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

Multiparticle Entanglement Dynamics of Quantum Chaos
in a Bose-Einstein condensate

Sheng-Chang Li1,2    Luca Pezzè2    Augusto Smerzi2 1. MOE Key Laboratory for Nonequilibrium Synthesis and Modulation of Condensed Matter, Shaanxi Province Key Laboratory of Quantum Information and Quantum Optoelectronic Devices, and School of Physics, Xi’an Jiaotong University, Xi’an 710049, China
2. QSTAR, INO-CNR and LENS, Largo Enrico Fermi 2, Firenze 50125, Italy
Résumé

We study the particle-entanglement dynamics witnessed by the quantum Fisher information (QFI) of a trapped Bose-Einstein condensate governed by the kicked rotor Hamiltonian. The dynamics is investigated with a beyond mean-field approach. We link the time scales of the validity of this approximation in, both, classical regular and chaotic regions, with the maximum Lyapunov exponents of the classical system. This establishes an effective connection between the classical chaos and the QFI. We finally study the critical point of a quantum phase transition using the beyond mean-field approximation by considering a two-mode bosonic Josephson junction with attractive interparticle interaction.

I Introduction

Quantum chaos aims to investigate the properties of quantum systems that have a chaotic dynamics in the classical limit. First works on quantum chaos can be traced back to 1950s [1] or even earlier [2]. While systematic theoretical studies emerged in the late 1980s [3], the experimental activity – requiring highly controllable many-body systems – has begun in the last years [4]. Atomic Bose-Einstein condensates (BECs) can be trapped in time-dependent potentials having almost arbitrary geometries and are therefore emerging as ideal candidates for the study of quantum chaos in many-body systems [5, 6, 7, 8]. Several efforts have been devoted to investigate chaotic dynamics in both scalar and spinor BECs, including driven two-mode [9], tilted triple-well [10], kicked ring [11], kicked optical-lattice [12], barrier-height-modulated double-well [13], dissipative double-well [14], dissipative optical-lattice [3], spin-1 [15], spin-2 [16], and coupled spinor atom-molecule BECs [17]. A paradigmatic model to study quantum chaos is the kicked rotor [18, 19] which can be realized with trapped BECs [20, 21, 22, 23, 24]. This model has been widely studied to illustrate classical and quantum chaos and the transition between them [25, 26].

An intriguing problem emerging in the recent literature is the study of the interplay between quantum chaos and entanglement. It has been indeed recognized that chaotic dynamics and entanglement generation are intimately related in many systems [29, 28, 27]. In particular, Ref. [29] presented an unifying picture connecting the multiparticle entanglement growth to the quantifiers of classical and quantum chaos. In this paper, we study the multiparticle entanglement dynamics of a BEC governed by the quantum kicked rotor (QKR) model. Multipartite entanglement is witnessed by studying the quantum Fisher information (QFI) [30, 31, 32, 33, 34, 35, 36, 37]. The QFI [38, 39] has proved a quite valuable quantity for the analysis of entanglement in complex systems [29, 28, 40, 47, 41, 42, 43, 44, 45, 46]. Here, we study the chaotic dynamics of the QFI under the framework of a beyond mean-field (BMF) approximation [48]. We show that the time scalings of the validity of the BMF dynamics are related with the maximum Lyapunov exponents of the corresponding classical chaotic regimes. This provides an effective relation between the time scales of quantum evolution in a chaotic system and the QFI. We further employ the BMF method to the analysis of the ground state of the system. We show that the QFI evaluated within the BMF method is in excellent agreement with exact results able to characterize the quantum phase transition in finite-size systems. The multipartite entanglement detected by the QFI is a resource for parameter estimation beyond the sensitivity limit achievable with separable (or classically-correlated) states [30, 33, 34, 35, 36, 37]. In this respect, our results show that the generation of metrologically useful entanglement is much faster in chaotic systems, when compared to entanglement generation in integrable systems. As shown in Ref. [49], quantum chaos can also enhance the sensitivity in the estimation of Hamiltonian parameters.

II Quantum Fisher information within a beyond mean-field approach

In the following, we study the dynamics of a two-mode model governed by the Hamiltonian (=1\hbar=1)

H^(t)=a(t)J^x+cNJ^z2,\hat{H}(t)=a(t)\hat{J}_{x}+\frac{c}{N}\hat{J}_{z}^{2}, (1)

where J^x=(a^1a^2+a^2a^1)/2\hat{J}_{x}=(\hat{a}_{1}^{\dagger}\hat{a}_{2}+\hat{a}_{2}^{\dagger}\hat{a}_{1})/2, J^y=(a^1a^2a^2a^1)/2i\hat{J}_{y}=(\hat{a}_{1}^{\dagger}\hat{a}_{2}-\hat{a}_{2}^{\dagger}\hat{a}_{1})/2{i}, and J^z=(a^1a^1a^2a^2)/2\hat{J}_{z}=(\hat{a}_{1}^{\dagger}\hat{a}_{1}-\hat{a}_{2}^{\dagger}\hat{a}_{2})/2 are the SU(2) generators satisfying angular momentum commutation relation [J^x,J^y]=iJ^z[\hat{J}_{x},\hat{J}_{y}]=i\hat{J}_{z} with cyclic permutation. Here, a^1,2\hat{a}_{1,2} (a^1,2\hat{a}_{1,2}^{\dagger}) are bosonic annihilation (creation) operators for the two modes (labelled as “1” and “2”). The time-dependent parameter a(t)a(t) describes the coupling between the two modes, cc depends on the two-body interaction, and NN is the total number of particles. In atomic BEC experiments, the two-mode model (1) can be realized by using either a double-well trapping potential [50, 51, 52, 53, 54, 56, 57, 55, 58, 59, 60, 61] or two hyperfine atomic states [62, 63, 64, 65]. As initial state, we consider a coherent spin state [66]

|ψs=\displaystyle|\psi_{s}\rangle= μ=N2N2(NN2+μ)(cosθ2)N2μ(sinθ2)N2+μ\displaystyle\sum_{\mu=-\frac{N}{2}}^{\frac{N}{2}}\sqrt{\bigg{(}\begin{array}[]{c}N\\ \frac{N}{2}+\mu\end{array}\bigg{)}}\bigg{(}\cos\frac{\theta}{2}\bigg{)}^{\frac{N}{2}-\mu}\bigg{(}\sin\frac{\theta}{2}\bigg{)}^{\frac{N}{2}+\mu} (5)
×ei(N2+μ)ϕ|μ,\displaystyle\times e^{i(\frac{N}{2}+\mu)\phi}|\mu\rangle,

where |μ|\mu\rangle are eigenstates of J^z\hat{J}_{z} and s=(sinθcosϕ;sinθsinϕ;cosθ)s=(\sin\theta\cos\phi;\sin\theta\sin\phi;\cos\theta) is the mean spin direction. We will focus on the dynamical evolution of the initial state sx=1s_{x}=1 (sy=sz=0s_{y}=s_{z}=0). This corresponds to the coherent spin state given with all NN spins pointing along the +x+x direction in the Bloch sphere. For the initial state, the collective spin mean values and variances are J^x=N/2\langle\hat{J}_{x}\rangle={N}/{2} and J^=0\langle\hat{J}_{\perp}\rangle=0, (ΔJ^x)2=0(\Delta\hat{J}_{x})^{2}=0 and (ΔJ^)2=N/4(\Delta\hat{J}_{\perp})^{2}={N}/{4}, respectively, where \perp indicates an arbitrary direction orthogonal to the mean spin direction.

We characterize the quantum dynamics by studying the QFI for unitary parametric transformations generated by an operator J^\hat{J} [39, 35, 37]. The QFI of a generic state ρ^\hat{\rho} is formally defined as the variance FQ[ρ^,J^n]=(ΔL^)2F_{Q}[\hat{\rho},\hat{J}_{n}]=(\Delta\hat{L})^{2} of the symmetric logarithmic derivative L^\hat{L}, satisfying 2[J^,ρ^]={L^,ρ^}2[\hat{J},\hat{\rho}]=\{\hat{L},\hat{\rho}\} [39, 35, 37]. In terms of the spectral decomposition of ρ^t\hat{\rho}_{t} at time tt (namely ρ^t=λpλλ||λ\hat{\rho}_{t}=\sum_{\lambda}p_{\lambda}\langle\lambda||\lambda\rangle, with pλ0p_{\lambda}\geq 0 and |λ|\lambda\rangle, eigenvalues and eigenvectors of ρ^\hat{\rho}, respectively), the QFI reads [39, 35, 37]

FQ(t)=2λ,λ(pλpλ)2pλ+pλ|λ|J^|λ|2.F_{Q}(t)=2\sum_{\lambda,\lambda^{\prime}}\frac{(p_{\lambda}-p_{\lambda^{\prime}})^{2}}{p_{\lambda}+p_{\lambda^{\prime}}}|\langle\lambda|\hat{J}|\lambda^{\prime}\rangle|^{2}. (6)

It has been shown [41] that Eq. (6) can be related to the dynamical susceptibility, see also [44, 46]. For a pure quantum state |ψt|\psi_{t}\rangle at time tt, Eq. (6) reduces to

FQ(t)=4(ΔJ^)2,F_{Q}(t)=4(\Delta\hat{J})^{2}, (7)

where (ΔJ^)2=ψt|J^2|ψtψt|J^|ψt2(\Delta\hat{J})^{2}=\langle\psi_{t}|\hat{J}^{2}|\psi_{t}\rangle-\langle\psi_{t}|\hat{J}|\psi_{t}\rangle^{2}. On the experimental side, the relevant quantity to detect is Eq. (6). The variance of the operator J^\hat{J} captures the behaviour of the QFI in the limit where the dynamics generating the quantum state is characterized by vanishing noise.

In the following we will restrict to the calculations of the QFI for pure states and for the operators J^=J^x,y,z\hat{J}=\hat{J}_{x,y,z} introduced above. In particular, the optimization of the QFI over J^x,y,z\hat{J}_{x,y,z} can be obtained as the maximum eigenvalue of the 3×33\times 3 covariance matrix

[𝚲Q]kl=2[(J^kJ^l+J^lJ^k)2J^kJ^l],[\mathbf{\Lambda}_{Q}]_{kl}=2[(\langle\hat{J}_{k}\hat{J}_{l}\rangle+\langle\hat{J}_{l}\hat{J}_{k}\rangle)-2\langle\hat{J}_{k}\rangle\langle\hat{J}_{l}\rangle], (8)

with k,l=x,y,zk,l=x,y,z, and the optimal direction is given by the corresponding eigenvector.

The QFI is related to multipartite entanglement [31, 32, 30, 33, 34, 35, 36, 37]: a QFI satisfying FQ(t)>𝓈k2+𝓇2F_{Q}(t)>\mathscr{s}k^{2}+\mathscr{r}^{2}, where 𝓈=N/k\mathscr{s}=\left\lfloor{N}/{k}\right\rfloor (being the maximum integer smaller than or equal to N/k{N}/{k}) and 𝓇=N𝓈k\mathscr{r}=N-\mathscr{s}k, is a witness of (k+1)(k+1)-partite entanglement among the NN bosonic particles in the state |ψt|\psi_{t}\rangle. Multipartite entanglement generated the dynamical evolution of an initially coherent spin state via Eq. (1) has been largely studied in the literature. Theoretical works [30, 67] have focused on the one-axis twisting (OAT) Hamiltonian, corresponding to a(t)=0a(t)=0 in Eq. (1[68, 69]. The OAT dynamics has been realized experimentally by exploiting interatomic interaction in BECs [65, 70, 71] or the atom-light interactions in cold thermal atoms [73, 72, 74, 75, 76], see Ref. [37] for a comprehensive review. Another model of interest is the twist-and-turn (TAT) Hamiltonian, corresponding to a(t)=A>0a(t)=A>0 in Eq. (1[51, 62, 63, 77]. The TAT dynamics is characterized by the competition between the twisting (due to the quadratic term in the Hamiltonian) and the linear coupling between the two modes that turns the state along an orthogonal direction. It has been shown, both experimentally [78, 79] and theoretically [47] that the TAT model creates entanglement faster than the OAT model. More recently, multipartite entanglement witnessed by the QFI has been studied for chaotic systems, where a(t)a(t) has a δ\delta-function time dependence [29].

Furthermore, as shown in Refs. [27, 80] the QFI can be also directly connected to the out-of-time-order correlations. Let us introduce the quantity t(ϕ)=tr[ρ^0ρ^t(ϕ)]=tr[W^t(ϕ)ρ^0W^t(ϕ)ρ^0]\mathscr{F}_{t}(\phi)={\rm tr}[\hat{\rho}_{0}\hat{\rho}_{t}(\phi)]={\rm tr}[\hat{W}_{t}^{\dagger}(\phi)\hat{\rho}_{0}^{\dagger}\hat{W}_{t}(\phi)\hat{\rho}_{0}], where ρ^0\hat{\rho}_{0} is the initial state, ρ^t(ϕ)=eiH^teiJ^ϕeiH^tρ^0eiH^teiJ^ϕeiH^t\hat{\rho}_{t}(\phi)=e^{i\hat{H}t}e^{-i\hat{J}\phi}e^{-i\hat{H}t}\hat{\rho}_{0}e^{i\hat{H}t}e^{i\hat{J}\phi}e^{-i\hat{H}t}, and W^t(ϕ)=eiH^teiJ^ϕeiH^t\hat{W}_{t}(\phi)=e^{i\hat{H}t}e^{-i\hat{J}\phi}e^{-i\hat{H}t} and W^t(ϕ)=eiH^teiJ^ϕeiH^t\hat{W}^{\dagger}_{t}(\phi)=e^{i\hat{H}t}e^{i\hat{J}\phi}e^{-i\hat{H}t}. Notice that, in the definition above we have considered a time-independent Hamiltonian H^\hat{H}, as relevant in our case (see discussion in Sec. III). Expanding in Taylor series for ϕ1\phi\ll 1, we have t(ϕ)=tr[ρ^t2]ϕ2tr[ρ^t2J^2(ρ^tJ^)2]+O(ϕ3)\mathscr{F}_{t}(\phi)={\rm tr}[\hat{\rho}_{t}^{2}]-\phi^{2}{\rm tr}[\hat{\rho}_{t}^{2}\hat{J}^{2}-(\hat{\rho}_{t}\hat{J})^{2}]+O(\phi^{3}), where ρ^t=eiH^tρ^0eiH^t\hat{\rho}_{t}=e^{i\hat{H}t}\hat{\rho}_{0}e^{-i\hat{H}t}. The second derivative with respect to ϕ\phi is [81, 80, 27]

2d2t(ϕ)dϕ2=4tr[ρ^t2J^2(ρ^tJ^)2]FQ(t),-2\frac{d^{2}\mathscr{F}_{t}(\phi)}{d\phi^{2}}=4{\rm tr}[\hat{\rho}_{t}^{2}\hat{J}^{2}-(\hat{\rho}_{t}\hat{J})^{2}]\leq F_{Q}(t), (9)

with equality for pure states ρ^0=|ψ0ψ0|\hat{\rho}_{0}=|\psi_{0}\rangle\langle\psi_{0}|. It should be noticed that, for pure states, t(ϕ)\mathscr{F}_{t}(\phi) writes as an out-of-time-order correlator, t(ϕ)=W^t(ϕ)V^0W^t(ϕ)V^0\mathscr{F}_{t}(\phi)=\big{\langle}\hat{W}_{t}^{\dagger}(\phi)\hat{V}_{0}^{\dagger}\hat{W}_{t}(\phi)\hat{V}_{0}\big{\rangle} for V^0=|ψ0ψ0|\hat{V}_{0}=|\psi_{0}\rangle\langle\psi_{0}| and the mean value calculated on the initial state |ψ0|\psi_{0}\rangle.

While exact numerical calculations of the QFI for the model (1) can be obtained for long times and large systems, it is interesting to study various approximations to the quantum dynamics of the QFI. For instance, Ref. [47] has considered an exact quantum phase model, Refs. [82, 83] have used the a Fock state expansion [54, 84] for the study of the spin-squeezing parameter, which is a lower bound to the QFI [37], while Ref. [29] has proposed a semiclassical approximation. In this paper, we study the time evolution of the QFI within a BMF approach [48].

Here, rather than using a Schrödinger picture, we consider the dynamics of the spin operators. According to the Heisenberg equation for the operator J^l\hat{J}_{l}, i(dJ^l/dt)=[J^l,H^(t)]i({d\hat{J}_{l}}/{dt})=[\hat{J}_{l},\hat{H}(t)], the expectation value of J^l\hat{J}_{l} depends not only on themselves, but also on the second-order moments J^lJ^k\langle\hat{J}_{l}\hat{J}_{k}\rangle. Similarly, the time evolution of the second-order moments depends on third-order moments, and so on. Consequently, we can obtain the hierarchy of equations of motion for the expectation values J^l\langle\hat{J}_{l}\rangle, J^lJ^k\langle\hat{J}_{l}\hat{J}_{k}\rangle, \dots. The second-order truncation of the hierarchy equations and the approximation J^lJ^kJ^jJ^lJ^kJ^j+J^lJ^kJ^j+J^lJ^jJ^k2J^lJ^kJ^j\langle\hat{J}_{l}\hat{J}_{k}\hat{J}_{j}\rangle\simeq\langle\hat{J}_{l}\hat{J}_{k}\rangle\langle\hat{J}_{j}\rangle+\langle\hat{J}_{l}\rangle\langle\hat{J}_{k}\hat{J}_{j}\rangle+\langle\hat{J}_{l}\hat{J}_{j}\rangle\langle\hat{J}_{k}\rangle-2\langle\hat{J}_{l}\rangle\langle\hat{J}_{k}\rangle\langle\hat{J}_{j}\rangle will give the equations of motion of the expectation values for the first- and second-order moments, which are [48]

ddt(sxsyszΔxzΔyzΔxyΔxxΔyyΔzz)=(0csz00c20000csz0a(t)c2000000a(t)00000000000csza(t)00csy000csz000a(t)csxa(t)000csxa(t)csy0cszcsz00002csy02csz00000002(csxa(t))2csz00000002a(t)0000)(sxsyszΔxzΔyzΔxyΔxxΔyyΔzz),\frac{d}{dt}\left(\begin{array}[]{c}s_{x}\\ s_{y}\\ s_{z}\\ \Delta_{xz}\\ \Delta_{yz}\\ \Delta_{xy}\\ \Delta_{xx}\\ \Delta_{yy}\\ \Delta_{zz}\\ \end{array}\right)=\left(\begin{array}[]{ccccccccc}0&-cs_{z}&0&0&-\frac{c}{2}&0&0&0&0\\ cs_{z}&0&-a(t)&\frac{c}{2}&0&0&0&0&0\\ 0&a(t)&0&0&0&0&0&0&0\\ 0&0&0&0&-cs_{z}&a(t)&0&0&-cs_{y}\\ 0&0&0&cs_{z}&0&0&0&a(t)&cs_{x}-a(t)\\ 0&0&0&cs_{x}-a(t)&-cs_{y}&0&cs_{z}&-cs_{z}&0\\ 0&0&0&-2cs_{y}&0&-2cs_{z}&0&0&0\\ 0&0&0&0&2(cs_{x}-a(t))&2cs_{z}&0&0&0\\ 0&0&0&0&2a(t)&0&0&0&0\\ \end{array}\right)\left(\begin{array}[]{c}s_{x}\\ s_{y}\\ s_{z}\\ \Delta_{xz}\\ \Delta_{yz}\\ \Delta_{xy}\\ \Delta_{xx}\\ \Delta_{yy}\\ \Delta_{zz}\\ \end{array}\right), (10)

where (sx,sy,sz)=2(J^x,J^y,J^z)/N(s_{x},s_{y},s_{z})={2}(\langle\hat{J}_{x}\rangle,\langle\hat{J}_{y}\rangle,\langle\hat{J}_{z}\rangle)/{N} and Δlk=4(J^lJ^k+J^kJ^l2J^lJ^k)/N2\Delta_{lk}={4}(\langle\hat{J}_{l}\hat{J}_{k}+\hat{J}_{k}\hat{J}_{l}\rangle-2\langle\hat{J}_{l}\rangle\langle\hat{J}_{k}\rangle)/{N^{2}}. This BMF model is also referred to as the Bogoliubov backreaction model [48, 85]. Taking N2Δlk/8{N^{2}}\Delta_{lk}/{8} as the matrix elements can also form a similar covariance matrix 𝚲B\mathbf{\Lambda}_{B} with Δlk=Δkl\Delta_{lk}=\Delta_{kl} and four times the maximum eigenvalue gives the QFI within the BMF approximation, namely, FBF_{B}.

III Benchmark models

While the main focus of this paper is on the QKR model, which is discussed in the next section, we first analyze here the QFI calculated within the BMF approach for the OAT and TAT models.

III.1 One-axis-twisting dynamics

For the OAT model, it is possible to obtain an analytical expression for the QFI [68, 30, 37, 86] at all times and for arbitrary number of particles. For times ctNπ/22Nct\leq N\pi/2-2\sqrt{N}, the QFI is given by

FQ(t)N=1+(N1)4[αQ(t)+αQ2(t)+βQ2(t)],\frac{F_{Q}(t)}{N}=1+\frac{(N-1)}{4}\big{[}\alpha_{Q}(t)+\sqrt{\alpha_{Q}^{2}(t)+\beta_{Q}^{2}(t)}\big{]}, (11)

where αQ(t)=1cosN2(2ct/N)\alpha_{Q}(t)=1-\cos^{N-2}({2ct}/{N}) and βQ(t)=4sin(ct/N)cosN2(ct/N)\beta_{Q}(t)=4\sin({ct}/{N})\cos^{N-2}({ct}/{N}). The short-time evolution of the QFI is characterized by a sudden increase starting from FQ(0)=NF_{Q}(0)=N, see Fig. 1. We consider N1N\gg 1, such that N1N2NN-1\simeq N-2\simeq N. A Taylor expansion gives

FQ(t)N=1+ct+c2t22+c3t383c3t34N+O(c4t4).\frac{F_{Q}(t)}{N}=1+ct+\frac{c^{2}t^{2}}{2}+\frac{c^{3}t^{3}}{8}-\frac{3c^{3}t^{3}}{4N}+O\big{(}c^{4}t^{4}\big{)}. (12)

Notice that, up to the second order in ctct, the Taylor expansion is analogous to that of ecte^{ct}. For times ct1ct\geq 1 we have αQ(t)βQ(t)\alpha_{Q}(t)\gg\beta_{Q}(t) and Eq. (11) becomes

FQ(t)N1+NαQ(t)2.\frac{F_{Q}(t)}{N}\simeq 1+\frac{N\alpha_{Q}(t)}{2}. (13)

For times 1ct2N1\leq ct\leq 2\sqrt{N} we have αQ(t)2c2t2/N\alpha_{Q}(t)\simeq 2c^{2}t^{2}/N such that

FQ(t)Nc2t2,\frac{F_{Q}(t)}{N}\simeq c^{2}t^{2}, (14)

having a quadratic scaling with time. For ct2Nct\geq 2\sqrt{N}, we have cosN2(2ct/N)1\cos^{N-2}({2ct}/{N})\ll 1 and thus obtain αQ(t)1\alpha_{Q}(t)\simeq 1. Equation (13) thus predicts a constant plateau FQ(t)=N2/2F_{Q}(t)=N^{2}/2 as a function of time, up to ctNπ/22Nct\leq N\pi/2-2\sqrt{N}. Finally, for ctNπ/22Nct\geq N\pi/2-2\sqrt{N} the QFI further increases and reaches the maximum value FQ(t)=N2F_{Q}(t)=N^{2} at time ct=Nπ/2ct=N\pi/2. It should be mentioned that the behavior of QFI given by Eq. (11) is obviously πN\pi N-periodic.

We compare the exact QFI to that calculated within the BMF approach, which can also be solved analytically. For the initial state sx(0)=1s_{x}(0)=1, the covariance matrix 𝚲B\mathbf{\Lambda}_{B} reads

𝚲B(t)=N28(00002+N(1cos2ctN)N2sinctNN002N).\mathbf{\Lambda}_{B}(t)=\frac{N^{2}}{8}\left(\begin{array}[]{ccc}0&0&0\\ 0&\frac{2+N(1-\cos\frac{2ct}{\sqrt{N}})}{N}&\frac{2\sin\frac{ct}{\sqrt{N}}}{\sqrt{N}}\\ 0&0&\frac{2}{N}\\ \end{array}\right). (15)

The QFI is given by four times the maximum eigenvalue of 𝚲B(t)\mathbf{\Lambda}_{B}(t), namely

FB(t)N=1+NαB(t)+8NαB(t)+N2αB2(t)4,\frac{F_{B}(t)}{N}=1+\frac{N\alpha_{B}(t)+\sqrt{8N\alpha_{B}(t)+N^{2}\alpha_{B}^{2}(t)}}{4}, (16)

where αB(t)=2sin2(ct/N)\alpha_{B}(t)=2\sin^{2}({ct}/{\sqrt{N}}). Expanding Eq. (16) in a Taylor series for ct/N1{ct}/{N}\ll 1, we obtain

FB(t)N=1+ct+c2t22+c3t38c3t36N+O(c4t4).\frac{F_{B}(t)}{N}=1+ct+\frac{c^{2}t^{2}}{2}+\frac{c^{3}t^{3}}{8}-\frac{c^{3}t^{3}}{6N}+O\big{(}c^{4}t^{4}\big{)}. (17)

Equation (17) coincides with Eq. (12) up to the second order in ctct. The difference in the third order is the term 7c3t3/12N7c^{3}t^{3}/12N, which is negligible asymptotically for large NN. Similar considerations can be obtained for higher order terms in the Taylor expansion. For time ct1ct\geq 1 we have αB2(t)αB(t)\alpha_{B}^{2}(t)\gg\alpha_{B}(t) and Eq. (16) becomes

FB(t)N1+NαB(t)2,\frac{F_{B}(t)}{N}\simeq 1+\frac{N\alpha_{B}(t)}{2}, (18)

in analogy with Eq. (13). A Taylor expansion for ct/N1ct/\sqrt{N}\gg 1, predicts the quadratic scaling

FB(t)Nc2t2,\frac{F_{B}(t)}{N}\simeq c^{2}t^{2}, (19)

for 1ctN1\leq ct\leq\sqrt{N}, in agreement with Eq. (14). For longer times, Eq. (16) is characterized by harmonic oscillations of period Nπ/c\sqrt{N}\pi/c. At half period, namely, t=mNπ/2ct={m\sqrt{N}\pi}/{2c} with m=1,3,m=1,3,\dots, αB=2N\alpha_{B}=2N, Eq. (16) (wrongly) predicts the Heisenberg limit FBN2F_{B}\simeq N^{2}, see Fig. 1. We thus conclude that the BMF approach fails for times ctNct\geq\sqrt{N}.

A different approximation to the exact QFI dynamics has been recently studied in Refs. [87, 29] using the Holstein-Primakoff (HP) transformation around the instantaneous mean spin polarization (lying on the xx axis, in the present case). Hence, the HP transformation is performed by mapping the transverse (zz and yy) spin components to canonical bosonic variables

J^x=N2n^,J^yN2p^,J^zN2q^,\hat{J}_{x}=\frac{N}{2}-\hat{n},\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \hat{J}_{y}\simeq\sqrt{\frac{N}{2}}\hat{p},\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \hat{J}_{z}\simeq\sqrt{\frac{N}{2}}\hat{q}, (20)

where the bosonic variables q^\hat{q} and p^\hat{p} satisfy [q^,p^]=i[\hat{q},\hat{p}]=i, n^=(q^2+p^21)/2\hat{n}=({\hat{q}^{2}+\hat{p}^{2}-1})/{2}. The quantum number nn (=0,1,2,=0,1,2,\dots) labels the quantized spin projection along the xx direction of the classical minimum. Using Eqs. (20), the Hamiltonian (1) can be approximated by neglecting terms of order 1/N{1}/{N},

H^(t)a(t)(N2n^)+cq^22,\hat{H}(t)\simeq a(t)\left(\frac{N}{2}-\hat{n}\right)+c\frac{\hat{q}^{2}}{2}, (21)

and thus one can easily obtain the equations of motion for the quantum fluctuations

dq^dt=a(t)p^,dp^dt=[a(t)c]q^.\frac{d\hat{q}}{dt}=-a(t)\hat{p},\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \frac{d\hat{p}}{dt}=[a(t)-c]\hat{q}. (22)

This evolution of the quantum fluctuations is exact at the quadratic level and corresponds to the classical one. We start from initial coherent spin state sx(0)=1s_{x}(0)=1, for which q^2(0)=p^2(0)=1/2\langle\hat{q}^{2}(0)\rangle=\langle\hat{p}^{2}(0)\rangle={1}/{2} and q^(0)p^(0)=p^(0)q^(0)=0\langle\hat{q}(0)\hat{p}(0)\rangle=\langle\hat{p}(0)\hat{q}(0)\rangle=0. As a result, the QFI can be given by [87, 29]

FHP(t)N=1+2n^exc(t)+2n^exc(t)(n^exc(t)+1),\frac{F_{\rm HP}(t)}{N}=1+2\langle\hat{n}_{\mathrm{exc}}(t)\rangle+2\sqrt{\langle\hat{n}_{\mathrm{exc}}(t)\rangle(\langle\hat{n}_{\mathrm{exc}}(t)\rangle+1)}, (23)

where n^exc(t)=(q^2+p^21)/2\langle\hat{n}_{\mathrm{exc}}(t)\rangle=({\langle\hat{q}^{2}\rangle+\langle\hat{p}^{2}\rangle-1})/{2} is the number of collective excitations. For the OAT model, n^exc(t)=c2t2/4\langle\hat{n}_{\mathrm{exc}}(t)\rangle=c^{2}t^{2}/4. The Taylor expansion at short times leads to [87, 29, 88]

FHP(t)N=1+ct+c2t22+c3t38+O(c5t5).\frac{F_{\rm HP}(t)}{N}=1+ct+\frac{c^{2}t^{2}}{2}+\frac{c^{3}t^{3}}{8}+O(c^{5}t^{5}). (24)

The HP approximation is expected to hold for time sufficiently long as far as the mean spin length is J^xN/2\big{\langle}\hat{J}_{x}\big{\rangle}\simeq N/2, namely, as far as we can locally approximate the generalize (NN-spin) Bloch sphere with the tangent plane [37]. Using the exact result J^x=(N/2)cosN1(ct)\big{\langle}\hat{J}_{x}\big{\rangle}=(N/2)\cos^{N-1}(ct) [68], we obtain the HP approximation is expected to hold up to times ct=O(N)ct=O(\sqrt{N}). At intermediate times, for 1ctN1\ll ct\ll\sqrt{N}, we can approximate

FHP(t)Nc2t2,\frac{F_{\rm HP}(t)}{N}\simeq c^{2}t^{2}, (25)

predicting a quadratic scaling of the QFI with time, in agreement with Eqs. (14) and (19).

In Fig. 1(a), we compare the exact FQ(t)F_{Q}(t), Eq. (12), the BMF approximation FB(t)F_{B}(t), Eq. (17), and the HP approximation, Eq. (24), as a function of time, up to ctNct\leq\sqrt{N}. As discussed above, we see that all models agree very well for short times, see also Refs. [47, 83]. We also see that, as expected, the BMF approach is closer than the HP approximation to the exact solution. This is mainly because the BMF approximation (10) considers both the coupling between quantum fluctuations and the mean spin value and the coupling between fluctuations, while the semiclassical method based on HP transformation only considers the coupling between quantum fluctuations.

Refer to caption
Figure 1: QFI dynamics for (a) the OAT model [a(t)=0a(t)=0 in Eq. (1)] and (b) the TAT model [for A=π/2A={\pi}/{2}]. The solid black line is the exact solution (see text), the dashed line is the BMF approach FB(t)F_{B}(t), while the FHPF_{\rm HP} is the semiclassical HP Eq. (23). We have set c=πc=\pi, N=400N=400 and an initial coherent spin state pointing along the xx axis (sx=1s_{x}=1).

III.2 Twist-and-turn dynamics

The exact dynamics for FQ(t)F_{Q}(t) is not analytical, except for weak interaction [89]. The most interesting situation is that corresponding to the unstable regime [47] (i.e., c>Ac>A). The quantity FB(t)F_{B}(t) is obtained by diagonalizing the covariance matrix 𝚲B\mathbf{\Lambda}_{B} on the basis of solving Eq. (10) numerically. For short time, we can obtain an analytical solution when taking sx(t)1(A2+λ2)2t2/2NA2s_{x}(t)\simeq 1-(A^{2}+\lambda^{2})^{2}t^{2}/2NA^{2} with λ2=A(cA)\lambda^{2}=A(c-A). The Taylor expansion reads

FB(t)N=\displaystyle\frac{F_{B}(t)}{N}= 1+ct+c2t22+cλ2t36+c3t38c3t32N\displaystyle 1+ct+\frac{c^{2}t^{2}}{2}+\frac{c\lambda^{2}t^{3}}{6}+\frac{c^{3}t^{3}}{8}-\frac{c^{3}t^{3}}{2N} (26)
+c2λ2t46c4t42N+O(c5t5),\displaystyle+\frac{c^{2}\lambda^{2}t^{4}}{6}-\frac{c^{4}t^{4}}{2N}+O(c^{5}t^{5}),

that agrees with Ref. [47]. Similarly to the OAT [83], the Taylor expansion up to the second order is analogous to that of ecte^{ct}. Nevertheless, the third order of Eq. (26) is larger than that of Eq. (17) by a factor λ2/c2=A(1A/c)/c{\lambda^{2}}/{c^{2}}={A}(1-{A}/{c})/{c}, that has a maximum at c/A=2c/A=2 [47]. We can thus conclude that the BMF approximation breaks down for times of the order of

tc1λlog(Nλ2cA),t_{c}\simeq\frac{1}{\lambda}\log\bigg{(}\frac{N\lambda^{2}}{cA}\bigg{)}, (27)

where tct_{c} has been identified in Ref. [77] as the time needed to generate a macroscopic superposition state.

We can compare this prediction with the semiclassical HP approach of Ref. [29]. The analytical solution of FHP(t)F_{\rm HP}(t) is still given by Eq. (23) where, in the unstable regime, n^exc(t)=(A/λ+λ/A)2sinh2(λt)/4\langle\hat{n}_{\mathrm{exc}}(t)\rangle=({A}/{\lambda}+{\lambda}/{A})^{2}\sinh^{2}(\lambda t)/{4} [88]. The Taylor expansion of FHP(t)F_{\rm HP}(t) at short times reads [87, 29, 88]

FHP(t)N=1+ct+c2t22+cλ2t36+c3t38+c2λ2t46+O(c5t5).\frac{F_{\rm HP}(t)}{N}=1+ct+\frac{c^{2}t^{2}}{2}+\frac{c\lambda^{2}t^{3}}{6}+\frac{c^{3}t^{3}}{8}+\frac{c^{2}\lambda^{2}t^{4}}{6}+O(c^{5}t^{5}). (28)

Similarly to the OAT case, the agreement between Eqs. (26) and (28) is up to the second order of ctct. The difference in the third order term is c3t3/2Nc^{3}t^{3}/2N, which is negligible asymptotically for large NN. Similar considerations can be obtained for higher order terms in the Taylor expansion. Furthermore, neglecting subleading terms for N1N\gg 1, Eqs. (26) and (28) agree with the results of Ref. [47]. Finally, in the limit n^exc(t)1\langle\hat{n}_{\mathrm{exc}}(t)\rangle\gg 1, that is obtained after a transient time, c/λctlog(Nλ2/cA)c/\lambda\leq ct\leq\log(N\lambda^{2}/cA), the HP approach predicts [29, 88]

FHP(t)N(Aλ+λA)e2λt,\frac{F_{\rm HP}(t)}{N}\simeq\bigg{(}\frac{A}{\lambda}+\frac{\lambda}{A}\bigg{)}e^{2\lambda t}, (29)

where the largest instability parameter λ/c\lambda/c is obtained for c/A=2c/A=2. A comparison between the exact FQ(t)F_{Q}(t) and the approximations FB(t)F_{B}(t) and FHP(t)F_{\rm HP}(t) is shown in Fig. 1(b).

IV Quantum kicked rotor model

In the following we consider the case

a(t)=ATτ1n=0δ(tτn),a(t)=\frac{AT}{\tau_{1}}\sum_{n=0}^{\infty}\delta(t-\tau_{n}), (30)

such that the two-mode time-dependent Hamiltonian (1) becomes

H^(t)=ATτ1n=0δ(tτn)J^x+cNJ^z2,\hat{H}(t)=\frac{AT}{\tau_{1}}\sum_{n=0}^{\infty}\delta(t-\tau_{n})\hat{J}_{x}+\frac{c}{N}\hat{J}_{z}^{2}, (31)

where AA is the strength of the kicked field. The Dirac δ\delta-function presents a rectangular-pulse-type kicked field with τn[nT,nT+τ1]\tau_{n}\in[nT,nT+\tau_{1}]. Here, T=τ0+τ1T=\tau_{0}+\tau_{1} is the period of the kicked field with τ0\tau_{0} and τ1\tau_{1} being the interval and the width of the kicks, respectively. Equation (31) corresponds to the QKR model [18, 19]. The periodicity of the two-mode QKR model (31) allows us to describe stroboscopically the evolution of quantum state with discrete time in units of TT, i.e.,

|ψ(n)=UR(T)|ψ(n1)=URn(T)|ψ(0),|\psi(n)\rangle=U_{R}(T)|\psi(n-1)\rangle=U_{R}^{n}(T)|\psi(0)\rangle, (32)

with the unitary Floquet operator

UR(T)=ei[cNJ^z2]τ0ei[ATτ1J^x+cNJ^z2]τ1.U_{R}(T)={e}^{-i[\frac{c}{N}\hat{J}_{z}^{2}]\tau_{0}}{e}^{-i[\frac{AT}{\tau_{1}}\hat{J}_{x}+\frac{c}{N}\hat{J}_{z}^{2}]\tau_{1}}. (33)

When τ10\tau_{1}\rightarrow 0, τ0T\tau_{0}\rightarrow T, the kicked field becomes a δ\delta-function-type form and the Floquet operator reduces to

Uδ(T)=ei[cNJ^z2]Tei[AJ^x]T.U_{\delta}(T)={e}^{-i[\frac{c}{N}\hat{J}_{z}^{2}]T}{e}^{-i[{A}\hat{J}_{x}]T}. (34)

In our calculations, we adopt a rectangle-pulse-type kicked field by setting τ0=1\tau_{0}=1 and τ1=0.01\tau_{1}=0.01, which is very close to the δ\delta-function-type kicked field.

Refer to caption
Figure 2: (a) Maximum Lyapunov exponents for the initial state sx(0)=1s_{x}(0)=1 [i.e., (sz,ϕ)=(0,0)(s_{z},\phi)=(0,0)]. (b), (c), and (d) show the Poincaré sections with A/π=0.4A/\pi=0.4 for c/π=0.2c/\pi=0.2, 0.80.8, and 1.41.4, respectively. Green solid points indicate the initial states adopted in (a).

IV.1 Mean-field approximation and chaotic motion

Here, the total number NN is conserved and 1/N1/N can be identified with an effective Planck constant. The mean-field (MF) limit NN\rightarrow\infty (and thus the interaction c/N0c/N\rightarrow 0) corresponds to the classical limit, where (sx,sy,sz)(s_{x},s_{y},s_{z}) become classical variables that form a single-particle Bloch vector confined to the unit sphere and thus the norm is conserved [48]. (sz,ϕ)(s_{z},\phi) can be identified with classical phase-space variables, where ϕ\phi is the azimuthal angle. The first-order truncation of the hierarchy equations together with the approximations J^lJ^kJ^lJ^k\langle\hat{J}_{l}\hat{J}_{k}\rangle\simeq\langle\hat{J}_{l}\rangle\langle\hat{J}_{k}\rangle give the equations of motion of the expectation values for the first-order moments, namely, the MF equation [48]

ddt(sxsysz)=(0csz0csz0a(t)0a(t)0)(sxsysz),\frac{d}{dt}\left(\begin{array}[]{c}s_{x}\\ s_{y}\\ s_{z}\\ \end{array}\right)=\left(\begin{array}[]{ccc}0&-cs_{z}&0\\ cs_{z}&0&-a(t)\\ 0&a(t)&0\\ \end{array}\right)\left(\begin{array}[]{c}s_{x}\\ s_{y}\\ s_{z}\\ \end{array}\right), (35)

where a(t)a(t) is given by Eq. (30). For such a periodic-driven system, the classical motion can usually be divided into two classes: i) stable and periodic motion and ii) unstable and chaotic motion. In the unstable case, the magnification of the initially small deviation is exponential in time, which can be measured by the maximum Lyapunov exponent, i.e., [90]

λL(A,c)=limm1mTn=1mln(δsx(nT),δsy(nT),δsz(nT))(δsx(0),δsy(0),δsz(0)),\lambda_{L}(A,c)=\lim_{m\rightarrow\infty}\frac{1}{mT}\sum_{n=1}^{m}\ln\frac{||(\delta s_{x}(nT),\delta s_{y}(nT),\delta s_{z}(nT))||}{||(\delta s_{x}(0),\delta s_{y}(0),\delta s_{z}(0))||}, (36)

where δsx,y,z(0)=|sx,y,z(0)sx,y,z(0)|\delta s_{x,y,z}(0)=|s^{\prime}_{x,y,z}(0)-s_{x,y,z}(0)| denote the initial deviations and δsx,y,z(nT)=|sx,y,z(nT)sx,y,z(nT)|\delta s_{x,y,z}(nT)=|s^{\prime}_{x,y,z}(nT)-s_{x,y,z}(nT)| denote the deviations at t=nTt=nT, where sx,y,z(t)s_{x,y,z}(t) and sx,y,z(t)s^{\prime}_{x,y,z}(t) respectively represent the two classical trajectories without perturbation and with perturbation initially. In Fig. 2(a) we plot the maximum Lyapunov exponents in the (A,c)(A,c)-plane for the initial state sx(0)=1s_{x}(0)=1. In our numerical calculations we set δsx(0)=105\delta s_{x}(0)=10^{-5} and evolve the system for 500 periods. For the stable motion, λL\lambda_{L} is zero (dark blue region), while for the chaotic regime λL\lambda_{L} is positive. For example, in Fig. 2(a) for A/π=0.4A/\pi=0.4 and c/π=0.2c/\pi=0.2, the initial state is in stable region, while A/π=0.4A/\pi=0.4 both for c/π=0.8c/\pi=0.8 and c/π=1.4c/\pi=1.4 the initial states are in chaotic region. However, the point (A,c)/π=(0.4,1.4)(A,c)/\pi=(0.4,1.4) is “more chaotic” than (A,c)/π=(0.4,0.8)(A,c)/\pi=(0.4,0.8) is, because (A,c)/π=(0.4,1.4)(A,c)/\pi=(0.4,1.4) (λL1.524\lambda_{L}\simeq 1.524) has a larger Lyapunov exponent than (A,c)/π=(0.4,0.8)(A,c)/\pi=(0.4,0.8) (λL0.966\lambda_{L}\simeq 0.966), meaning that δsx,y,z(nT)\delta s_{x,y,z}(nT) deviate faster from the initial conditions.

To show the property of classical motion of the kicked system, we plot the Poincaré sections by using the stroboscopic points at t=nTt=nT in Figs. 2(b)-2(d) for different interaction strengths c/π=0.2c/\pi=0.2, 0.80.8, and 1.41.4 with same kicked strength A/π=0.4A/\pi=0.4. We see that for the integrable situation [e.g., (A,c)/π=(0.4,0.2)(A,c)/\pi=(0.4,0.2), see Fig. 2(b)], all trajectories (periodic orbit and elliptic fixed point) are stable with the maximum Lyapunov exponents being zero. In the completely nonintegrable case [e.g., (A,c)/π=(0.4,1.4)(A,c)/\pi=(0.4,1.4), see Fig. 2(d)], we see that the phase space is full of chaotic trajectories with largely positive λL\lambda_{L}. In the mixed case [e.g., (A,c)/π=(0.4,0.8)(A,c)/\pi=(0.4,0.8), see Fig. 2(c)], the phase space illustrates one big-size island, six middle-size islands, and four small-size islands (see the blank regions). It is clear that the motions inside the islands are stable and the motions outside the islands are unstable and mainly chaotic with medium positive λL\lambda_{L}. In Figs. 2(b)-2(d), the green points mark the initial state sx(0)=1s_{x}(0)=1.

IV.2 Quantum Fisher Information and maximum Lyapunov exponent

Refer to caption
Figure 3: Time evolution of QFI for the QKR (A/π=0.4A/\pi=0.4, black solid lines), the OAT (A=0A=0, red dashed lines), and the TAT (A/π=0.4A/\pi=0.4, green dash-dotted lines) models with different interaction strengths: (a) c/π=0.2c/\pi=0.2 (the initial state being a stable point), (b) c/π=0.5c/\pi=0.5 (the initial state being a saddle point), (c) c/π=0.8c/\pi=0.8 [The initial state is in chaotic regime, see Fig. 2(c)], and (d) c/π=2.0c/\pi=2.0 (The initial state is in deep chaotic regime). We have set N=200N=200 and sx(0)=1s_{x}(0)=1.

To show the influence of the chaotic motions on the dynamics of entanglement, we calculate the QFI for the same initial state [sx(0)=1s_{x}(0)=1] and different interacting parameters. In Fig. 3, for a fixed value of the parameter cc, we compare the time-evolution of the QFI between the QKR model (solid black line; obtained for parameter AA with δ\delta-kicked time dependence), the OAT model (dashed red line; A=0A=0), and the TAT model (dot-dashed green line; for constant AA). When the initial quantum state corresponds to a stable point (i.e., c/π=0.2c/\pi=0.2) or a saddle point (i.e., c/π=0.5c/\pi=0.5) in the Poincaré section of the classical model, the QKR model does not show any advantage in entanglement generation because of the small oscillation [see Fig. 3(a)] or the slow growth [see Fig. 3(b)] of the QFI with time. However, when the initial quantum state is located in the classical chaotic region (i.e., c/π=0.8c/\pi=0.8, and 2.0), compared with the QFI for the OAT model, the QFI for the QKR model can reach and exceed N2/2{N^{2}}/{2} in a very short period of time (e.g., 2 or 3 kicks), which indicates that the quantum chaos can significantly accelerate the generation of entanglement [see Fig. 3(c) and 3(d)]. When compared with the TAT model, the QFI for the QKR model is slower in all cases. In particular, it was shown in Ref. [47] that the maximally unstable TAT dynamics (for c/A=2c/A=2) is characterized by the fastest generation of entanglement within quadratic Hamiltonians as in Eq. (1).

Refer to caption
Figure 4: Comparison of QFI dynamics for the QKR model with the initial state sx(0)=1s_{x}(0)=1 being (a) a stable fixed point (i.e., c/π=0.2c/\pi=0.2), (b) a saddle point (i.e., c/π=0.5c/\pi=0.5), (c) in chaotic regime (i.e., c/π=0.8c/\pi=0.8), and (d) in deep chaotic regime (i.e., c/π=1.4c/\pi=1.4). The solid black line is the exact solution FQ(t)F_{Q}(t), the dashed line is the BMF approach FB(t)F_{B}(t), while the FHPF_{\rm HP} is the semiclassical HP Eq. (23). We have set A/π=0.4A/\pi=0.4 and N=400N=400.
Refer to caption
Figure 5: Top [Bottom] row panels show the break time tIBt_{IB}, as defined in Eq. (37) [tHPt_{\rm HP}, Eq. (42)], as a function of the total particle number NN for the QKR model. The different panels corresponds to the initial state being (a,d) a classical stable point, (b,e) a classical saddle point, and (c,f) in the deep chaotic region. In all panels, symbols are numerical results while lines are fits, see text. Values of Hamiltonian parameters are reported in each panel, while parameters of the parameters of the fitting lines are reoported in the text. Panel (a,d) shows the results for the OAT model (black diamonds) and the QKR model in the stable region (red squares). Panel (b,e) shows the results for the TAT model (black diamonds) and the QKR model at the saddle point (red squares). Panel (c,f) shows the results of the QKR model in the chaotic regime.
Refer to caption
Figure 6: Break time tIBt_{IB} as a function of the maximum Lyapunov exponent λL\lambda_{L}. Symbols are numerical results, lines are fits according to Eq. (41). In panel (a), the parameter AA is fixed and the Lyapunov exponent only depends on cc. The fitting lines are 2.83×e1.46λL2.83\times{e}^{-1.46\lambda_{L}} (solid line), 2.38×e1.22λL2.38\times{e}^{-1.22\lambda_{L}} (dash line), and 1.82×e1.03λL1.82\times{e}^{-1.03\lambda_{L}} (dash-dot line), respectively. In panel (b), the parameter cc is fixed and the maximum Lyapunov exponent is a double-valued function of AA. The display of the lines between the scattered points is only for guiding eyes. The initial state is sx(0)=1s_{x}(0)=1 and we set N=300N=300.

In Fig. 4, we compare the exact FQ(t)F_{Q}(t) with both the BMF approximation FB(t)F_{B}(t) and the HP approximation FHP(t)F_{\rm HP}(t) for the periodic-kicked system with the initial state sx(0)=1s_{x}(0)=1. When the initial state is a classical stable point, the three QFI agree well for a long time [see Fig. 4(a)]. When the initial state is an unstable saddle point [see Fig. 4(b)] or in the classical chaotic region [see Figs. 4(c) and 4(d)] the three results remain consistent only for a short time (a few kicks in the figure). While the FHP(t)F_{\rm HP}(t) diverges after a transient time, the BMF follows more closely the exact results for longer times. It is thus interesting to introduce the criterion

|FQ(tIB)FB(tIB)|FQ(tIB)=gIB,\frac{|F_{Q}(t_{IB})-F_{B}(t_{IB})|}{F_{Q}(t_{IB})}=g_{IB}, (37)

to quantitatively define a break time tIBt_{IB} where the BMF approximation departs from the exact result (by a relative amount gIBg_{IB}). This time is closely related to the instability degree of the initial state, which we will relate below to the maximum Lyapunov exponent. In practice, we set gIB=0.01g_{IB}=0.01 in the following. The main result of this paper is to study numerically the break time tIBt_{IB} in different chaotic regimes and determine characteristic behaviours. We consider different regimes:

  • \bullet

    If the initial condition is in the classical stable region, we find

    tIBαst(A,c)×N.t_{IB}\simeq\alpha_{\rm st}(A,c)\times\sqrt{N}. (38)

    In particular, we can demonstrate analytically that the O(N)O(\sqrt{N}) behaviour holds for the OAT dynamics. To support the claim (38) we show numerical results in Fig. 5(a), obtained for the QKR model for A/π=0.4A/\pi=0.4 and c/π=0.2c/\pi=0.2. There, we plot the break time tIBt_{IB} as a function of the number of particles NN. The dashed line in Fig. 5(a) is N/4.7\sqrt{N}/4.7 fitting the OAT results, while the solid line is N/15\sqrt{N}/15 fitting the QKR results for large NN, in agreement with Eq. (38).

  • \bullet

    If the initial condition is at the saddle point, we find

    tIBαsp(A,c)×lnN.t_{IB}\simeq\alpha_{\rm sp}(A,c)\times\ln{N}. (39)

    In particular, this behaviour holds for the TAT dynamics as well as in the QKR model at the saddle point [see Fig. 5(b)]. The dashed line in Fig. 5(b) is lnN/2.02\ln N/2.02 fitting the TAT results, while the solid line is lnN/2.2\ln N/2.2 fitting the QKR results for large NN, in agreement with Eq. (39). Notice that, for the TAT model with initial conditions at the saddle point, we expect tIBtct_{IB}\propto t_{c}, where tc(1/λ)logNt_{c}\sim(1/\lambda)\log N (for N1N\gg 1), as given in Eq. (27), with λ=A(cA)\lambda=\sqrt{A(c-A)}.

  • \bullet

    If the initial conditions are in the deep chaotic regime we find

    tIBαch(A,c)×(lnN)4,t_{IB}\simeq\alpha_{\rm ch}(A,c)\times(\ln{N})^{4}, (40)

    see Fig. 5(c). The solid line in Fig. 5(c) is the fit (lnN)4/1590(\ln N)^{4}/1590, the dashed line is (lnN)4/2810(\ln N)^{4}/2810, the dash-dotted line is (lnN)4/3200(\ln N)^{4}/3200, and the dash-double-dotted line is (lnN)4/4000(\ln N)^{4}/4000. In the chaotic regime, all results agree well with Eq. (40). Notice that we numerically do not see a smooth transition between the O(logN)O(\log N) behaviour at the saddle point and the O((logN)4)O((\log N)^{4}) behaviour in the deep chaotic regime, but rather a sharp transition. Furthermore, we can connect the prefactor αch(A,c)\alpha_{\rm ch}(A,c) in Eq. (40) and the asymptotically maximum Lyapunov exponent λL(A,c)\lambda_{L}(A,c) (depending on both parameters cc and AA). For strong chaos and large NN, we find numerically

    α(A,c)chη(A)eγ(A)λL(A,c),\alpha(A,c)_{\rm ch}\simeq\eta(A){e}^{-\gamma(A)\lambda_{L}(A,c)}, (41)

    where the coefficients γ(A)\gamma(A) and η(A)\eta(A) depend only on the kick strength AA. A numerical investigation of the Eq. (41) is shown in Fig. 6. In Fig. 6(a) we show the break time as a function of λL(A,c)\lambda_{L}(A,c) and for different values of kicked strength, A/π=0.2A/\pi=0.2, 0.30.3, and 0.40.4. Equation (41) predicts that, while fixing AA, the break time tIBt_{IB} depends on cc through an exponential function of the maximum Lyapunov exponent. This behaviour (lines) is well reproduced by the numerical calculations (symbols) for sufficiently large values of λL(A,c)\lambda_{L}(A,c), as shown in Fig. 6(a). Furthermore, when fixing the parameter cc, Fig. 2(a) shows that the maximum Lyapunov exponent in the chaotic region is a double-valued function of AA. Therefore, Eq. (41) also predicts that tIBt_{IB} is double-valued function of AA, for fixed cc. This prediction is confirmed in Fig. 6(b), where we plot tIBt_{IB} as a function of λL(A,c)\lambda_{L}(A,c) for different values of cc.

In Ref. [29], the time evolution of the HP approximation to the QFI, FHP(t)F_{\rm HP}(t), has been related to the local (i.e., finite-time) maximum Lyapunov exponent λt\lambda_{t}, showing that FHP(t)e2λttF_{\rm HP}(t)\sim e^{2\lambda_{t}t}. This relation holds in a transient time window, up to the Ehrenfest time. We notice that the HP approach of Ref. [29] agrees well with our results. In Fig. 5 we plot the break time of the HP approximation, tHPt_{\rm HP}, defined in analogy to tIBt_{IB} as

|FQ(tHP)FB(tHP)|FQ(tHP)=gHP.\frac{|F_{Q}(t_{\rm HP})-F_{B}(t_{\rm HP})|}{F_{Q}(t_{\rm HP})}=g_{\rm HP}. (42)

Panels (d)-(f) of Fig. 5 show tHPt_{\rm HP} as a function of NN. Panel (d) corresponds to the initial state in the classical stable region. Lines in Fig. 5(d) are numerical fit tHP=O(N)t_{\rm HP}=O(\sqrt{N}): the dashed line is N/5.8\sqrt{N}/5.8 fitting the OAT results, while the solid line is N/13\sqrt{N}/13 fitting the QKR results for large NN. Panel (e) corresponds to the saddle point, where tHP=O(logN)t_{\rm HP}=O(\log N): the dashed line is lnN/2.34\ln N/2.34 fitting the TAT results, while the solid line is lnN/2.34\ln N/2.34 fitting the QKR results for large NN. Panel (f) corresponds to the deep chaotic regime, where tHP=O((logN)4)t_{\rm HP}=O((\log N)^{4}): the solid line is the fit (lnN)4/2020(\ln N)^{4}/2020, the dashed line is (lnN)4/3540(\ln N)^{4}/3540, the dash-dotted line is (lnN)4/4070(\ln N)^{4}/4070, and the dash-double-dotted line is (lnN)4/5100(\ln N)^{4}/5100. The scaling with NN is thus analogous to Eqs. (38), (39) and (40), respectively.

Refer to caption
Figure 7: Adiabatic dynamics comparison between full quantum model (FQM) and BMF model. To obtain the population imbalance probability sz(t)s_{z}(t), we adiabatically evolve the system from the initial ground states sz(0)=±1s_{z}(0)=\pm 1 according to both quantum Schrödinger equation and BMF equations with A=vtA=vt (v=103)v=10^{-3}) and c=1c=1.

V Application in identifying a quantum phase transition

So far, we have shown how the BMF theory can describe the entanglement dynamics of quantum many-body kicked system up to transient times. Here, we discuss the application of the BMF approach to identify the critical point of a second-order quantum phase transition (QPT) that occurs in a bosonic Josephson junction. This QPT has been recently explored with a BEC with tuneable interparticle interaction [91, 60] (see Refs. [92, 40, 41, 44] for studies of entanglement in this system). The system is described by the Hamiltonian

H^=AJ^xcNJ^z2,\hat{H}=A\hat{J}_{x}-\frac{c}{N}\hat{J}_{z}^{2}, (43)

that is characterized by a QPT between a gapped symmetric state (for A/c>1A/c>1) to two degenerate asymmetric states (i.e., A/c<1A/c<1), with critical point (in the thermodynamic limit) at A/c=1A/c=1. Instead of direct diagonalization, here we use adiabatic dynamical method to describe and identify the transition between the two regimes in a system with finite number of particles. Initially, we assume that the coupling between the two modes is very weak, and the eigenstates of J^z\hat{J}_{z}, i.e., |μ=±N/2|\mu=\pm{N}/{2}\rangle (or sz(0)=±1s_{z}(0)=\pm 1) are the ground state of the system, which are double degenerate, where all atoms occupy in one mode. Then we adiabatically evolve the system by slowly increasing the coupling strength between the two modes as A=vtA=vt. If the coupling strength changes slowly enough, i.e., v0v\rightarrow 0, the quantum adiabatic theorem ensures that the system is always in the instantaneous ground state. When the coupling exceeds a certain pseudo-phase transition critical point, the ground state of the system will change from mainly occupying a single mode to having two modes equally occupied. This adiabatic dynamics shown in Fig. 7 can be used to illustrate the process of phase transition. Obviously, increasing the number of particles, the consistency of the results between the full quantum model and the BMF model calculations becomes better and better. The oscillations of the order parameter szs_{z} for the phase transition after passing the pseudo-critical point will be suppressed with the decrease of adiabatic evolution velocity vv.

Refer to caption
Figure 8: QFI comparison between full quantum model (FQM) and BMF model. To obtain FB,QF_{B,Q}, we adiabatically evolve the system from the initial ground state sz(0)=1s_{z}(0)=1 according to both quantum Schrödinger equation and BMF equations with A=vtA=vt (v=103)v=10^{-3}) and c=1c=1.

The corresponding QFI of the instantaneous ground state during adiabatic evolution is also calculated and the results are shown in Fig. 8. We find that the QFI has a maximum value at the pseudo-critical point of phase transition. This can be used as an effective signal to recognize the phase transition of finite size systems. It is seen that the BMF approximation becomes more precise when cranking up the number NN. This means that for quantum many-body system with large but finite number of particles, the simple BMF theory can be used to accurately predict the position of phase transition without solving the complex quantum many-body model.

VI Conclusion

To conclude, we have investigated the entanglement dynamics of a quantum kicked rotor model by using a beyond mean-field approach. We have found numerically characteristic scaling laws for the break time tIBt_{IB} when the quantum Fisher information calculated within the BMF approach departs from the exact value. The scaling depends on the properties of the initial conditions: if the initial conditions correspond to a classical regular regime, tIB=O(N)t_{IB}=O(\sqrt{N}); if the initial conditions correspond to a classical saddle point, tIB=O(logN)t_{IB}=O(\log N); otherwise, if the initial conditions correspond to a classical chaotic regime, tIB=O((logN)4)t_{IB}=O((\log N)^{4}). We thus conclude that the entanglement dynamics of unstable saddle point is the most irregular, in the sense that the moment when the quantum effects cannot be captured by the BMF approach occurs earlier than in both regular and chaotic evolution. Furthermore, in the chaotic regime, we have shown a relation between tIBt_{IB} and the maximum Lyapunov exponent, tIBeγλL×(logN)4t_{IB}\sim e^{-\gamma\lambda_{L}}\times(\log N)^{4}, see Eqs. (40) and (41). In particular, we have found that the QFI has a maximum value at the pseudo-critical point of phase transition in a bosonic Josephson junction, which suggests an application in accurately identifying the critical point of quantum phase transition in finite-size systems. While our study is restricted for pure states and it is expected to hold for vanishing decoherence, a study of mixed states will be discussed in a future work.

Acknowledgement

We acknowledge discussions with Alessio Lerose and Silvia Pappalardi. This work is supported by the National Natural Science Foundation of China (Grant No. 11974273) and the Natural Science Fundamental Research Program of Shaanxi Province of China (Grant No. 2019JM-004).

Références

  • [1] E. P. Wigner, On the statistical distribution of the widths and spacings of nuclear resonance levels, Math. Proc. Cambr. Philos. Soc. 47, 790 (1951).
  • [2] A. Einstein, Zum Quantensatz von Sommerfeld und Epstein, Verh. Dtsch. Phys. Ges. 19, 82 (1917). Reprinted in The Collected Papers of Albert Einstein, A. Engel translator (Princeton University Press, Princeton, 1997).
  • [3] A. A. Bychek, P. S. Muraev, and A. R. Kolovsky, Probing quantum chaos in many-body quantum systems by the induced dissipation, Phys. Rev. A 100, 013610 (2019).
  • [4] S. Chaudhury, A. Smith, B. Anderson, S. Ghose, and P. S. Jessen. Quantum signatures of chaos in a kicked top. Nature 461, 768-771 (2009).
  • [5] H.-J. Stöckmann and J. Stein, “Quantum” Chaos in Billiards Studied by Microwave Absorption, Phys. Rev. Lett. 64, 2215 (1990).
  • [6] H.-D. Gräf, H. L. Harney, H. Lengeler, C. H. Lewenkopf, C. Rangacharyulu, A. Richter, P. Schardt, and H. A. Weidenmüller, Distribution of Eigenmodes in a Superconducting Stadium Billiard with Chaotic Dynamics, Phys. Rev. Lett. 69, 1296 (1992).
  • [7] V. Milner, J. L. Hanssen, W. C. Campbell, and M. G. Raizen, Optical Billiards for Atoms, Phys. Rev. Lett. 86, 1514 (2001).
  • [8] N. Friedman, A. Kaplan, D. Carasso, and N. Davidson, Observation of Chaotic and Regular Dynamics in Atom-Optics Billiards, Phys. Rev. Lett. 86, 1518 (2001).
  • [9] Q. Zhang, P. Hanggi, and J. B. Gong, Nonlinear Landau-Zener processes in a periodic driving field, New J. Phys. 10, 073008 (2008).
  • [10] M. Lepers, V. Zehnle, and J. C. Garreau, Tracking Quasiclassical Chaos in Ultracold Boson Gases, Phys. Rev. Lett. 101, 144103 (2008).
  • [11] J. Liu, C. Zhang, M. G. Raizen, and Q. Niu, Transition to instability in a periodically kicked Bose-Einstein condensate on a ring, Phys. Rev. A 73, 013601 (2006).
  • [12] J. Martin, B. Georgeot, and D. L. Shepelyansky, Time Reversal of Bose-Einstein Condensates, Phys. Rev. Lett. 101, 074102 (2008).
  • [13] R. A. Kidd, M. K. Olsen, and J. F. Corney, Quantum chaos in a Bose-Hubbard dimer with modulated tunneling, Phys. Rev. A 100, 013625 (2019).
  • [14] P. Coullet and N. Vandenberghe, Chaotic dynamics of a Bose-Einstein condensate in a double-well trap, J. Phys. B: At. Mol. Opt. Phys. 35, 1593 (2002).
  • [15] J. Cheng, Chaotic dynamics in a periodically driven spin-1 condensate, Phys. Rev. A 81, 023619 (2010).
  • [16] J. Kronjäger, K. Sengstock, and K. Bongs, Chaotic dynamics in spinor Bose-Einstein condensates, New J. Phys. 10, 045028 (2008).
  • [17] J. Cheng, Regular and irregular spin-mixing dynamics in coupled spin-1 atomic and molecular Bose-Einstein condensates, Phys. Rev. A 80, 023608 (2009).
  • [18] F. Haake, M. Kuś, and R. Scharf, Classical and quantum chaos for a kicked top Z. Phys. B Condens. Matter 65, 381 (1987).
  • [19] F. Haake, Quantum Signatures of Chaos (Springer, Berlin, 2010).
  • [20] F. L. Moore, J. C. Robinson, C. F. Bharucha, B. Sundaram, and M. G. Raizen, Atom Optics Realization of the Quantum δ\delta-Kicked Rotor, Phys. Rev. Lett. 75, 4598 (1995).
  • [21] L. Fallani, L. De Sarlo, J. E. Lye, M. Modugno, R. Saers, C. Fort, and M. Inguscio, Observation of Dynamical Instability for a Bose-Einstein Condensate in a Moving 1D Optical Lattice, Phys. Rev. Lett. 93, 140406 (2004).
  • [22] M. Christiani, O. Morsch, N. Malossi, M. Jona-Lasinio, M. Anderlini, E. Courtade, and E. Arimondo, Instabilities of a Bose-Einstein condensate in a periodic potential: an experimental investigation, Opt. Express 12, 4 (2004).
  • [23] G. J. Duffy, S. Parkins, T. Muller, M. Sadgrove, R. Leonhardt, and A. C. Wilson, Experimental investigation of early-time diffusion in the quantum kicked rotor using a Bose-Einstein condensate, Phys. Rev. E 70, 056206 (2004).
  • [24] C. E. Creffield, G. Hur, and T. S. Monteiro, Localization-Delocalization Transition in a System of Quantum Kicked Rotors, Phys. Rev. Lett. 96, 024103 (2006).
  • [25] B. V. Chirikov, A universal instability of many-dimensional oscillator systems, Phys. Rep. 52, 263 (1979).
  • [26] H. J. Stockmann, Quantum Chaos: An Introduction (Cambridge University Press, Cambridge, 1999).
  • [27] R. Lewis-Swan, A. Safavi-Naini, J. Bollinger, and A. Rey, Unifying scrambling, thermalization and entanglement through measurement of fidelity out-of-time-order correlators in the Dicke model, Nat. Comm. 10, 1581 (2019).
  • [28] K. Gietka, J. Chwedeńczuk, T. Wasak, and F. Piazza, Multipartite entanglement dynamics in a regular-to-ergodic transition: Quantum Fisher information approach, Phys. Rev. B 99, 064303 (2019).
  • [29] A. Lerose and S. Pappalardi, Bridging entanglement dynamics and chaos in semiclassical systems, Phys. Rev. A 102, 032404 (2020).
  • [30] L. Pezzè and A. Smerzi, Entanglement, Nonlinear Dynamics, and the Heisenberg Limit, Phys. Rev. Lett. 102, 100401 (2009).
  • [31] The notion of multipartite entanglement is related to that of kk-producibility. A pure state is kk-partite entangled if it is not (k1)(k-1)-producible, namely if it can be written as a product j|ψj\otimes_{j}|\psi_{j}\rangle containing at least one state |ψj|\psi_{j}\rangle of k1k-1 particles that does not factorize. The definition is then extended to mixed states by convexity, see [32, 33, 34].
  • [32] O Gühne and G Tóth, Entanglement detection, Phys. Rep. 474, 1 (2009).
  • [33] P. Hyllus, W. Laskowski, R. Krischek, C. Schwemmer, W. Wieczorek, H. Weinfurter, L. Pezzè, and A. Smerzi, Fisher information and multiparticle entanglement, Phys. Rev. A 85, 022321 (2012).
  • [34] G. Tóth, Multipartite entanglement and high-precision metrology, Phys. Rev. A 85, 022322 (2012).
  • [35] G. Tóth and I. Apellaniz, Quantum metrology from a quantum information science perspective, J. Phys. A 47, 424006 (2014).
  • [36] L. Pezzè, Y. Li, W. Li, and A. Smerzi, Witnessing entanglement without entanglement witness operators, Proc. Natl. Acad. Sci. U.S.A. 113, 11459 (2016).
  • [37] L. Pezzè, A. Smerzi, M. K. Oberthaler, R. Schmied, and P. Treutlein, Quantum metrology with nonclassical states of atomic ensembles, Rev. Mod. Phys. 90, 035005 (2018).
  • [38] C. W. Helstrom, Quantum Detection and Estimation Theory (Academic Press, New York, 1976).
  • [39] S. L. Braunstein, and C. M. Caves, Statistical distance and the geometry of quantum states, Phys. Rev. Lett. 72, 3439 (1994).
  • [40] J. Ma and X. Wang, Fisher information and spin squeezing in the Lipkin-Meshkov-Glick model, Phys. Rev. A 80 012318 (2009).
  • [41] P. Hauke, M. Heyl, L. Tagliacozzo, and P. Zoller, Measuring multipartite entanglement through dynamic susceptibilities. Nat. Phys. 12, 778 (2016).
  • [42] L. Pezzè, M. Gabbrielli, L. Lepori and A. Smerzi, Multipartite entanglement in topological quantum phases, Phys.Rev.Lett. 119 250401 (2017)
  • [43] S. Pappalardi, S. Russomanno, A. Silva, and R. Fazio, Multipartite entanglement after a quantum quench, J. Stat. Mech. 2017, 053104 (2017).
  • [44] M. Gabbrielli, A. Smerzi and L. Pezzè, Multipartite Entanglement at Finite Temperature, Sci. Rep. 8, 15663 (2018).
  • [45] M. Gabbrielli, L. Lepori and L. Pezzè, Multipartite-entanglement tomography of a quantum simulator New J. Phys. 21, 033039 (2019).
  • [46] I. Frérot and T. Roscilde, Reconstructing the quantum critical fan of strongly correlated systems using quantum correlations, Nat. Comm 10, 577 (2019).
  • [47] G. Sorelli, M. Gessner, A. Smerzi, and L. Pezzè, Fast and optimal generation of entanglement in bosonic Josephson junctions, Phys. Rev. A 99, 022329 (2019).
  • [48] J. R. Anglin and A. Vardi, Dynamics of a two-mode Bose-Einstein condensate beyond mean-field theory, Phys. Rev. A 64, 013605 (2001).
  • [49] L. J. Fiderer and D. Braun, Quantum metrology with quantum-chaotic sensors, Nat. Commun. 9, 1351 (2018).
  • [50] J. Javanainen, Oscillatory exchange of atoms between traps containing Bose condensates, Phys. Rev. Lett. 57, 3164 (1986).
  • [51] G. J. Milburn, J. Corney, E. M. Wright, and D. F. Walls, Quantum dynamics of an atomic Bose-Einstein condensate in a double-well potential, Phys. Rev. A 55, 4318 (1997).
  • [52] I. Zapata, F. Sols, and A. J. Leggett, Josephson effect between trapped Bose-Einstein condensates, Phys. Rev. A 57, R28 (1998).
  • [53] A. Smerzi, S. Fantoni, S. Giovanazzi, and S. R. Shenoy, Quantum coherent atomic tunneling between two trapped Bose-Einstein condensates, Phys. Rev. Lett. 79, 4950 (1997).
  • [54] J. Javanainen and M. Yu. Ivanov, Splitting a trap containing a Bose-Einstein condensate: Atom number fluctuations, Phys. Rev. A 60, 2351 (1999).
  • [55] M. Albiez, R. Gati, J. Fölling, S. Hunsmann, M. Cristiani, and M. K. Oberthaler, Direct observation of tunneling and nonlinear self-trapping in a single bosonic Josephson junction, Phys. Rev. Lett. 95, 010402 (2005).
  • [56] T. Schumm, S. Hofferberth, L. M. Andersson, S. Wildermuth, S. Groth, I. Bar-Joseph, J. Schmiedmayer, and P. Krüger, Matter-wave interferometry in a double well on an atom chip, Nat. Phys. 1, 57 (2005).
  • [57] L. Pezzè, L. A. Collins, A. Smerzi, G. P. Berman, and A. R. Bishop, Sub-shot-noise phase sensitivity with a Bose-Einstein condensate Mach-Zehnder interferometer, Phys. Rev. A 72, 043612 (2005).
  • [58] D. Ananikian and T. Bergeman, Gross-Pitaevskii equation for Bose particles in a double-well potential: Two-mode models and beyond, Phys. Rev. A 73, 013604 (2006).
  • [59] R. Gati and M. K. Oberthaler, A bosonic Josephson junction, J. Phys. B: At., Mol. Opt. Phys. 40, R61 (2007).
  • [60] A. Trenkwalder, et al., Quantum phase transitions with parity-symmetry breaking and hysteresis, Nat. Phys. 12, 826 (2016).
  • [61] G. Spagnolli, et al., Crossing over from attractive to repulsive interactions in a tunneling bosonic josephson junction, Phys. Rev. Lett. 118, 230403 (2017).
  • [62] J. I. Cirac, M. Lewenstein, K. Mølmer, and P. Zoller, Quantum superposition states of Bose-Einstein condensates, Phys. Rev. A 57, 1208 (1998).
  • [63] M. J. Steel and M. J. Collett, Quantum state of two trapped Bose-Einstein condensates with a Josephson coupling, Phys. Rev. A 57, 2920 (1998).
  • [64] T. Zibold, E. Nicklas, C. Gross, and M. K. Oberthaler, Classical bifurcation at the transition from Rabi to Josephson dynamics. Phys. Rev. Lett. 105, 204101 (2010).
  • [65] C. Gross, T. Zibold, E. Nicklas, J. Estève, and M. K. Oberthaler, Nonlinear atom interferometer surpasses classical precision limit, Nature 464, 1165 (2010).
  • [66] F. T. Arecchi, E. Courtens, R. Gilmore, and H. Thomas, Atomic Coherent States in Quantum Optics, Phys. Rev. A 6, 2211 (1972).
  • [67] G. Ferrini, D. Spehner, A. Minguzzi, and F. W. J. Hekking, Effect of phase noise on quantum correlations in Bose-Josephson junctions, Phys. Rev. A 84, 043628 (2011).
  • [68] M. Kitagawa and M. Ueda, Squeezed spin states, Phys. Rev. A 47, 5138 (1993).
  • [69] A. S. Sørensen, L.-M. Duan, J. I. Cirac, and P. Zoller, Many-particle entanglement with Bose-Einstein condensates, Nature 409, 63 (2001).
  • [70] M. F. Riedel, P. Böhi, Y. Li, T. W. Hänsch, A. Sinatra, and P. Treutlein, Atom-chip-based generation of entanglement for quantum metrology, Nature 464, 1170 (2010).
  • [71] R. Schmied, J.-D. Bancal, B. Allard, M. Fadel, V. Scarani, P. Treutlein, and N. Sangouard, Bell correlations in a Bose-Einstein condensate, Science 352, 441 (2016).
  • [72] M. Takeuchi, S. Ichihara, T. Takano, M. Kumakura, T. Yabuzaki, and Y. Takahashi, Spin Squeezing Via One-Axis Twisting with Coherent Light, Phys. Rev. Lett. 94, 023003 (2005).
  • [73] I. D. Leroux, M. H. Schleier-Smith, and V. Vuletić, Implementation of Cavity Squeezing of a Collective Atomic Spin, Phys. Rev. Lett. 104, 073602 (2010).
  • [74] M. H. Schleier-Smith, I. D. Leroux, and V. Vuletić, Squeezing the collective spin of a dilute atomic ensemble by cavity feedback, Phys. Rev. A 81, 021804 (2010).
  • [75] J. G. Bohnet, K. C. Cox, M. A. Norcia, J. M. Weiner, Z. Chen, and J. K. Thompson, Reduced spin measurement back-action for a phase sensitivity ten times beyond the standard quantum limit, Nat. Photonics 8, 731 (2014).
  • [76] O. Hosten, N. J. Engelsen, R. Krishnakumar, and M. A. Kasevich, Measurement noise 100 times lower than the quantum projection limit using entangled atoms, Nature 529, 505 (2016).
  • [77] A. Micheli, D. Jaksch, J. I. Cirac, and P. Zoller, Many-particle entanglement in two-component Bose-Einstein condensates, Phys. Rev. A 67, 013607 (2003).
  • [78] H. Strobel, W. Muessel, D. Linnemann, T. Zibold, D. B. Hume, L. Pezzè, A. Smerzi, and M. K. Oberthaler, Fisher information and entanglement of non-Gaussian spin states, Science 345, 424 (2014).
  • [79] W. Muessel, H. Strobel, D. Linnemann, T. Zibold, B. Juliá-Diaz, and M. K. Oberthaler, Twist-and-turn spin squeezing in Bose-Einstein condensates, Phys. Rev. A 92, 023603 (2015).
  • [80] M. Gärrtner, P. Hauke, and A. M. Rey, Relating Out-of-Time-Order Correlations to Entanglement via Multiple-Quantum Coherences, Phys. Rev. Lett. 120, 040402 (2018).
  • [81] D. Girolami and B. Yadin, Witnessing Multipartite Entanglement by Detecting Asymmetry, Entropy 19, 124 (2017).
  • [82] B. Julià-Díaz, E. Torrontegui, J. Martorell, J. G. Muga, and A. Polls, Fast generation of spin-squeezed states in bosonic Josephson junctions, Phys. Rev. A 86, 063623 (2012).
  • [83] B. Julià-Díaz, T. Zibold, M.K. Oberthaler, M. Mele-Messeguer, J. Martorell, and A. Polls, Dynamic generation of spin- squeezed states in bosonic Josephson junctions, Phys. Rev. A 86, 023615 (2012).
  • [84] V. S. Shchesnovich and M. Trippenbach, Fock-space WKB method for the boson Josephson model describing a Bose-Einstein condensate trapped in a double-well potential, Phys. Rev. A 78, 023611 (2008).
  • [85] A. Lerose, J. Marino, B. Zunkovic, A. Gambassi, and A. Silva, Chaotic Dynamical Ferromagnetic Phase Induced by Nonequilibrium Quantum Fluctuations, Phys. Rev. Lett. 120, 130603 (2018); A. Lerose, B. Zunkovic, J. Marino, A. Gambassi, and A. Silva, Impact of nonequilibrium fluctuations on prethermal dynamical phase transitions in long-range interacting spin chains, Phys. Rev. B 99, 045128 (2019).
  • [86] The QFI is given by the maximum between Eq. (11) and
    FQ(t)N=N(1cos2N2ctN)(N1)αQ(t)2.\frac{F_{Q}(t)}{N}=N\left(1-\cos^{2N-2}\frac{ct}{N}\right)-(N-1)\frac{\alpha_{Q}(t)}{2}. (44)
    Equation (44) is larger than Eq. (11) for ctNπ/22Nct\simeq N\pi/2-2\sqrt{N} and even values of NN.
  • [87] A. Lerose and S. Pappalardi, Origin of the slow growth of entanglement entropy in long-range interacting spin systems, Phys. Rev. Research 2, 012041(R) (2020).
  • [88] A. Lerose and S. Pappalardi, private communication. S. Pappalardi, Entanglement dynamics and chaos in long-range quantum systems PhD thesis (2020). Available at https://iris.sissa.it.
  • [89] C. K. Law, H. T. Ng, and P. T. Leung, Coherent control of spin squeezing, Phys. Rev. A 63, 055601 (2001).
  • [90] J. Liu, W. Wang, C. Zhang, Q. Niu, and B. Li, Fidelity for the quantum evolution of a Bose-Einstein condensate, Phys. Rev. A 72, 063623 (2005).
  • [91] V. V. Ulyanov and O. B. Zaslavskii, New methods in the theory of quantum spin systems, Phys. Rep. 216, 179 (1992).
  • [92] J. Vidal, G. Palacios and R. Mosseri, Entanglement in a second-order quantum phase transition, Phys. Rev. A 69, 022107 (2004).