hyperangular integral ϕls|ϕls\bra{\phi_{ls}}\ket{\phi_{ls^{\prime}}}.

The hyperradial integral is given Srivastava et al. (2003)

Fqs(R)|Fqs(R)=\displaystyle\bra{F_{qs}(R)}\ket{F_{q^{\prime}s^{\prime}}(R)}=
aμ22(q+sq)(q+ss21q)Γ(s+s+22)F23(q,s+s+22,ss+22;s+1,ss+22q;1).\displaystyle\qquad\frac{a_{\mu}^{2}}{2}\binom{q+s^{*}}{q}\binom{q^{\prime}+\frac{s^{\prime}-s^{*}}{2}-1}{q^{\prime}}\Gamma\left(\frac{s^{*}+s^{\prime}+2}{2}\right){}_{3}F_{2}\bigg{(}-q,\frac{s^{*}+s^{\prime}+2}{2},\frac{s^{*}-s^{\prime}+2}{2};s^{*}+1,\frac{s^{*}-s^{\prime}+2}{2}-q^{\prime};1\bigg{)}.\qquad (23)

For s=ss=s^{\prime} and qqq\neq q^{\prime} the hyperradial integral vanishes. Evaluating the hyperangular integral is not as straightforward as for the hyperradial integral due to the permutation operator. The hyperangle α\alpha is defined in terms of the Jacobi vectors r\vec{r} and ρ\vec{\rho}, which are defined in terms of r1,r2\vec{r}_{1},\vec{r}_{2}, and r3\vec{r}_{3}. We have defined r\vec{r} as being between particles one and two but it is possible to define r\vec{r} between any pair of particles. There are then three possible ways to define the Jacobi vectors for the three-body system, these Jacobi sets are related to one another by a “kinematic rotation” Nielsen et al. (2001), a coordinate transform in other words. We perform the hyperangular integral by taking advantage of these relations and “rotating” the terms acted upon by P^23\hat{P}_{23} into the same Jacobi set as the term not acted upon by P^23\hat{P}_{23} Fedorov and Jensen (1993, 2001); Braaten and Hammer (2006); Thøgersen (2009). However this restricts us to the l=0l=0 case due to the presence of the spherical harmonic term making the coordinate transform intractable for l>0l>0 as the hyperangular part of the wavefunction becomes a function of ρ^\hat{\rho} in addition to α\alpha. For the general mass case the overlap of ϕ0s\phi_{0s} and ϕ0s\phi_{0s^{\prime}} is given,

ϕ0s(α)|ϕ0s(α)=8π0π/2((1P^23)φ0s(α)sin(2α))((1P^23)φ0s(α)sin(2α))sin2(2α)𝑑α,\displaystyle\bra{\phi_{0s}(\alpha)}\ket{\phi_{0s^{\prime}}(\alpha)}=8\pi\int_{0}^{\pi/2}\left((1-\hat{P}_{23})\frac{\varphi_{0s}(\alpha)}{\sin(2\alpha)}\right)^{*}\left((1-\hat{P}_{23})\frac{\varphi_{0s^{\prime}}(\alpha)}{\sin(2\alpha)}\right)\sin^{2}(2\alpha)d\alpha, (24)
=16π[0π/2φ0s(α)φ0s(α)𝑑α(1+κ)22κ1+2κ0π/2φ0s(α)[|arctan(1κ(2+1κ))α|π/2|π/2arctan(1κ(2+1κ))α|φ0s(α)𝑑α]𝑑α].\displaystyle=16\pi\Bigg{[}\int_{0}^{\pi/2}\varphi_{0s}^{*}(\alpha)\varphi_{0s^{\prime}}(\alpha)d\alpha-\frac{(1+\kappa)^{2}}{2\kappa\sqrt{1+2\kappa}}\int_{0}^{\pi/2}\varphi_{0s}^{*}(\alpha)\Bigg{[}\int_{|\arctan(\sqrt{\frac{1}{\kappa}(2+\frac{1}{\kappa})})-\alpha|}^{\pi/2-|\pi/2-\arctan(\sqrt{\frac{1}{\kappa}(2+\frac{1}{\kappa})})-\alpha|}\varphi_{0s^{\prime}}(\alpha^{\prime})d\alpha^{\prime}\Bigg{]}d\alpha\Bigg{]}.\>\> (25)

For l=0l=0 Eq. (13) reduces to Werner (2008); Fedorov and Jensen (2001)

φ0s(α)sin(s(π2α)).\displaystyle\varphi_{0s}(\alpha)\propto\sin(s\left(\frac{\pi}{2}-\alpha\right)).\quad (26)

Note that combining Eqs. (25) and (26) for κ=1\kappa=1 (the equal mass case) does not give the same result as in Refs. Colussi (2019); Werner (2008) and only agrees when s=ss=s^{\prime} is one of the solutions to Eq. (16) for κ=1\kappa=1. This is because the latter references appropriately substitute Eq. (16) for κ=1\kappa=1 into the result for the hyperangular integral to simplify the expression.

However, for a general value of κ\kappa we have to be more careful. For a specific κ\kappa the ss eigenvalues produced by Eq. (16) are orthogonal to one another when Eq. (25) is evaluated only for that specific value of κ\kappa, i.e. the ss-eigenspectrum for κ=x\kappa=x produces an orthogonal set of states only when Eq. (25) is evaluated using κ=x\kappa=x. However, the non-interacting values of ss (s=4,6,8,s=4,6,8,\dots) are orthogonal to one another whatever the value of κ\kappa.

Now that we can calculate the wavefunction overlaps we can evaluate the Ramsey signal for the forwards and backwards quenches for any initial state with l=0l=0 and any mass imbalance. In Figs. 1 and 2 we plot the forwards and backwards quenches respectively for a variety of κ\kappa and initial states.

Before we discuss the properties of the Ramsey signals, in Figs. 1 and 2, it is worth exploring what we may expect in a general sense. Assuming in Eq. (22) that the sum is dominated by a few terms, the general form for S(t)S(t) can be represented by S(t)Aeait+Bebit+CecitS(t)\approx Ae^{-ait}+Be^{-bit}+Ce^{-cit}. The magnitude of the signal oscillates with angular frequencies (ab),(bc),(ac)(a-b),(b-c),(a-c) with the more heavily weighted terms being more significant. However the phase of the signal is dominated by the angular frequencies of the most heavily weighted terms not the difference between angular frequencies of different terms. In our case the angular frequencies (a,b,c,a,b,c,\dots) are the differences between a post-quench eigenenergy and the initial energy, and the difference between the angular frequencies ((ab),(bc),(ac),(a-b),(b-c),(a-c),\dots) is the difference between different post-quench eigenenergies.

Refer to caption
Refer to caption
Refer to caption
Figure 1: Ramsey signal for the forwards quench for a variety initial states and mass imbalances. In all panels the dot-dashed red line corresponds to (qi,si)=(0,4)(q_{\rm i},s_{\rm i})=(0,4), the dashed green line corresponds to (qi,si)=(1,4)(q_{\rm i},s_{\rm i})=(1,4), and the solid blue line corresponds to (qi,si)=(2,4)(q_{\rm i},s_{\rm i})=(2,4). The upper, middle, and lower plots correspond to κ=0.1,1\kappa=0.1,1, and 1010 (heavy impurity, equal mass and light impurity) respectively. Each Ramsey signal is obtained by evaluating Eq. (22) with 10 terms in each sum, i.e. 100 terms total. In this limit we find that the sum is convergent.

In Fig. 1 we plot the Ramsey signals of the forwards quench for a variety of initial states and mass imbalances. We find an irregularly repeating Ramsey signal, this is in contrast to similar calculations performed for the two-body problem Kerin and Martin (2020) where the signal is periodic with period π/ω\pi/\omega. This irregular periodicity arises from the unitary ss-eigenspectrum. In this case angular frequencies of each term in Eq. (22) are irrational and so are the differences between them, in general. These irrational angular frequencies lead to the irregular phase and magnitude.

For the κ=0.1\kappa=0.1 (heavy impurity) forwards quench the most heavily weighted terms in Eq. (22) for the (qi,si)=(0,4)(q_{\rm i},s_{\rm i})=(0,4) initial state (red line in the upper panel of Fig. 1 are the |F0,4ϕ0,4|F0,2.004ϕ0,2.004|20.539|\bra{F_{0,4}\phi_{0,4}}\ket{F_{0,2.004}\phi_{0,2.004}}|^{2}\approx 0.539, |F0,4ϕ0,4|F0,5.817ϕ0,5.817|20.134|\bra{F_{0,4}\phi_{0,4}}\ket{F_{0,5.817}\phi_{0,5.817}}|^{2}\approx 0.134 and |F0,4ϕ0,4|F1,2.004ϕ0,2.004|20.179|\bra{F_{0,4}\phi_{0,4}}\ket{F_{1,2.004}\phi_{0,2.004}}|^{2}\approx 0.179 terms. The corresponding angular frequencies are 2ω\approx-2\omega, 1.8ω\approx 1.8\omega, 0.004ω\approx 0.004\omega respectively. Hence the phase displays features with a period of π/ω\approx\pi/\omega, and the magnitude has three main modes with periods 0.5π/ω\approx 0.5\pi/\omega, π/ω\pi/\omega and 1.1π/ω\approx 1.1\pi/\omega.

For the κ=1\kappa=1 (equal mass) forwards quench the most significant terms for the (qi,si)=(0,4)(q_{\rm i},s_{\rm i})=(0,4) initial state are |F0,4ϕ0,4|F0,2.166ϕ0,2.166|20.499|\bra{F_{0,4}\phi_{0,4}}\ket{F_{0,2.166}\phi_{0,2.166}}|^{2}\approx 0.499, |F0,4ϕ0,4|F0,5.127ϕ0,5.127|20.297|\bra{F_{0,4}\phi_{0,4}}\ket{F_{0,5.127}\phi_{0,5.127}}|^{2}\approx 0.297, and |F0,4ϕ0,4|F1,2.166ϕ1,2.166|20.13|\bra{F_{0,4}\phi_{0,4}}\ket{F_{1,2.166}\phi_{1,2.166}}|^{2}\approx 0.13. The corresponding angular frequencies are 1.8ω\approx-1.8\omega, 1.1ω\approx 1.1\omega and 0.17ω\approx 0.17\omega respectively. The phase is then dominated by a mode with period 1.1π/ω\approx 1.1\pi/\omega and the magnitude is dominated by modes with periods 2π/3ω,π/ω\approx 2\pi/3\omega,\pi/\omega and 2π/ω\approx 2\pi/\omega.

For the κ=10\kappa=10 (light impurity) forwards quench the most significant terms for the (qi,si)=(0,4)(q_{\rm i},s_{\rm i})=(0,4) initial state are |F0,4ϕ0,4|F0,3.3169ϕ0,3.3169|20.607|\bra{F_{0,4}\phi_{0,4}}\ket{F_{0,3.3169}\phi_{0,3.3169}}|^{2}\approx 0.607, and |F0,4ϕ0,4|F0,4.707ϕ0,4.707|20.350|\bra{F_{0,4}\phi_{0,4}}\ket{F_{0,4.707}\phi_{0,4.707}}|^{2}\approx 0.350. The corresponding angular frequencies are 0.7ω/π\approx-0.7\omega/\pi and 0.7ω/π\approx 0.7\omega/\pi respectively hence the phase has an approximate periodicity of 2.8π/ω2.8\pi/\omega and the magnitude has a period of 1.4π/ω\approx 1.4\pi/\omega.

Refer to caption
Refer to caption
Refer to caption
Figure 2: Ramsey signal for the backwards quench for a variety of initial states and mass imbalances. In all panels the dot-dashed red line corresponds to qi=0q_{\rm i}=0, the dashed green line corresponds to qi=1q_{\rm i}=1, and the solid blue line corresponds to qi=2q_{\rm i}=2. The upper, middle and lower plots correspond to κ=0.1,1\kappa=0.1,1, and 1010 (heavy impurity, equal mass and light impurity) and si=2.004,2.166s_{\rm i}=2.004\dots,2.166\dots, and 3.3163.316\dots respectively. Each Ramsey signal is obtained by evaluating Eq. (22) with 10 terms in each sum, i.e. 100 terms total. In this limit we find that the sum is convergent.

In Fig. 2 we plot the Ramsey signals of the backwards quench for a variety of initial states and mass imbalances. We find a strongly periodic magnitude. In the non-interacting regime we have s=4,6,8,s=4,6,8,\dots for every value of κ\kappa. This means that every oscillatory term in the magnitude has an angular frequency that is a multiple of two, hence the period of π/ω\pi/\omega. However the behaviour of the phase is still dominated by the largest terms in the summation. For κ=0.1\kappa=0.1 (heavy impurity) and (qi=0,si=2.00449)(q_{\rm i}=0,s_{\rm i}=2.00449\dots) the largest term is |F0,2.004ϕ0,2.004|F0,4ϕ0,4|20.539|\bra{F_{0,2.004}\phi_{0,2.004}}\ket{F_{0,4}\phi_{0,4}}|^{2}\approx 0.539 and this defines the period of π/ω\approx\pi/\omega we see in the phase. Similarly for κ=1\kappa=1 (equal mass) and (qi=0,si=2.166)(q_{\rm i}=0,s_{\rm i}=2.166\dots) the largest term is |F0,2.166ϕ0,2.166|F0,4ϕ0,4|20.499|\bra{F_{0,2.166}\phi_{0,2.166}}\ket{F_{0,4}\phi_{0,4}}|^{2}\approx 0.499 defining a period of 1.1π/ω\approx 1.1\pi/\omega for the phase, and for κ=10\kappa=10 (light impurity) and (qi=0,si=3.3169)(q_{\rm i}=0,s_{\rm i}=3.3169\dots) the largest term is |F0,3.3169ϕ0,3.3169|F0,4ϕ0,4|20.607|\bra{F_{0,3.3169}\phi_{0,3.3169}}\ket{F_{0,4}\phi_{0,4}}|^{2}\approx 0.607 defining a period of 2.8π/ω\approx 2.8\pi/\omega for the phase.

III.2 Particle separation

The Ramsey signal is not the only quench observable we investigate. We can also calculate the expectation value of R~\tilde{R}, i.e. the particle separation.

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(t)}\tilde{R}\ket{\Psi(t)}
=j,jΨi(0)|ΨjΨj|Ψi(0)Ψj|R~|Ψjei(EjEj)t/,\displaystyle=\sum_{j,j^{\prime}}\bra{\Psi_{\rm i}(0)}\ket{\Psi^{\prime}_{j}}\bra{\Psi^{\prime}_{j^{\prime}}}\ket{\Psi_{\rm i}(0)}\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\Psi 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 sum over jj and jj^{\prime} taken over all post-quench eigenstates.

The COM wavefunction is independent of the interparticle interaction and does not change when the system is quenched 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,qsFqisiϕlisi|FqsϕlisFqsϕlis|Fqisiϕlisi\displaystyle\sum_{q^{\prime},q}\sum_{s}\bra{F_{q_{\rm i}s_{\rm i}}\phi_{l_{\rm i}s_{\rm i}}}\ket{F_{q^{\prime}s}\phi_{l_{\rm i}s}}\bra{F_{qs}\phi_{l_{\rm i}s}}\ket{F_{q_{\rm i}s_{\rm i}}\phi_{l_{\rm i}s_{\rm i}}} (28)
×Fqsϕlis|R~|Fqsϕlisei(EqlisEqlis)t/.\displaystyle\times\bra{F_{q^{\prime}s}\phi_{l_{\rm i}s}}\tilde{R}\ket{F_{qs}\phi_{l_{\rm i}s}}e^{-i(E_{ql_{\rm i}s}-E_{q^{\prime}l_{\rm i}s})t/\hbar}.

All integrals required for calculating Eq. (28) are calculated in Sec. III.1 except for Fqs|R~|Fqs\bra{F_{q^{\prime}s}}\tilde{R}\ket{F_{q^{\prime}s}}. This is given Srivastava et al. (2003)

Fqs|R~|Fqs=\displaystyle\bra{F_{qs}}\tilde{R}\ket{F_{q^{\prime}s^{\prime}}}=
aμ22(q+sq)(q+s2s232q)Γ(s+s+32)\displaystyle\frac{a_{\mu}^{2}}{2}\binom{q+s}{q}\binom{q^{\prime}+\frac{s^{\prime}}{2}-\frac{s}{2}-\frac{3}{2}}{q^{\prime}}\Gamma\left(\frac{s+s^{\prime}+3}{2}\right)
×F23(q,s+s+32,ss+32;s+1,ss+32q;1).\displaystyle\times{}_{3}F_{2}\left(-q,\frac{s^{\prime}+s+3}{2},\frac{s-s^{\prime}+3}{2};s+1,\frac{s-s^{\prime}+3}{2}-q^{\prime};1\right).

This result combined with the results of Sec. III.1 allows us to calculate R~(t)\langle\tilde{R}(t)\rangle for the forwards and backwards quench for any initial state with l=0l=0 and any mass imbalance.

Refer to caption
Refer to caption
Refer to caption
Figure 3: R~(t)\langle\tilde{R}(t)\rangle for the forwards quench from the non-interacting ground state, (qi,si)=(0,4)(q_{\rm i},s_{\rm i})=(0,4), for κ=0.1\kappa=0.1 (upper panel, heavy impurity), κ=1\kappa=1 (middle panel equal mass) and κ=10\kappa=10 (lower panel light impurity). 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 to Nmax=12N_{\rm max}=12 and the dotted black line to Nmax=24N_{\rm max}=24. As can be seen from the plots the sum is convergent with NmaxN_{\rm max}.

In Sec. III.1 we note that the magnitude and phase of the Ramsey signal for the forwards quench is irregular and this is because the ss eigenvalues are irrational at unitarity. Additionally the phase for the backwards quench is also irregular and this is due to the irrationality of sis_{\rm i}. However the angular frequencies of the terms in Eq. (28) are always even integers because the ss contributions to the energies cancel leaving angular frequencies proportional to (2q2q)ω(2q^{\prime}-2q)\omega and q,q0q,q^{\prime}\in\mathbb{Z}_{\geq 0}. This results in the angular frequency of every term in the summation being a multiple of 2ω2\omega causing R~(t)\langle\tilde{R}(t)\rangle to have a period of π/ω\pi/\omega in both the forwards and backwards quench.

In Fig. 3 we have plotted R~(t)\langle\tilde{R}(t)\rangle for the forwards quench for κ=0.1\kappa=0.1, κ=1\kappa=1 and κ=10\kappa=10 (heavy impurity, equal mass, and light impurity) in the upper, middle and lower panels respectively. For each κ\kappa we have taken the initial state to be the ground state and calculated the sum in Eq. (28) over qq, qq^{\prime} and ss up to Nmax=3,6,12N_{\rm max}=3,6,12 and 2424, where NmaxN_{\rm max} is the number of terms in each individual sum, so there are Nmax3N_{\rm max}^{3} terms total. In each case we find that the results for R~(t)\langle\tilde{R}(t)\rangle have converged at Nmax=24N_{\rm max}=24. As discussed above R~(t)\langle\tilde{R}(t)\rangle oscillates with a period π/ω\pi/\omega.

Additionally we observe that as κ\kappa increases (impurity becomes lighter) the amplitude of the oscillation decreases. In the κ\kappa\rightarrow\infty limit the unitary ss-eigenspectrum approaches s4,6,8s\rightarrow 4,6,8\dots Kerin and Martin (2022) and the initial non-interacting ground state overlaps perfectly with the unitary ground state. As such the amplitude decreases as κ\kappa increases until eventually R~\langle\tilde{R}\rangle reaches a constant value of 2.182.18\dots which is R~\langle\tilde{R}\rangle for q=0q=0, s=4s=4. This is also R~(t=0)\langle\tilde{R}(t=0)\rangle, the maximum particle separation does not change with κ\kappa but the minimum increases to decrease the amplitude. Conversely in the κ0\kappa\rightarrow 0 (heavy impurity) limit the unitary ss-eigenspectrum approaches s2,6,10,14s\rightarrow 2,6,10,14\dots, analytically this presents a problem. The non-interacting ground state, si=4s_{\rm i}=4, is orthogonal to the unitary ss-eigenspectrum because it is a subset of the non-interacting eigenspectrum, except s=2s=2 which is forbidden for l=0l=0 because it causes the hyperangular part of the wavefunction to be zero. Numeric investigations for κ\kappa as small as 10310^{-3} imply that the minimum R~(t)\langle\tilde{R}(t)\rangle asymptotes to 1.7\approx 1.7\dots as κ\kappa becomes very small. The maximum particle separation remains at 2.18aμ2.18\dots a_{\mu} whatever the value of κ\kappa because the initial non-interacting state does not depend on κ\kappa.

Refer to caption
Refer to caption
Refer to caption
Figure 4: R~(t)\langle\tilde{R}(t) for the backwards quench from the interacting ground state for a variety of κ\kappa. In all panels qi=0q_{\rm i}=0, in the upper panel (κ=0.1\kappa=0.1, heavy impurity) si=2.004s_{\rm i}=2.004\dots, in the middle panel (κ=1\kappa=1, equal mass) si=2.166s_{\rm i}=2.166\dots, and in the lower panel (κ=10\kappa=10, light impurity) si=3.3169s_{\rm i}=3.3169\dots. 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 to Nmax=12N_{\rm max}=12 and the dotted black line to Nmax=24N_{\rm max}=24. This sum diverges logarithmically with NmaxN_{\rm max} away from t=nπ/ωt=n\pi/\omega.

In Fig. 4 we have considered R~(t)\langle\tilde{R}(t)\rangle for the backwards quench for the κ=0.1\kappa=0.1, κ=1\kappa=1 and κ=10\kappa=10 cases, (heavy impurity, equal mass, and light impurity) in the upper, middle and lower panels respectively. For each κ\kappa we have taken the initial state to be the ground state and calculated the sum in Eq. (28) over qq, qq^{\prime} and ss up to Nmax=3,6,12N_{\rm max}=3,6,12 and 2424. Similar to r=|r1r2|r=|\vec{r}_{1}-\vec{r}_{2}| in the two body case Kerin and Martin (2020) we observe in the backwards quench that the particle separation diverges away from ωt=nπ\omega t=n\pi. For the three-body case we find that R~[t=(n+1/2)π/ω]\langle\tilde{R}[t=(n+1/2)\pi/\omega]\rangle diverges logarithmically with NmaxN_{\rm max}. More specifically we find that if the number of terms in the sums over qq and qq^{\prime} is fixed then as more terms are included in the sum over ss R~(t)\langle\tilde{R}(t)\rangle converges. The divergence in R~(t)\langle\tilde{R}(t)\rangle comes from the sums over qq and qq^{\prime} and therefore from the hyperradial wavefunction. This divergence warrants further scrutiny, to this end we investigate the evolution of the probability distribution of R(t)R(t),

P(R,t)=Ψ(t)|δ(RR)|Ψ(t).\displaystyle P(R^{\prime},t)=\bra{\Psi(t)}\delta(R-R^{\prime})\ket{\Psi(t)}. (30)
Refer to caption
Refer to caption
Refer to caption
Figure 5: The evolution of the hyperradial probability distribution, Eq. (30), following a forwards quench. Each plotted curve is P(R,t)P(R,t) for a specified value of tt and the upper, middle and lower panels correspond to κ=0.1,1\kappa=0.1,1 and 1010, the heavy impurity, equal mass and light impurity cases, respectively. The red dot-dashed line corresponds to t=0t=0, the dashed green line to t=0.17π/ωt=0.17\pi/\omega, the solid blue line to t=0.34π/ωt=0.34\pi/\omega and the dotted black line to t=0.5π/ωt=0.5\pi/\omega. For all plots the initial state is (qi,si)=(0,4)(q_{\rm i},s_{\rm i})=(0,4) and calculations are performed with Nmax=24N_{\rm max}=24. Each curve is convergent with NmaxN_{\rm max}.

In Fig. 5 we plot P(R,t)P(R,t) at t=0,0.17π/ω,0.34π/ωt=0,0.17\pi/\omega,0.34\pi/\omega, and π/ω\pi/\omega for κ=0.1,1\kappa=0.1,1, and 1010 (heavy impurity, equal mass, and light impurity respectively) as a function of RR for the forwards quench. We find that the peak of the probability distribution shifts inwards as the system evolves, reaching its innermost point at t=π/2ωt=\pi/2\omega then evolving in reverse back to its original configuration and continuing to oscillate in this way with period t=π/ωt=\pi/\omega. The magnitude of oscillation is smaller for larger κ\kappa. This oscillatory behaviour and its dependence on κ\kappa is unsurprising given the behaviour observed in Fig. 3. However the double peak, which is present for κ5\kappa\lesssim 5, is unusual. We can illuminate the double peak structure by looking at the overlaps of the pre- and post-quench wavefunctions. Looking at the κ=0.1\kappa=0.1 (heavy impurity) case the largest overlaps are |F0,4ϕ0,4|F0,2.004ϕ0,2.004|20.539|\bra{F_{0,4}\phi_{0,4}}\ket{F_{0,2.004}\phi_{0,2.004}}|^{2}\approx 0.539, |F0,4ϕ0,4|F0,5.817ϕ0,5.817|20.134|\bra{F_{0,4}\phi_{0,4}}\ket{F_{0,5.817}\phi_{0,5.817}}|^{2}\approx 0.134 and |F0,4ϕ0,4|F1,2.004ϕ0,2.004|20.179|\bra{F_{0,4}\phi_{0,4}}\ket{F_{1,2.004}\phi_{0,2.004}}|^{2}\approx 0.179. R~\langle\tilde{R}\rangle of these states projected onto are 1.66\approx 1.66, 2.56\approx 2.56, and 2.07\approx 2.07 respectively. Compare this to the κ=10\kappa=10 (light impurity) case where the most significant states are |F0,4ϕ0,4|F0,3.3169ϕ0,3.3169|20.607|\bra{F_{0,4}\phi_{0,4}}\ket{F_{0,3.3169}\phi_{0,3.3169}}|^{2}\approx 0.607, and |F0,4ϕ0,4|F0,4.707ϕ0,4.707|20.350|\bra{F_{0,4}\phi_{0,4}}\ket{F_{0,4.707}\phi_{0,4.707}}|^{2}\approx 0.350 and R~\langle\tilde{R}\rangle of the states projected onto are 2.01\approx 2.01 and 2.33\approx 2.33. For small κ\kappa the initial state most heavily overlaps with a few states with a relatively large variation in R~\langle\tilde{R}\rangle for large κ\kappa the overlaps are with states that are more closely clustered in R~\langle\tilde{R}\rangle hence P(R,t=π/2ω)P(R,t=\pi/2\omega) is singly peaked in the latter case. Physically speaking the oscillation amplitude grows smaller for a lighter impurity particle because the less mass (and therefore momenta) it has the less it is able to affect the positions of the two fermions. For the forwards quench we find P(R,t)P(R,t) is convergent as NmaxN_{\rm max}\rightarrow\infty at all points in time, as expected given R(t)\langle R(t)\rangle is convergent for the forwards quench. We now turn to the divergent case, the reverse quench.

Refer to caption
Refer to caption
Refer to caption
Figure 6: The evolution of the hyperradial probability distribution, Eq. (30), following a backwards quench. Each plotted curve is P(R,t)P(R,t) for a specified value of tt and the upper, middle and lower panels correspond to κ=0.1,1\kappa=0.1,1 and 1010, the heavy impurity, equal mass and light impurity cases, respectively. The red dot-dashed line corresponds to t=0t=0, the dashed green line to t=0.17π/ωt=0.17\pi/\omega, the solid blue line to t=0.34π/ωt=0.34\pi/\omega and the dotted black line to t=0.5π/ωt=0.5\pi/\omega. The initial state is given qi=0q_{\rm i}=0 for all plots, for the upper, middle and lower plots si=2.004,2.166s_{\rm i}=2.004\dots,2.166\dots and 3.31693.3169\dots respectively and all calculations are performed with Nmax=24N_{\rm max}=24. Only the t=0t=0 curve (red dot-dashed) is convergent with NmaxN_{\rm max}.
Refer to caption
Figure 7: The tail of P(R,t=π/2ω)P(R,t=\pi/2\omega) in the reverse quench with κ=1\kappa=1 (equal mass) and (qi,si)=(0,2.166)(q_{\rm i},s_{\rm i})=(0,2.166\dots) 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. 6 we plot P(R,t)P(R,t) at t=0,0.17π/ω,0.34π/ωt=0,0.17\pi/\omega,0.34\pi/\omega, and π/ω\pi/\omega for κ=0.1,1\kappa=0.1,1, and 1010 (heavy impurity, equal mass, and light impurity respectively) as a function of RR for the backwards quench. As with the forwards quench P(R,t)P(R,t) oscillates with period π/ω\pi/\omega however the peak initially moves outwards rather than inwards. The magnitude of the oscillations grows smaller for larger κ\kappa (lighter impurity) but the behaviour is qualitatively similar whatever the mass imbalance. For t=nπ/ωt=n\pi/\omega P(R,t)P(R,t) is approximately a Gaussian and converges with NmaxN_{\rm max}, however away from t=nπ/ωt=n\pi/\omega the probability distribution develops a long tail. The behaviour of the tail at t=π/2ωt=\pi/2\omega for κ=1\kappa=1 is plotted in Fig. 7 for various NmaxN_{\rm max} and the behaviour is qualitatively similar for different κ\kappa. The long tail decays approximately as 1/R~21/\tilde{R}^{2} before becoming exponentially suppressed, a “cut-off”. The larger NmaxN_{\rm max} is the later the cut-off occurs. While P(R,t=π/2ω)P(R,t=\pi/2\omega) is normalised as NmaxN_{\rm max}\rightarrow\infty this increasing long tail means that integral of RP(R,t=π/2ω)RP(R,t=\pi/2\omega) from R=0R=0 to RR\rightarrow\infty is divergent, hence R~\langle\tilde{R}\rangle diverges.

In the two-body case it has been suggested that the divergence in r=|r1r2|\langle r\rangle=\langle|\vec{r}_{1}-\vec{r}_{2}|\rangle is due to the 1/r1/r divergence in the two-body relative wavefunctionKerin and Martin (2020). In the three-body case the hyperradial part of Eq. (8) does not have a 1/R1/R divergence and has a cusp at R~0\tilde{R}\rightarrow 0 for the unitary κ=0.1\kappa=0.1 (heavy impurity) and κ=1\kappa=1 (equal mass) ground states but not for the κ=10\kappa=10 (light impurity) ground state. Nonetheless the divergence in R~(t)\langle\tilde{R}(t)\rangle is present in all three cases. More specifically we find that for si<3s_{\rm i}<3 the initial wavefunction exhibits a cusp, whilst for si3s_{\rm i}\geq 3 there is no cusp in the initial wavefunction. Regardless of which regime the initial state is in the logarithmic divergence of R~(t)\langle\tilde{R}(t)\rangle persists.

However, it is clear that the finite range of the interaction in a real system provides a natural cut-off in the sum in Eq. (28). A finite range of interaction defines a minimum de Broglie wavelength which in turn defines a maximum energy which defines a maximum number of terms in the sums of Eq. (28) and thus a maximum R~(t)\langle\tilde{R}(t)\rangle. For a system of three sodium atoms (i.e. κ=1\kappa=1) in a 1kHz trap and an interaction cut-off of 10910^{-9}m we obtain a cut off energy of Erel8.7×106ωE_{\rm rel}\approx 8.7\times 10^{6}\hbar\omega and thus expect a maximum R~(t)\langle\tilde{R}(t)\rangle of 11\approx 11. With this cut-off R~(t)\langle\tilde{R}(t)\rangle for the backwards quench oscillates between 1.5\approx 1.5 and 11\approx 11, an amplitude of 5\approx 5. This is significantly larger than in the forwards quench where R~(t)\langle\tilde{R}(t)\rangle oscillates between 1.8\approx 1.8 and 2.2\approx 2.2. In light of the divergence it is natural to consider the effect of using finite-range interaction models rather than zero-range as done here, and the case of two-bodies with a soft-core interaction has been solved analytically in a one-dimensional harmonic trap Kościk and Sowiński (2018). In the limit of small interaction range it is likely that the dynamics will be similar, but the effects of longer range interactions on the dynamics are unclear. If the source of the R~(t)\langle\tilde{R}(t)\rangle divergence is indeed the zero-range nature of the interaction then the finite-range model may not have the divergence present in the zero-range model.

However the zero-range interaction is not the only non-physical aspect of this model. We assume that the quench in asa_{\rm s} is instantaneous, in experiment asa_{\rm s} will change continuously over some finite time. The instantaneous quench we consider here may also be related to the R~(t)\langle\tilde{R}(t)\rangle divergence in the backwards quench, but it is difficult to calculate a non-instantaneous quench in this formalism. In the two-body formalism it is possible to quench between any two scattering lengths Kerin and Martin (2020) and thus numerically calculate a non-instantaneous quench. It would be interesting to see how this would affect the dynamics of the system and if this affects the divergence.

IV Conclusion

In this work we examine the time evolution of quenched systems. We consider a harmonically trapped system of two identical fermions plus a distinct particle interacting via a contact interaction where the system is quenched from non-interacting to strongly interacting or vice versa. We calculate the static wavefunction in both the non-interacting and strongly interacting regimes for general mass and use these solutions to calculate two time-dependent post-quench observables; the Ramsey signal, S(t)S(t), and the particle separation, R~(t)\langle\tilde{R}(t)\rangle. These observables are calculated for both the forwards and backwards quenches.

For the Ramsey signal we find an irregularly repeating signal for the forwards quench. This irregularity is due to the irrationality of the unitary energy spectrum and this irregular behaviour is more pronounced for small κ\kappa (heavy impurity). For the reverse quench the magnitude of the Ramsey signal oscillates with period π/ω\pi/\omega as the non-interacting energies are integer multiples and the phase has an irregular period due to the irrational initial energy.

For the particle separation the forwards quench yields the expected oscillating result, however the period is π/ω\pi/\omega as the irrational contributions to the unitary eigenenergies cancel. For the backwards quench we find that the particle separation diverges logarithmically, similar to divergence in r=|r2r1|r=|\vec{r}_{2}-\vec{r}_{1}| for the same quench performed on the two-body system Kerin and Martin (2020). By enforcing a cut-off based on the van der Waals interaction range we estimate a very large amplitude of oscillation, 5aμ\approx 5a_{\mu} for κ=1\kappa=1. From a physical perspective it is unclear why the divergence occurs for the backwards but not forwards quench but it is likely related to the zero-range interaction and/or the instantaneous quench.

Finally we note that experimental testing of these theoretical predictions is within reach. Few-atom systems can be reliably constructed with modern techniques Serwane et al. (2011); Murmann et al. (2015); Zürn et al. (2013, 2012); Stöferle et al. (2006), the quench in s-wave scattering length can be achieved using Feshbach resonance Fano (1935); Feshbach (1958); Tiesinga et al. (1993); Chin et al. (2010), and the Ramsey signal can be measured using Ramsey interferometry techniques Cetina et al. (2016). Notably Ref. Guan et al. (2019) measured the particle separation of two harmonically trapped 6Li atoms following a quench in trap geometry rather than s-wave scattering length. Additionally there have been theoretical advances that allow for the four-body wavefunction to be characterised analytically in the untrapped case for 3+1 and 2+2 fermi systems Endo and Castin (2015); Castin et al. (2010). These advances may allow for this work to be generalised to the four-body 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.

References

  • 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).
  • D’Incao et al. (2018) J. P. D’Incao, J. Wang,  and V. 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).
  • Kerin and Martin (2022) A. Kerin and A. Martin, arXiv preprint arXiv:2204.09205  (2022).
  • 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).
  • Colussi et al. (2018) V. Colussi, J. P. Corson,  and J. P. D’Incao, Physical review letters 120, 100401 (2018).
  • Colussi et al. (2019) V. Colussi, B. van Zwol, J. D’Incao,  and S. Kokkelmans, Physical Review A 99, 043604 (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).
  • Kerin and Martin (2020) A. D. Kerin and A. M. Martin, Physical Review A 102, 023311 (2020).
  • Werner and Castin (2006a) F. Werner and Y. Castin, Physical Review Letters 97, 150401 (2006a).
  • Werner (2008) F. Werner, Theses, Université Pierre et Marie Curieâ Paris VI, Paris France  (2008).
  • Werner and Castin (2006b) F. Werner and Y. Castin, Physical Review A 74, 053604 (2006b).
  • 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).
  • Srivastava et al. (2003) H. M. Srivastava, H. A. Mavromatis,  and R. S. Alassar, Applied Mathematics Letters 16, 1131 (2003).
  • 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).
  • Colussi (2019) V. E. Colussi, Atoms 7, 19 (2019).
  • Kościk and Sowiński (2018) P. Kościk and T. Sowiński, Scientific Reports 8, 1 (2018).
  • 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).
  • 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).
  • Endo and Castin (2015) S. Endo and Y. Castin, Physical Review A 92, 053624 (2015).
  • Castin et al. (2010) Y. Castin, C. Mora,  and L. Pricoupenko, Physical review letters 105, 223201 (2010).