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

Evaluation of metrology using self-consistent approach in the MCTDH theory

Jae-Gyun Baak    Uwe R. Fischer Seoul National University, Department of Physics and Astronomy,
Center for Theoretical Physics, Seoul 08826, Korea

Self-consistent many-body metrology

Jae-Gyun Baak    Uwe R. Fischer Seoul National University, Department of Physics and Astronomy,
Center for Theoretical Physics, Seoul 08826, Korea
Abstract

We investigate performing classical and quantum metrology and parameter estimation by using interacting trapped bosons, which we theoretically treat by a self-consistent many-body approach of the multiconfigurational Hartree type. Focusing on a tilted double-well geometry, we compare a self-consistently determined and monitored two-mode truncation, with dynamically changing orbitals, to the conventional two-mode approach of fixed orbitals, where only Fock space coefficients evolve in time. We demonstrate that, as a consequence, various metrological quantities associated to a concrete measurement such as the classical Fisher information and the maximum likelihood estimator are deeply affected by the orbitals’ change during the quantum evolution. Self-consistency of the quantum many-body dynamics of interacting trapped ultracold gases thus fundamentally affects the attainable parameter estimation accuracy of a given metrological protocol.

Within the currently emerging quantum era, quantum metrology [1, 2, 3, 4, 5, 6, 7, 8] has proven itself to be a powerful tool for the accurate estimation of even very small physical parameters, such as gravitational wave amplitudes [9], or to limit the attainable measurement accuracy of fundamental constants like the speed of light [10]. As a result, quantum metrology promises to revolutionize the existing technologies of measurement.

While quantum metrology has frequently been employed in the quantum optical context [11, 12, 13, 14], more recently the corresponding experiments and theory are also exploring coherent matter waves cf., e.g., Refs. [15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 30]. Photons freely propagating in the quantum vacuum are to a very good approximation noninteracting particles and are well described by plane waves of definite momentum. Matter waves forming Bose-Einstein condensates at very low temperatures are, however, interacting by the scattering of their elementary atomic or molecular constituents, and are spatially confined (trapped) by arbitrary scalar potentials. In what follows, we show that the full self-consistency of the quantum many-body evolution of such a system needs in general to be taken into account, to yield reliable parameter estimation. We demonstrate that the interplay of Fock space amplitudes and time-dependent field operator modes (\coloneqq orbitals), representing self-consistent many-body evolution, is crucial. This interplay is not obtained when fixing the orbitals’ shape, thereby significantly restricting the associated Hilbert space.

We take as an archetypical model system and for concreteness a tilted double well, where the parameter to be estimated is the linear slope p4p_{4} which could, e.g., represent exposing the gas to constant gravitational acceleration (see Fig. 1). To facilitate comparison with conventional interferometry, we operate within a (continuously monitored) two-mode approximation (TMA), corresponding to two interferometric arms. We consider a simple (Mach-Zehnder type) experiment which counts the final number of particles on the left and right. It is demonstrated that while a non-self-consistent evolution yields a null result for p4p_{4} [zero classical Fisher information (CFI)], a self-consistent quantum many-body evolution gives finite CFI, enabling p4p_{4} estimation. We therefore conclusively show that self-consistency is crucial for the correct interpretation of parameter estimation data in a trapped interacting many-body system. We note that Heisenberg-limit number scaling of estimation precision with 1/N1/N [instead of the shot noise (standard quantum) limit 1/N\propto 1/\sqrt{N}], for cat state distributions in Fock space [28], or an enhanced NN-scaling relying on kk-body interactions [31, 32], are not our aim in the present work. The focus is on the fundamental imprint of the self-consistency of many-body evolution on the accuracy of parameter estimation from the metrological protocol.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 1: Top: Trap potential. Left: t<0t<0, symmetric double well. Right: At t=0t=0, the tilt is switched on. Bottom: Initially localized orbitals in the large-barrier double-well self-consistently evolve in time after tilting, and become delocalized, whereas without self-consistency they remain localized. The orbitals are vertically offset for clarity.

To determine the many-body evolution self-consistently, we perform a multiconfigurational time-dependent Hartree (MCTDH) analysis. This is a well tested method to describe the dynamics of spatially confined quantum systems, such as the one under investigation here, for which self-consistency is crucial [33, 34, 35, 36, 37]. The general NN-body state reads

|Ψ(t)=nCn(t)|n(t),|\Psi(t)\rangle=\sum\limits_{\vec{n}}C_{\vec{n}}(t)|\vec{n}(t)\rangle\,,\vspace*{-0.25em} (1)

where n|Cn|2=1\sum_{\vec{n}}|C_{\vec{n}}|^{2}=1 for state normalization and n\vec{n} denotes the set of occupation numbers {ni|i=1,2,,M}\{n_{i}\,|\,i=1,2,\cdots,M\} in each mode (orbital), with i=1Mni=N\sum_{i=1}^{M}n_{i}=N. The time-dependent Fock basis state |n(t)|\vec{n}(t)\rangle indicates that the orbitals change in time as a result of finite MM (see, Fig. 1). Their dynamics follows the system of nonlinear coupled integrodifferential Eqs. (S1) in the supplement [38]. Numerical solution enables the determination of Fock space coefficients CnC_{\vec{n}} and orbitals in a self-consistent manner at any time [39].

The quantum metrological approach to parameter estimation, see for example Refs. [40, 41, 42, 43, 44], proceeds essentially as follows. An initial state |ψ|\psi\rangle experiences a dynamical evolution, e.g., eiH^Xte^{-i\hat{H}_{X}t}, during the time tt and the final state |ψX|\psi_{X}\rangle contains the information of the parameter XX. One chooses an appropriate measurement on |ψX|\psi_{X}\rangle to estimate XX. Previous studies on quantum metrology with ultracold atoms [15, 16, 42, 17, 18, 19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 30] have focused on the coefficients Cn(X;t)C_{\vec{n}}(X;t) in Eq. (1), and have calculated the quantum Fisher information (QFI) 𝔉X{\mathfrak{F}}_{X} from Cn(X;t)C_{\vec{n}}(X;t) only. However, since the orbitals also evolve by Eqs. (S1) and the time evolution relies on XX, the |n(X;t)|\vec{n}(X;t)\rangle must be considered in the calculation of both the QFI and the CFI. To evaluate the sensitivity of a quantum mechanical state to a parameter change thus requires full exploitation of the information encoded in the state. Here, we make full use of the parameter dependence of the state, reflected in both coefficients and orbitals. As a result, we establish numerically exact many-body parameter estimation for trapped interacting quantum gases.

A scalar bosonic gas with contact interactions, trapped in a quasi-one-dimensional (quasi-1D) double-well potential, is described by the Hamiltonian

H^=j=1N{122xj2+V(xj)}+gj<kNδ(xjxk),\hat{H}=\sum_{j=1}^{N}\Big{\{}-\frac{1}{2}\frac{\partial^{2}}{\partial x_{j}^{2}}+V(x_{j})\Big{\}}+g\sum^{N}_{j<k}\delta(x_{j}-x_{k}), (2)

where VV is the trap potential. We render H^\hat{H} in a dimensionless form, fixing a unit length LL; the unit of time is then L2L^{2} for =m=1\hbar=m=1 111In Rb87{}^{87}\text{Rb}, L=1μL=1\,\,\mum yields a time unit Δt=1.366\Delta t=1.366  msec.. The quasi-1D interaction coupling gg is controllable by either Feshbach or geometric scattering resonances [46]. Exact solutions of the Schrödinger equation associated to Eq. (2) exist for homogeneous gases under periodic boundary conditions or in a box trap; their metrology was explored in [47].

We assume a trap potential of the form V(x)=Vdw(x)+p4x=12p1x2+p2exp[x2/(2p32)]+p4xV(x)=V_{\rm dw}(x)+p_{4}x=\frac{1}{2}p_{1}x^{2}+p_{2}\exp[-x^{2}/(2\,p_{3}^{2})]+p_{4}\,x, where p4=0p_{4}=0 initially, cf. Fig. 1. The remaining parameters, p1=0.5p_{1}=0.5, p2=50p_{2}=50, and p3=1p_{3}=1, are fixed throughout the evolution. The two lowest-energy single-particle states are symmetric and antisymmetric with respect to the origin, respectively, and their addition and subtraction results in two well-localized orbitals: left ϕL(x)\phi_{\rm L}(x) and right ϕR(x)\phi_{\rm R}(x), see Fig. 1. These orbitals, approximate ground states of each well, furnish the two modes [48].

Refer to caption
Refer to caption
Figure 2: Monitoring the two-mode truncation after turning on p4p_{4}, verifying whether ρtm=(ρ1+ρ2)/N1\rho_{\rm tm}=(\rho_{1}+\rho_{2})/N\lesssim 1 (we put N=10N=10), for cat state (left) and spin-coherent state (right).

The estimation process of p4p_{4} proceeds with an initial state in the form of Eq. (1). With ϕL(x)\phi_{\rm L}(x) and ϕR(x)\phi_{\rm R}(x), we employ two coefficient distributions: A NOON (cat) state, |ψ0=(|N,0+|0,N)/2|\psi_{0}\rangle=(|N,0\rangle+|0,N\rangle)/\sqrt{2}, and a spin-coherent state |ψ0=k=0NN!k!(Nk)!cosNk(π4)sink(π4)|Nk,k|\psi_{0}\rangle=\sum_{k=0}^{N}\sqrt{\frac{N!}{k!\,(N-k)!}}\cos^{N-k}(\frac{\pi}{4})\sin^{k}(\frac{\pi}{4})|N-k,k\rangle. Then a small p4p_{4} tilt is switched on. The state then evolves according to Eqs. (S1); finally, the number of particles in each well is measured and p4p_{4} is estimated from the set of outcomes.

As the relative interaction strength gNgN (typical ratio of interaction over single-particle energies), increases, more modes than two are required to correctly reproduce the many-body dynamics [37]. Keeping gNgN fixed when taking the limit NN\rightarrow\infty has been demonstrated to reproduce the Gross-Pitaevskii ground state energy [49, 50]. We thus keep in the following gNgN fixed when varying NN, to remain close to the TMA. In order to adequately compare self-consistent (SC) evolution to conventional SU(2) two-mode interferometry (TMI), which operates with changing Fock space coefficients only, we maintain TMA validity throughout the time evolution. This can be assessed by evaluating ρtm(ρ1+ρ2)/N\rho_{\rm tm}\coloneqq(\rho_{1}+\rho_{2})/N, cf. Fig. 2, where ρj\rho_{j} is the jjth largest eigenvalue of the reduced one-body density matrix, ρ(1)(x,x;t)\rho^{(1)}(x,x^{\prime};t); see (S2) in the supplement [38]. After diagonalization, ρ(1)(x,x)=jρj(t)ϕj(no)(x,t)ϕj(no)(x,t)\rho^{(1)}(x,x^{\prime})=\sum_{j}\rho_{j}(t)\,\phi_{j}^{({\rm no})\ast}(x^{\prime},t)\,\phi_{j}^{({\rm no})}(x,t), with the natural orbitals {ϕj(no)(x,t)|j=1,2,}\{\phi_{j}^{({\rm no})}(x,t)|\,j=1,2,\cdots\}. When a single ρ1=O(N)\rho_{1}=O(N) (in the formal limit NN\rightarrow\infty), a Bose-Einstein condensate is obtained. For several ρj\rho_{j} of O(N)O(N), we have a fragmented condensate [51]. The validity of the TMA (two-fold fragmented condensate) depends on whether ρ1ρ2N/2\rho_{1}\simeq\rho_{2}\simeq N/2 and ρtmO(1)\rho_{\rm tm}\simeq O(1) hold. In the supplement [38], we provide further evidence for the appropriateness of using the TMA, by demonstrating the rapid convergence of our results with increasing MM.

We define an initial state using four orbitals (M=4M=4), which are including ϕL(x)\phi_{\rm L}(x) and ϕR(x)\phi_{\rm R}(x), and then monitor the natural occupations ρj\rho_{j} under self-consistent evolution at nonzero p4p_{4}, for both cat and spin-coherent states. Fig. 2 shows ρtm\rho_{\rm tm}, where we observe it is close to unity. Also, both ρ1\rho_{1} and ρ2\rho_{2} are macroscopically occupied during the evolution, with negligible occupations ρ3\rho_{3} and ρ4\rho_{4}. This however also depends on the parameter regime used. When gN=0.1gN=0.1, two modes are sufficient, but when gN=1gN=1, ρtm\rho_{\rm tm} discernibly dips below unity. Increasing p4p_{4} further, the TMA fails. An appropriate regime of parameters where ρtm1\rho_{\rm tm}\simeq 1 is obtained by fixing gN=0.1gN=0.1 and p4=0.1p_{4}=0.1. Also, even though natural orbitals are used in the discussion above, we can apply the two-mode criterion to the left/right orbitals or their time-evolved forms, i.e., ϕ1(x,t)\phi_{1}(x,t) and ϕ2(x,t)\phi_{2}(x,t); there always exists a unitary transformation such that ϕj(x,t)=jkUjkϕk(no)(x,t)\phi_{j}(x,t)=\sum_{jk}U_{jk}\,\phi_{k}^{({\rm no})}(x,t).

Expanding the second-quantized form of Eq. (2) with Ψ^(x)=b^LϕL(x)+b^RϕR(x)\hat{\Psi}(x)=\hat{b}_{\rm L}\,\phi_{\rm L}(x)+\hat{b}_{\rm R}\,\phi_{\rm R}(x), a two-site single-band Bose-Hubbard Hamiltonian is obtained. In terms of the usual SU(2) Pauli matrices J^x=12(b^Lb^R+b^Rb^L)\hat{J}_{x}=\frac{1}{2}(\hat{b}_{\rm L}\hat{b}_{\rm R}+\hat{b}^{\dagger}_{\rm R}\hat{b}_{\rm L}) and J^z=12(b^Lb^Lb^Rb^R)\hat{J}_{z}=\frac{1}{2}(\hat{b}_{\rm L}\hat{b}_{\rm L}-\hat{b}^{\dagger}_{\rm R}\hat{b}_{\rm R}),

H^=τJ^x+ϵJ^z+UJ^z2,\hat{H}=-\tau\hat{J}_{x}+\epsilon\,\hat{J}_{z}+U\hat{J}_{z}^{2}\,, (3)

where τ\tau is tunneling amplitude, ϵ\epsilon an energy offset between wells, and UgU\propto g an interaction coupling, all depending on integrals involving the two orbitals. The implementation of quantum metrological protocols using the above Hamiltonian was carried out, e.g., in Refs. [19, 24, 44]: An initial state |ψ0|\psi_{0}\rangle, as defined by the distribution of coefficients CnC_{\vec{n}}, evolves as exp(iH^t)|ψ0\exp(-i\hat{H}t)|\psi_{0}\rangle, and the parameter of interest, e.g., ϵ\epsilon, is estimated from the population imbalance [19, 24]. In our setup, the strong barrier renders the initial τ\tau exponentially small compared to ϵ\epsilon and UU, and ϵ0\epsilon\neq 0, as the symmetry of V(x)V(x) is broken by p4p_{4}. Hence the TMI time evolution operator is, to very good accuracy, exp(i(ϵJ^z+UJ^z2)t)\exp(-i(\epsilon\hat{J}_{z}+U\hat{J}_{z}^{2})t), and the QFI can be analytically calculated by 𝔉ϵ=4ψ0|(ΔJ^z)2|ψ0\mathfrak{F}_{\epsilon}=4\,\langle\psi_{0}|(\Delta\hat{J}_{z})^{2}|\psi_{0}\rangle. When the cat state is used, 𝔉ϵ=N2t2\mathfrak{F}_{\epsilon}=N^{2}\,t^{2}, which is denoted as the Heisenberg limit. For the spin-coherent state, 𝔉ϵ=Nsin2(θ)t2\mathfrak{F}_{\epsilon}=N\sin^{2}(\theta)\,t^{2}, representing the shot-noise (standard quantum) limit. Any nonzero τ\tau deteriorates the NN-scaling of 𝔉ϵ\mathfrak{F}_{\epsilon}, confirmed by numerically calculating the QFI [19]. Note that here only the change of Fock space coefficients has been considered, while the orbitals are fixed in TMI. Because of the latter fact, the still exponentially small τ\tau and UU are kept constant during the evolution, and ϵ\epsilon is abruptly switched on at t=0t=0. In our setting, ϵ(0.7,0.6)\epsilon\in(-0.7,-0.6), and U(0.002,0.03)U\in(0.002,0.03), with concrete values determined by gNgN and NN, on which in turn the initial orbitals ϕL(x)\phi_{\rm L}(x) and ϕR(x)\phi_{\rm R}(x) depend. Recall that our target parameter is p4p_{4}, not ϵ\epsilon, thus by using the chain rule, 𝔉p4=𝔉ϵ×(h1h2)2\mathfrak{F}_{p_{4}}=\mathfrak{F}_{\epsilon}\times(h_{1}-h_{2})^{2}, where the single-particle energies hi𝑑xϕi(x)[122x2+V(x)]ϕi(x)h_{i}\coloneqq\int dx\,\phi_{i}^{\ast}(x)\left[-\frac{1}{2}\frac{\partial^{2}}{\partial x^{2}}+V(x)\right]\phi_{i}(x).

First, we compare the QFIs of the SC approach and the conventional TMI. The QFI with respect to a given parameter XX inscribed onto a pure state |ψX|\psi_{X}\rangle, is 𝔉X=4(XψX|XψX|ψX|XψX|2)\mathfrak{F}_{X}=4\big{(}\langle\partial_{X}\psi_{X}|\partial_{X}\psi_{X}\rangle-|\langle\psi_{X}|\partial_{X}\psi_{X}\rangle|^{2}\big{)}, and insertion of Eq. (1) into |ψX|\psi_{X}\rangle gives Eq. (S4), which facilitates calculation of the QFI using the ingredients of MCTDH from the expansion in Eq. (1), also cf. [38]. The first row in Fig. 3 shows QFI versus time tt and particle number NN, respectively. For each initial state, the SC method reproduces very well the QFI predicted by TMI. The influence of self-consistency thus plays a subdominant role for 𝔉X\mathfrak{F}_{X}. The latter completely depends on the final state, and self-consistency (changing orbitals) essentially represents fitting that state more exactly. When the bosons weakly interact (gN=0.1gN=0.1) and the disturbance to the system is small (p4=0.1p_{4}=0.1), conventional TMI therefore approximates well the QFI [52].

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 3: First and second row display quantum (𝔉\mathfrak{F}) and classical (F) Fisher information, respectively, plotted versus time tt (left) and particle number NN (right), for cat and coherent (coh) states, respectively. At the maximum on the lower left plot, which is at t=1.77t=1.77, the tunneling amplitude in the SC evolution has increased to τ0.09\tau\simeq 0.09 (for N=10N=10).

We now turn to the CFI [53, 54], associated to a concrete measurement, for which, as we show, the impact of self-consistency becomes manifest. Counting the number of bosons in each well constitutes our measurement, so that the CFI is defined as

Fp4=𝐧P(𝐧|p4)(logP(𝐧|p4))p4)2,F_{p_{4}}=\sum_{\vec{\mathbf{n}}}P(\vec{\mathbf{n}}|p_{4})\Big{(}\frac{\partial\log P(\vec{\mathbf{n}}|p_{4}))}{\partial\,p_{4}}\Big{)}^{2}\,, (4)

where P(𝐧|p4)P(\vec{\mathbf{n}}|p_{4}) is the probability distribution (likelihood) 𝐧=(nL,nR)\vec{\mathbf{n}}=(n_{\rm L},n_{\rm R}), given p4p_{4}, and nLn_{\rm L} and nRn_{\rm R} are the numbers of particles that reside in the left and the right well, respectively. An appropriate P(𝐧|p4)P(\vec{\mathbf{n}}|p_{4}) has to be constructed to calculate the CFI and the conventional TMI approach considers that P(𝐧|p4)|n(t)|Ψ(t)|2P(\vec{\mathbf{n}}|p_{4})\coloneqq|\langle\vec{n}(t)|\Psi(t)\rangle|^{2}. Then, the CFI always exactly vanishes, irrespective of the initial state, as the Hamiltonian contains only J^z\hat{J}_{z} and J^z2\hat{J}_{z}^{2}. TMI time evolution therefore just changes the phases of the coefficients: Ck(t)=exp(iϵN2k2t)exp(iU(N2k2)2t)Ck(0)C_{k}(t)=\exp(-i\epsilon\frac{N-2k}{2}t)\,\exp(-iU(\frac{N-2k}{2})^{2}t)\,C_{k}(0), where the state |ψ(t)=k=0NCk(t)|Nk,k=exp(i(ϵJ^z+UJ^z2)t)|ψ(0)|\psi(t)\rangle=\sum_{k=0}^{N}C_{k}(t)|N-k,k\rangle=\exp(-i(\epsilon\hat{J}_{z}+U\hat{J}_{z}^{2})t)|\psi(0)\rangle and J^z|Nk,k=N2k2|Nk,k\hat{J}_{z}|N-k,k\rangle=\frac{N-2k}{2}|N-k,k\rangle. Then P(𝐧=(Nk,k)|p4)=|Ck(t)|2=|Ck(0)|2P(\vec{\mathbf{n}}=(N-k,k)|p_{4})=|C_{k}(t)|^{2}=|C_{k}(0)|^{2}, independent of p4p_{4}, yielding vanishing CFI, cf. [38].

On the other hand, when both orbitals and Fock space coefficients evolve in the SC framework, the initial interpretation of the orbitals cannot be maintained throughout the time evolution. Initially well-localized orbitals, i.e., ϕL(x)=ϕ1(x,0)\phi_{\rm L}(x)=\phi_{1}(x,0) and ϕR(x)=ϕ2(x,0)\phi_{\rm R}(x)=\phi_{2}(x,0), correspond to particles being found in the left and right well, respectively. However, the orbitals change with time during the many-body evolution, so ϕ1(x,t)\phi_{1}(x,t) and ϕ2(x,t)\phi_{2}(x,t), do not necessarily imply left or right localization at the time of measurement. Therefore, simply computing P(𝐧|p4)=|n(t)|Ψ(t)|2P(\vec{\mathbf{n}}|p_{4})=|\langle\vec{n}(t)|\Psi(t)\rangle|^{2}, as in the TMI approach, is not applicable. In the bosonic field operator Ψ^(x)=jb^j(t)ϕj(x,t)\hat{\Psi}(x)=\sum_{j}\hat{b}_{j}(t)\,\phi_{j}(x,t), it is clear that the bosonic annihilation operator b^j(t)\hat{b}_{j}(t) corresponds to the time-evolving orbital ϕj(x,t)\phi_{j}(x,t), which delocalizes with increasing tt, see Fig. 1. Thus the Fock state in Eq. (1) |n(t)=(b^1(t))n1(b^2(t))n2(b^M(t))nMn1!n2!nM!|0|\vec{n}(t)\rangle=\frac{\big{(}\hat{b}_{1}^{\dagger}(t)\big{)}^{n_{1}}\big{(}\hat{b}_{2}^{\dagger}(t)\big{)}^{n_{2}}\cdots\big{(}\hat{b}_{M}^{\dagger}(t)\big{)}^{n_{M}}}{\sqrt{n_{1}!n_{2}!\cdots n_{M}!}}|0\rangle cannot by itself project the quantum state into any of |𝐧|\vec{\mathbf{n}}\rangle and n(t)|Ψ(t)\langle\vec{n}(t)|\Psi(t)\rangle cannot be interpreted as probability amplitude for each measurement outcome as in TMI. In other words, |n(t)|\vec{n}(t)\rangle and |𝐧|\vec{\mathbf{n}}\rangle, while initially identical, become different due to SC evolution. Hence we need to re-establish a connection between time-evolving orbitals and |𝐧|\vec{\mathbf{n}}\rangle that corresponds to a measurement outcome, resulting in the proper distribution P(𝐧|p4)=|𝐧|Ψ(t)|2|n(t)|Ψ(t)|2P(\vec{\mathbf{n}}|p_{4})=|\langle\vec{\mathbf{n}}|\Psi(t)\rangle|^{2}\neq|\langle\vec{n}(t)|\Psi(t)\rangle|^{2}.

Within SC time evolution after a trap tilt, ϕ1(x,t)\phi_{1}(x,t) and ϕ2(x,t)\phi_{2}(x,t) remain well-localized in the case of a cat state. That is, ϕ1(x,t)\phi_{1}(x,t) (ϕ2(x,t)\phi_{2}(x,t)) begins from ϕ1(x,0)=ϕL(x)\phi_{1}(x,0)=\phi_{\rm L}(x) [ϕ2(x,0)=ϕR(x)\phi_{2}(x,0)=\phi_{\rm R}(x)] and their absolute value remains nearly identical except slightly wider (narrower) width and shorter (taller) height, respectively. For the spin-coherent state, however, ϕ1(x,t)\phi_{1}(x,t) and ϕ2(x,t)\phi_{2}(x,t) spread into opposite wells while they evolve under nonzero p4p_{4}; see Fig. 1. Then even when a particle resides in ϕ1(x,t)\phi_{1}(x,t) or ϕ2(x,t)\phi_{2}(x,t), to assign it to the left or right well is ambiguous. One can however still define mathematically “left” or “right” by integrating the orbitals from -\infty to the center point x=0x=0 of V(x)V(x) and from the center point of V(x)V(x) to \infty, respectively: PLeft=0|ϕj(x,t)|2𝑑xP_{\text{Left}}=\int_{-\infty}^{0}\!\!|\phi_{j}(x,t)|^{2}\,dx and PRight=0|ϕj(x,t)|2𝑑xP_{\text{Right}}=\int_{0}^{\infty}\!\!|\phi_{j}(x,t)|^{2}\,dx. One can then construct P(𝐧|p4)P(\vec{\mathbf{n}}|p_{4}) by computing the permanent of a special matrix composed from PLeftP_{\text{Left}} and PRightP_{\text{Right}}. The computable NN range is limited, though, due to rapidly increasing algorithmic complexity when calculating a matrix permanent, see [38].

The second row in Fig. 3 shows the CFI versus tt and NN. Fixed orbitals results in vanishing CFI, as expected. Also, even within SC evolution, the cat state shows almost vanishing CFI, which is attributed to the fact that the orbitals stay localized in each well during the whole evolution time and the probabilities, i.e., PLeftP_{\text{Left}} and PRightP_{\text{Right}}, remain nearly constant (for small p4p_{4}). Thus under the given measurement the change of P(𝐧|p4)P(\vec{\mathbf{n}}|p_{4}) with respect to p4p_{4} is negligible. However, the spin-coherent state displays a significant change in orbitals and increasing CFI during the early stage. Bottom right in Fig. 3 shows the NN-scaling of the CFI; the SC approach with spin-coherent state shows an almost linearly increasing CFI. The complex fluctuation pattern appears because of the short time t=1.77t=1.77 after a nonzero p4p_{4} is suddenly applied, and for increasing tt these fluctuations smoothen out. Thus a SC approach may yield drastically different metrological predictions from a TMI based method.

Refer to caption
Figure 4: Single implementation of the MLE for an estimate of p4p_{4} using the spin-coherent state. Red solid line is for the SC approach and orange dashed line (a constant p4\forall\,p_{4}) for the TMI. In the inset, red squares represent the mean-square deviation and black rounds the Cramér-Rao lower bound from Eq. (5).

Fig. 4 shows our primary result: The implementation of parameter estimation at the final stage of the metrological protocol. The maximum likelihood estimator (MLE) [6, 55] is used as a concrete example, because it asymptotically saturates the Cramér-Rao bound for an infinite number of measurements (ν\nu\rightarrow\infty, see also [38])

(δXest)2X1νFX.\langle(\delta X_{\text{est}})^{2}\rangle_{X}\geq\frac{1}{\nu F_{X}}\,. (5)

Here, the parameter XX is our p4p_{4}, XestX_{\text{est}} is an estimator of XX, and δXest:=Xest/|XestX/X|X\delta X_{\text{est}}:=X_{\text{est}}/\big{|}\partial\langle X_{\text{est}}\rangle_{X}/\partial X\big{|}-X, with average X\langle\cdots\rangle_{X} taken with respect to P(𝐧|p4)P(\vec{\mathbf{n}}|p_{4}).

In Fig. 4, the likelihood function P(𝐧|p4)P(\vec{\mathbf{n}}|p_{4}) is displayed, supposing, for concreteness, that the measurement outcome is 𝐧=(7,3)\vec{\mathbf{n}}=(7,3), where nLn_{\rm L} and nRn_{\rm R} denote number of particles in left and right well, respectively. Red solid line shows the maximum of P(𝐧|p4)P(\vec{\mathbf{n}}|p_{4}) at p40.109p_{4}\simeq 0.109, thus the estimate of p4p_{4}, given 𝐧=(7,3)\vec{\mathbf{n}}=(7,3), is Xest0.109X_{\text{est}}\simeq 0.109. Similarly, every single outcome, 1111 in total, is connected to an estimate of p4p_{4}. The orange dashed line obtained by conventional TMI with a spin-coherent state stays flat, which means that the information given by the measurement outcome 𝐧=(7,3)\vec{\mathbf{n}}=(7,3) is zero, and an estimate of p4p_{4}, in accordance with Fp4=0F_{p_{4}}=0, is not possible.

When XestX=X\langle X_{\text{est}}\rangle_{X}=X holds, an estimator is unbiased. The present MLE is not unbiased except infinitesimally close to p4=0p_{4}=0. The larger p4p_{4}, the more bias its estimation acquires. For instance, in Fig. 4, the true value of p4p_{4} is 0.10.1 and the bias of the MLE is on the level of 10 %, XestX=0.10.11\langle X_{\text{est}}\rangle_{X=0.1}\simeq 0.11. To compensate for this bias, we calibrate the MLE by obtaining Xest(X)\langle X_{\text{est}}\rangle(X) from our MCTDH simulations, see [38]. Also, XestX/X|X=0.10.6\partial\langle X_{\text{est}}\rangle_{X}/\partial X|_{X=0.1}\simeq 0.6, which then provides the mean-square deviation (δXest)2X=0.1\langle(\delta X_{\text{est}})^{2}\rangle_{X=0.1}. The inset in Fig. 4 verifies that, asymptotically, (δXest)2X=0.1\langle(\delta X_{\text{est}})^{2}\rangle_{X=0.1} approaches the Cramér-Rao lower bound for large ν\nu.

In conclusion, we have found, using a self-consistent many-body approach, that many-body metrology utilizing interacting trapped bosons needs to conform to the final self-consistently computed many-body state. The QFI completely depends on the parameter dependence of the final state itself, and is thus relatively unaffected by self-consistency (in the weakly interacting regime). However, even in the latter regime, the CFI for a parameter estimation experiment is strongly affected by self-consistency due to its sensitive dependence on the orbitals’ time evolution. As a particularly notable example, fitting the outcome of a number-statistics experiment in a double well to conventional TMI gives a null result for estimating the slope parameter p4p_{4}. The SC approach we employ, however, enables p4p_{4} estimation. Metrology with trapped ultracold quantum gases thus in general requires self-consistency of dynamical evolution, to correctly predict the estimation precision that can be accomplished in a given metrological protocol.

This work has been supported by the National Research Foundation of Korea under Grants No. 2017R1A2A2A05001422 and No. 2020R1A2C2008103.

References

I Supplemental Material

I.1 I. Multiconfigurational time-dependent Hartree theory

Given a set of coefficients and a set of orbitals for an initial state, the time evolution in the MCTDH framework proceeds according to the following system of equations:

i𝐂(t)t\displaystyle i\,\frac{\partial\,\mathbf{C}(t)}{\partial\,t} =\displaystyle= 𝐇(t)𝐂(t),\displaystyle\mathbf{H}(t)\,\mathbf{C}(t)\,,
it|ϕj\displaystyle i\,\partial_{t}|\phi_{j}\rangle =\displaystyle= P^[h^|ϕj+k,s,q,l[ρ1]jkρksqlW^sl|ϕq],\displaystyle\hat{P}\Big{[}\hat{h}|\phi_{j}\rangle+\sum_{k,s,q,l}[\rho^{-1}]_{jk}\rho_{ksql}\hat{W}_{sl}|\phi_{q}\rangle\Big{]}, (S1)

which is derived by applying the time-dependent variational principle to the interacting NN-body Hamiltonian H^=j=1Nh^(xj)+j<kW^(xjxk)\hat{H}=\sum_{j=1}^{N}\hat{h}(x_{j})+\sum_{j<k}\hat{W}(x_{j}-x_{k})[34]. Here, 𝐂(t)\mathbf{C}(t) is a column vector that consists of all possible expansion coefficients Cn(t)C_{\vec{n}}(t) and 𝐇(t)\mathbf{H}(t) corresponds to the time-dependent Hamiltonian matrix in the basis {|n(t)}\{|\vec{n}(t)\rangle\}. Also, h^\hat{h} is a single-particle Hamiltonian, W^sl=𝑑xϕs(x)W^(xx)ϕl(x)\hat{W}_{sl}=\int dx^{\prime}\,\phi_{s}^{\ast}(x^{\prime})\,\hat{W}(x-x^{\prime})\,\phi_{l}(x^{\prime}), and P^=1j=1M|ϕjϕj|\hat{P}=1-\sum_{j=1}^{M}|\phi_{j}\rangle\langle\phi_{j}| is an projection operator to the subspace that is orthogonal to the one spanned by orbitals. The [ρ1]jk[\rho^{-1}]_{jk} is a matrix element of the inverse of reduced one-body density matrix:

ρ(x,x;t)\displaystyle\rho(x,x^{\prime};t) =\displaystyle= Ψ(t)|Ψ^(x)Ψ^(x)|Ψ(t)\displaystyle\langle\Psi(t)|\hat{\Psi}^{\dagger}(x^{\prime})\hat{\Psi}(x)|\Psi(t)\rangle (S2)
=\displaystyle= k,qϕk(x,t)ϕq(x,t)Ψ(t)|b^k(t)b^q(t)|Ψ(t)\displaystyle\sum_{k,q}\phi_{k}^{\ast}(x^{\prime},t)\phi_{q}(x,t)\langle\Psi(t)|\hat{b}_{k}^{\dagger}(t)\hat{b}_{q}(t)|\Psi(t)\rangle
=\displaystyle= k,qϕk(x,t)ϕq(x,t)ρkq(t),\displaystyle\sum_{k,q}\phi_{k}^{\ast}(x^{\prime},t)\phi_{q}(x,t)\rho_{kq}(t)\,,

where the ρkq\rho_{kq} is, for the cases of k=qk=q and kqk\neq q,

ρkk=n|Cn(t)|2nk,ρkq=nCn(t)Cnkq(t)nk(nq+1).\rho_{kk}=\sum_{\vec{n}}|C_{\vec{n}}(t)|^{2}n_{k},\quad\rho_{kq}=\sum_{\vec{n}}C_{\vec{n}}^{\ast}(t)C_{\vec{n}_{k}^{q}}(t)\sqrt{n_{k}(n_{q}+1)}\,.

Similarly, ρksql\rho_{ksql} is a matrix element of the reduced two-body matrix

ρ(x1,x2,x1,x2;t)\displaystyle\rho(x_{1},x_{2},x^{\prime}_{1},x^{\prime}_{2};t) =\displaystyle= Ψ(t)|Ψ^(x1)Ψ^(x2)Ψ^(x1)Ψ^(x2)|Ψ(t)\displaystyle\langle\Psi(t)|\hat{\Psi}^{\dagger}(x^{\prime}_{1})\hat{\Psi}^{\dagger}(x^{\prime}_{2})\hat{\Psi}(x_{1})\hat{\Psi}(x_{2})|\Psi(t)\rangle (S3)
=\displaystyle= k,s,q,lϕk(x1,t)ϕs(x2,t)ϕq(x1,t)ϕl(x2,t)ρksql(t),\displaystyle\sum_{k,s,q,l}\phi_{k}^{\ast}(x^{\prime}_{1},t)\phi_{s}^{\ast}(x^{\prime}_{2},t)\phi_{q}(x_{1},t)\phi_{l}(x_{2},t)\rho_{ksql}(t)\,,

where

ρkkkk\displaystyle\rho_{kkkk} =\displaystyle= n|Cn(t)|2nk(nk1),ρksks=n|Cn(t)|2nkns,\displaystyle\sum_{\vec{n}}|C_{\vec{n}}(t)|^{2}n_{k}(n_{k}-1)\,,\quad\rho_{ksks}=\sum_{\vec{n}}|C_{\vec{n}}(t)|^{2}n_{k}n_{s}\,,
ρkkqq\displaystyle\rho_{kkqq} =\displaystyle= nCn(t)Cnkkqq(t)(nk1)nk(nq+1)(nq+2),ρkkkl=nCn(t)Cnkl(t)(nk1)nk(nl+1),\displaystyle\sum_{\vec{n}}C_{\vec{n}}^{\ast}(t)\,C_{\vec{n}_{kk}^{qq}}(t)\sqrt{(n_{k}-1)n_{k}(n_{q}+1)(n_{q}+2)}\,,\quad\rho_{kkkl}=\sum_{\vec{n}}C_{\vec{n}}^{\ast}(t)\,C_{\vec{n}_{k}^{l}}(t)(n_{k}-1)\sqrt{n_{k}(n_{l}+1)}\,,
ρksss\displaystyle\rho_{ksss} =\displaystyle= nCn(t)Cnks(t)nsnk(ns+1),ρkkql=nCn(t)Cnkkql(t)(nk1)nk(nq+1)(nl+1),\displaystyle\sum_{\vec{n}}C_{\vec{n}}^{\ast}(t)\,C_{\vec{n}_{k}^{s}}(t)n_{s}\sqrt{n_{k}(n_{s}+1)}\,,\quad\rho_{kkql}=\sum_{\vec{n}}C_{\vec{n}}^{\ast}(t)\,C_{\vec{n}_{kk}^{ql}}(t)\sqrt{(n_{k}-1)n_{k}(n_{q}+1)(n_{l}+1)}\,,
ρksqq\displaystyle\rho_{ksqq} =\displaystyle= nCn(t)Cnksqq(t)nkns(nq+1)(nq+2),ρkssl=nCn(t)Cnkl(t)nsnk(nl+1),\displaystyle\sum_{\vec{n}}C_{\vec{n}}^{\ast}(t)\,C_{\vec{n}_{ks}^{qq}}(t)\sqrt{n_{k}n_{s}(n_{q}+1)(n_{q}+2)}\,,\quad\rho_{kssl}=\sum_{\vec{n}}C_{\vec{n}}^{\ast}(t)\,C_{\vec{n}_{k}^{l}}(t)n_{s}\sqrt{n_{k}(n_{l}+1)}\,,
ρksql\displaystyle\rho_{ksql} =\displaystyle= nCn(t)Cnksql(t)nkns(nq+1)(nl+1).\displaystyle\sum_{\vec{n}}C_{\vec{n}}^{\ast}(t)\,C_{\vec{n}_{ks}^{ql}}(t)\sqrt{n_{k}n_{s}(n_{q}+1)(n_{l}+1)}\,.

Infinite resources for numerical calculation makes it possible to assume the theoretical limit MM\rightarrow\infty, thus P^0^\hat{P}\rightarrow\hat{0} and t|ϕj=0\partial_{t}|\phi_{j}\rangle=0 in Eq. (S1), which means that a complete set of time-independent orbitals {ϕj(x)|j=1,2,}\{\phi_{j}(x)\,|\,j=1,2,\cdots\} can be composed and the dynamics of systems is fully described only by the set of coefficients {Cn(t)}\{C_{\vec{n}}(t)\}, from which all quantum metrological properties can be extracted.

If M=2M=2, with fixed orbitals, is adequate for the description of a system, the modes comprise the conventional TMI, using a SU(2) formulation. Optical systems have been used to realize such two-mode systems, e.g., a Mach-Zehnder interferometer, where only the Fock space coefficients matter to predict the number of photons in each interferometric arm. However, for interacting atoms, an exact description requires infinite MM, and truncating at finite MM is valid only approximately, cf. the error-controlled extension of multiconfigurational Hartree put forth in [56]. As the interaction becomes weaker, a description in terms of finite MM improves. The self-consistent MCTDH framework here goes significantly further further than a conventional TMI and introduces time-evolving orbitals of changing shape.We also note here that a Hartree-Fock method, using plane waves for the field operator expansion as appropriate in a translationally invariant system, will fail to capture a trapped system when, as necessary for finite computational resources, the expansion is truncated at a finite MM.

I.2 II. Quantum Fisher information of a pure state in the MCTDH framework

Because of the introduction of time-evolving orbitals, a formulation of the QFI is required which facilitates incorporating the result of solving the MCTDH time evolution in Eqs. (S1). The QFI, which is the ultimate limit of precision given by |ψX|\psi_{X}\rangle, is calculated by 𝔉X=4(XψX|XψX|ψX|XψX|2)\mathfrak{F}_{X}=4\big{(}\langle\partial_{X}\psi_{X}|\partial_{X}\psi_{X}\rangle-|\langle\psi_{X}|\partial_{X}\psi_{X}\rangle|^{2}\big{)} for general pure states, and for some state represented as Eq. (1), we have, for any number of modes,

𝔉X/4\displaystyle\mathfrak{F}_{X}/4 =\displaystyle= nXCnXCn|nCnXCn|2\displaystyle\sum_{\vec{n}}\partial_{X}C_{\vec{n}}^{\ast}\,\partial_{X}C_{\vec{n}}-\Big{|}\sum_{\vec{n}}C_{\vec{n}}^{\ast}\,\partial_{X}C_{\vec{n}}\Big{|}^{2} (S4)
+nk,q(XCnCnkqCnXCnkq)(X)kqζqkn(XCnCnCnXCn)k,q(X)kqρkq\displaystyle\,\,+\sum_{\vec{n}}\sum_{k,q}\big{(}\partial_{X}C_{\vec{n}}^{\ast}\,C_{\vec{n}_{k}^{q}}-C_{\vec{n}}^{\ast}\,\partial_{X}C_{\vec{n}_{k}^{q}}\big{)}(\partial_{X})_{kq}\,\zeta_{qk}-\sum_{\vec{n}}\big{(}\partial_{X}C_{\vec{n}}^{\ast}\,C_{\vec{n}}-C_{\vec{n}}^{\ast}\,\partial_{X}C_{\vec{n}}\big{)}\sum_{k,q}(\partial_{X})_{kq}\,\rho_{kq}
k,s,q(X)ks(X)sqρkq+(k,q(X)kqρkq)2k,s,q,l(X)kq(X)slρksql,\displaystyle\quad-\sum_{k,s,q}(\partial_{X})_{ks}(\partial_{X})_{sq}\,\rho_{kq}+\Big{(}\sum_{k,q}(\partial_{X})_{kq}\rho_{kq}\Big{)}^{2}-\!\sum_{k,s,q,l}(\partial_{X})_{kq}(\partial_{X})_{sl}\,\rho_{ksql}\,,

where ζqk:=nk(nq+1)\zeta_{qk}:=\sqrt{n_{k}(n_{q}+1)} or ζqk:=nk\zeta_{qk}:=n_{k} if qkq\neq k or q=kq=k, respectively, and (X)kq:=𝑑xϕk(x,t)Xϕq(x,t)(\partial_{X})_{kq}:=\int dx\,\phi^{\ast}_{k}(x,t)\,\partial_{X}\phi_{q}(x,t). Refer to Eq. (S2) and Eq. (S3) for the definitions of ρkq\rho_{kq} and ρksql\rho_{ksql}. The first two terms involve only the coefficients and the remaining terms are related to the changes of coefficients and orbitals, for infinitesimal increment of XX. In summary, Eq. (S4) completely incorporates the information orbitals as well as coefficients changing with XX.

I.3 III. Construction of the probability distribution (likelihood) in MCTDH for bosons

Here we explain how to construct the probability distribution of measurement outcomes. This process obviously depends on the specific systems and the choice of measurement. Here, the metrological implementation with the ultracold bosons trapped in a double-well potential is covered and the number of particles in each well is counted after the time evolution is finished, and considered as the measurement. The probability of a particle in ϕj(x,t)\phi_{j}(x,t) to be found at the left (LL) or the right (RR) is defined as

PL,j=0|ϕj(x,t)|2𝑑x,PR,j=0|ϕj(x,t)|2𝑑x,P_{L,j}=\int_{-\infty}^{0}\!\!|\phi_{j}(x,t)|^{2}\,dx\,,\qquad P_{R,j}=\int_{0}^{\infty}\!\!|\phi_{j}(x,t)|^{2}\,dx\,, (S5)

where we assume that the center of the 1D potential is at x=0x=0.

Next, we need to consider the combinatorial problem related to many particles and bosonic statistics. Let us take for simplicity the example of N=2N=2. There are three measurement outcomes: 𝐧:=(nL,nR)=(2,0)\vec{\mathbf{n}}:=(n_{\rm L},n_{\rm R})=(2,0), (1,1)(1,1), and (0,2)(0,2), in which nLn_{\rm L} and nRn_{\rm R} mean the numbers of particles found in the left well and in the right well, respectively. When the final state is nCn|n=k=02Ck|2k,k\sum_{\vec{n}}C_{\vec{n}}|\vec{n}\rangle=\sum_{k=0}^{2}C_{k}|2-k,k\rangle, the probability for each case is as follows:

P0\displaystyle P_{0} =\displaystyle= P(𝐧=(2,0))=|C0|2PL,12+|C1|2PL,1PL,2+|C2|2PL,22,\displaystyle P\big{(}\vec{\mathbf{n}}=(2,0)\big{)}=|C_{0}|^{2}P_{{\rm L},1}^{2}+|C_{1}|^{2}P_{{\rm L},1}P_{{\rm L},2}+|C_{2}|^{2}P_{{\rm L},2}^{2}\,,
P1\displaystyle P_{1} =\displaystyle= P(𝐧=(1,1))=2|C0|2PL,1PR,1+|C1|2(PL,1PR,2+PR,1PL,2)+2|C2|2PL,2PR,2,\displaystyle P\big{(}\vec{\mathbf{n}}=(1,1)\big{)}=2\,|C_{0}|^{2}P_{{\rm L},1}P_{{\rm R},1}+|C_{1}|^{2}(P_{{\rm L},1}P_{{\rm R},2}+P_{{\rm R},1}P_{{\rm L},2})+2\,|C_{2}|^{2}P_{{\rm L},2}P_{{\rm R},2}\,,
P2\displaystyle P_{2} =\displaystyle= P(𝐧=(0,2))=|C0|2PR,12+|C1|2PR,1PR,2+|C2|2PR,22,\displaystyle P\big{(}\vec{\mathbf{n}}=(0,2)\big{)}=|C_{0}|^{2}P_{{\rm R},1}^{2}+|C_{1}|^{2}P_{{\rm R},1}P_{{\rm R},2}+|C_{2}|^{2}P_{{\rm R},2}^{2}\,, (S6)

where it is trivial to show that P(𝐧=(2,0))+P(𝐧=(1,1))+P(𝐧=(0,2))=1P(\vec{\mathbf{n}}=(2,0))+P(\vec{\mathbf{n}}=(1,1))+P(\vec{\mathbf{n}}=(0,2))=1 using Pj,L+Pj,R=1P_{j,L}+P_{j,R}=1 and |C0|2+|C1|2+|C2|2=1|C_{0}|^{2}+|C_{1}|^{2}+|C_{2}|^{2}=1. After careful inspection, one can rewrite the above probabilities as

P0\displaystyle P_{0} =\displaystyle= P(𝐧=(2,0))=|C0|22{PL,1PL,1PL,1PL,1}+|C1|22{PL,1PL,2PL,1PL,2}+|C2|22{PL,2PL,2PL,2PL,2},\displaystyle P\big{(}\vec{\mathbf{n}}=(2,0)\big{)}=\frac{|C_{0}|^{2}}{2}\Big{\{}\begin{array}[]{cc}P_{{\rm L},1}&P_{{\rm L},1}\\ P_{{\rm L},1}&P_{{\rm L},1}\end{array}\Big{\}}+\frac{|C_{1}|^{2}}{2}\Big{\{}\begin{array}[]{cc}P_{{\rm L},1}&P_{{\rm L},2}\\ P_{{\rm L},1}&P_{{\rm L},2}\end{array}\Big{\}}+\frac{|C_{2}|^{2}}{2}\Big{\{}\begin{array}[]{cc}P_{{\rm L},2}&P_{{\rm L},2}\\ P_{{\rm L},2}&P_{{\rm L},2}\end{array}\Big{\}}, (S13)
P1\displaystyle P_{1} =\displaystyle= P(𝐧=(1,1))=|C0|2{PL,1PL,1PR,1PR,1}+|C1|2{PL,1PL,2PR,1PR,2}+|C2|2{PL,2PL,2PR,2PR,2},\displaystyle P\big{(}\vec{\mathbf{n}}=(1,1)\big{)}=|C_{0}|^{2}\Big{\{}\begin{array}[]{cc}P_{{\rm L},1}&P_{{\rm L},1}\\ P_{{\rm R},1}&P_{{\rm R},1}\end{array}\Big{\}}+|C_{1}|^{2}\Big{\{}\begin{array}[]{cc}P_{{\rm L},1}&P_{{\rm L},2}\\ P_{{\rm R},1}&P_{{\rm R},2}\end{array}\Big{\}}+|C_{2}|^{2}\Big{\{}\begin{array}[]{cc}P_{{\rm L},2}&P_{{\rm L},2}\\ P_{{\rm R},2}&P_{{\rm R},2}\end{array}\Big{\}}, (S20)
P2\displaystyle P_{2} =\displaystyle= P(𝐧=(0,2))=|C0|22{PR,1PR,1PR,1PR,1}+|C1|22{PR,1PR,2PR,1PR,2}+|C2|22{PR,2PR,2PR,2PR,2},\displaystyle P\big{(}\vec{\mathbf{n}}=(0,2)\big{)}=\frac{|C_{0}|^{2}}{2}\Big{\{}\begin{array}[]{cc}P_{{\rm R},1}&P_{{\rm R},1}\\ P_{{\rm R},1}&P_{{\rm R},1}\end{array}\Big{\}}+\frac{|C_{1}|^{2}}{2}\Big{\{}\begin{array}[]{cc}P_{{\rm R},1}&P_{{\rm R},2}\\ P_{{\rm R},1}&P_{{\rm R},2}\end{array}\Big{\}}+\frac{|C_{2}|^{2}}{2}\Big{\{}\begin{array}[]{cc}P_{{\rm R},2}&P_{{\rm R},2}\\ P_{{\rm R},2}&P_{{\rm R},2}\end{array}\Big{\}}, (S27)

in which {V}\{V\} means the permanent of a matrix VV. By tracing the factor in front of each term and by considering bosonic statistics, one can find a regular pattern and generalize as follows:

Pj=P(𝐧=(Nj,j))=1N!(Nj)k=0N|Ck|2{Vj,k},P_{j}=P\big{(}\vec{\mathbf{n}}=(N-j,j)\big{)}=\frac{1}{N!}\Big{(}\begin{array}[]{c}N\\ j\end{array}\Big{)}\sum_{k=0}^{N}|C_{k}|^{2}\{V_{j,k}\}, (S28)

where Vj,kV_{j,k} is a special N×NN\times N matrix, defined as below. The jj is the number of particles in the right well, i.e, 𝐧=(Nj,j)\vec{\mathbf{n}}=(N-j,j) and the kk means n=(Nk,k)\vec{n}=(N-k,k). The example above shows how to compose the matrix Vj,kV_{j,k}. In order to compose V1,2V_{1,2}, for example, “11” is represented as {L,R}\{L,R\} and the “22” is represented as {2,2}\{2,2\}. The former is an ordered set of NjN-j of LL and jj of RR, and the latter is a conversion of “how many particles there are in each mode” into an (ascending-)ordered set of the occupied mode numbers:

j=0:(2,0){L,L},j=1:(1,1){L,R},j=2:(0,2){R,R},\displaystyle j=0\,:\,(2,0)\quad\rightarrow\quad\{L,L\}\,,\qquad j=1\,:\,(1,1)\quad\rightarrow\quad\{L,R\}\,,\qquad j=2\,:\,(0,2)\quad\rightarrow\quad\{R,R\}\,,
k=0:(2,0){1,1},k=1:(1,1){1,2},k=2:(0,2){2,2}.\displaystyle k=0\,:\,(2,0)\quad\rightarrow\quad\{1,1\}\,,\qquad k=1\,:\,(1,1)\quad\rightarrow\quad\{1,2\}\,,\qquad k=2\,:\,(0,2)\quad\rightarrow\quad\{2,2\}\,.

Then the former set makes up the row indices and the latter set makes up the column indices:

22LR(PL,2PL,2PR,2PR,2)=V1,2,\displaystyle\begin{array}[]{ccc}&2&2\\ L&&\\ R&&\end{array}\quad\rightarrow\quad\left(\begin{array}[]{cc}P_{{\rm L},2}&P_{{\rm L},2}\\ P_{{\rm R},2}&P_{{\rm R},2}\end{array}\right)=V_{1,2}\,, (S34)

and its permanent is now readily obtained to be

{V1,2}={PL,2PL,2PR,2PR,2}=PL,2PR,2+PL,2PR,2.\displaystyle\{V_{1,2}\}=\Big{\{}\begin{array}[]{cc}P_{{\rm L},2}&P_{{\rm L},2}\\ P_{{\rm R},2}&P_{{\rm R},2}\end{array}\Big{\}}=P_{{\rm L},2}\,P_{{\rm R},2}+P_{{\rm L},2}\,P_{{\rm R},2}\,. (S37)

For another example, let us suppose that N=3N=3 and try to express V1,2V_{1,2}. The first subscript 11 is converted into {L,L,R}\{L,L,R\} and the second one 22 is converted into {1,2,2}\{1,2,2\}. Then

122LLR(PL,1PL,2PL,2PL,1PL,2PL,2PR,1PR,2PR,2)=V1,2,\displaystyle\begin{array}[]{cccc}&1&2&2\\ L&&&\\ L&&&\\ R&&&\end{array}\quad\rightarrow\quad\left(\begin{array}[]{ccc}P_{{\rm L},1}&P_{{\rm L},2}&P_{{\rm L},2}\\ P_{{\rm L},1}&P_{{\rm L},2}&P_{{\rm L},2}\\ P_{{\rm R},1}&P_{{\rm R},2}&P_{{\rm R},2}\end{array}\right)=V_{1,2}\,, (S45)

and the permanent is {V1,2}=4PL,1PL,2PR,2+2PL,22PR,1\{V_{1,2}\}=4P_{{\rm L},1}P_{{\rm L},2}P_{{\rm R},2}+2P_{{\rm L},2}^{2}P_{{\rm R},1}. Now we have all ingredients to construct the probability distribution of a measurement for which the number of particles in each well is counted. To calculate the permanent of a matrix, we used the advanced algorithm developed to reduce the algorithmic complexity in [57]; for an introduction see [58].

I.4 IV. Additional details on the MLE

I.4.1 Construction of estimator and likelihood function

The maximum likelihood estimator (abbreviated already in the main text as MLE) is a commonly used estimator in the field of statistics and is defined as follows:

Xest=argmaxXP(𝐧|X),X_{\text{est}}=\text{argmax}_{X}P(\vec{\mathbf{n}}|X)\,, (S46)

where 𝐧\vec{\mathbf{n}} is used to denote the measurement outcome. Also, argmaxX denotes, by definition of the MLE, the unique point in the domain of interest, at which the function values are maximized. Whenever an outcome 𝐧\vec{\mathbf{n}} is attained, one inserts it into the RHS of Eq. (S46) and finds a value of XX that maximizes P(𝐧|X)P(\vec{\mathbf{n}}|X). This is a one-shot estimate of XX, namely XestX_{\text{est}}. In order to implement the MLE, it is necessary to obtain the likelihood function, i.e., P(𝐧|X)P(\vec{\mathbf{n}}|X), the process of which we now describe.

In the conventional TMI, only considering the change of Fock space coefficients, P(𝐧|X)P(\vec{\mathbf{n}}|X) is calculated as P(n|X)=n|Ψ^(t)=|Cn(t)|2P(\vec{n}|X)=\langle\vec{n}|\hat{\Psi}(t)\rangle=|C_{\vec{n}}(t)|^{2}, where |Ψ(t)=nCn(t)|n|\Psi(t)\rangle=\sum_{\vec{n}}C_{\vec{n}}(t)|\vec{n}\rangle. For the two-mode (double-well) system covered in the main text, cf. Eq.(3), we may consider the general two-mode state |Ψ(t)=k=0NCk(t)|Nk,k|\Psi(t)\rangle=\sum_{k=0}^{N}C_{k}(t)|N-k,k\rangle. Then n=(n1,n2)\vec{n}=(n_{1},n_{2}), denoting that n1n_{1} particles are in ϕ1(x,t)\phi_{1}(x,t) and n2n_{2} particles in ϕ2(x,t)\phi_{2}(x,t), is identified as the measurement result that n1n_{1} particles are in the left well and n2n_{2} particles are in the right well: 𝐧=(nL=n1,nR=n2)\vec{\mathbf{n}}=(n_{\rm L}=n_{1},n_{\rm R}=n_{2}). In particular, with the metrological protocol adopted in the main text, i.e., |Ψ(t)=ei(ϵJ^z+UJ^z2)t|Ψ(0)|\Psi(t)\rangle=e^{-i(\epsilon\hat{J}_{z}+U\hat{J}_{z}^{2})t}|\Psi(0)\rangle, each Ck(t)C_{k}(t) changes only by a phase (but not by magnitude): Ck(t)=exp(iϵN2k2t)exp(iU(N2k2)2t)Ck(0)C_{k}(t)=\exp(-i\epsilon\frac{N-2k}{2}t)\,\exp(-iU(\frac{N-2k}{2})^{2}t)\,C_{k}(0), where ϵ\epsilon contains the information of XX. Hence the likelihood P(𝐧|X)P(\vec{\mathbf{n}}|X) is independent of XX and invariant, which leads to vanishing CFI, see also the constant orange dashed line (coh TMI) in Fig. 4.

In the self-consistent approach, however, the calculation of P(𝐧|X)P(\vec{\mathbf{n}}|X) depends on the specifics of each system and measurement considered, since the measurement results are affected by the changing orbitals as well as by the changing Fock space coefficients. A quantum state is now written as |Ψ(t)=nCn(t)|n(t)|\Psi(t)\rangle=\sum_{\vec{n}}C_{\vec{n}}(t)|\vec{n}(t)\rangle, indicating that the orbitals associated by the Fock space basis state |n(t)|\vec{n}(t)\rangle evolve in time. In our double-well system, now n=(n1,n2)\vec{n}=(n_{1},n_{2}) cannot be interpreted as “n1n_{1} particles in the left well and n2n_{2} particles in the right well” anymore. The correct statement now is “n1n_{1} particles are in ϕ1(x,t)\phi_{1}(x,t) and n2n_{2} particles are in ϕ1(x,t)\phi_{1}(x,t)”. The orbitals may delocalize as time passes, thus at the instant of measurement a particle in the orbital ϕ1(x,t)\phi_{1}(x,t) can be found in the left well or in the right well with some probabilities PL,1P_{{\rm L},1} or PR,1P_{{\rm R},1}, respectively, where PL,j:=0|ϕj(x,t)|2𝑑xP_{L,j}:=\int_{-\infty}^{0}|\phi_{j}(x,t)|^{2}dx and PR,j:=0|ϕj(x,t)|2𝑑xP_{R,j}:=\int_{0}^{\infty}|\phi_{j}(x,t)|^{2}dx. In other words, at time tt, it is necessary to take further probability distributions into account other than just |Ck(t)|2|C_{k}(t)|^{2}:

P(𝐧=(Nj,j)|X)\displaystyle P(\vec{\mathbf{n}}=(N-j,j)|X) =\displaystyle= |Ck(t)|2,Conventional Two-Mode Interferometry\displaystyle|C_{k}(t)|^{2}\,,\qquad\qquad\qquad\,\,\,\,\text{Conventional Two-Mode Interferometry} (S47)
P(𝐧=(Nj,j)|X)\displaystyle P(\vec{\mathbf{n}}=(N-j,j)|X) =\displaystyle= k=0Pj,k|Ck(t)|2,Self-Consistent Approach\displaystyle\sum_{k=0}P_{j,k}|C_{k}(t)|^{2}\,,\qquad\qquad\qquad\quad\text{Self-Consistent Approach} (S48)

where the probability coefficients in (S48) read

Pj,k1N!(Nj){Vj,k}.P_{j,k}\coloneqq\frac{1}{N!}\left(\begin{array}[]{c}N\\ j\end{array}\right)\{V_{j,k}\}\,.

We refer to Eq. (S28) and the discussion it follows for the definition and calculation of the special matrix {Vj,k}\{V_{j,k}\}. In summary, the difference in obtaining the probabilities P(𝐧|X)P(\vec{\mathbf{n}}|X) as outlined in the above leads to a discrepancy in the probability distribution (synonymously likelihood), and therefore in the CFI and the MLE.

Refer to caption
Refer to caption
Figure S1: Probability distributions for the measurement outcome 𝐧=(nL,nR)=(Nj,j)\mathbf{n}=(n_{\text{L}},n_{\text{R}})=(N-j,j) with self-consistency taken into account.

For further illustration of the importance of self-consistency, see Fig. S1. These plots show how the probability distribution, calculated by Eq. (S28), changes according to the double-well metrological scenario of the main text. For the cat state (left), delocalization of orbitals is weak, and the probability distribution almost does not change even within the self-consistent framework. Then self-consistent metrology simply repeats the results of conventional two-mode interferometry. However, in the case of a spin-coherent state (right), the probability distribution evolves in time because of delocalizing orbitals. The conventional two-mode interferometry, i.e., Eq. (S47), predicts that the probability distribution remains identical to the initial one at t=0t=0 (black crosses). On the other hand, self-consisten metrology, i.e., Eq. (S48), predicts that the likelihood gets biased towards the left, and this is accurate since the particles are mostly located at the left well when p4>0p_{4}>0. As p4p_{4} increases, the likelihood gets even more biased, and this dependence on p4p_{4} results in a nonvanishing CFI.

I.4.2 Further results on MLE statistics

Additional details on the MLE are supplied in Fig. S2. The top left shows the mean of the maximum likelihood estimator with respect to the final state that has evolved under the true value of p4p_{4}. To calculate the mean, the probability distribution first needs to be composed. One measurement outcome is used at a single time of estimation, i.e., ν=1\nu=1. When the true value of p4p_{4} is 0.10.1, XestX0.0926\langle X_{\text{est}}\rangle_{X}\simeq 0.0926. The top right plot shows the mean of the maximum likelihood estimator versus ν\nu when p4=0.1p_{4}=0.1, which is the number of measurement outcomes for a single estimation of p4p_{4}. As ν\nu increases, XestX\langle X_{\text{est}}\rangle_{X} converges at around 0.110.11, which implies a bias of MLE. The bottom left shows this bias of the MLE, where |XestX/X|1\big{|}\partial\langle X_{\text{est}}\rangle_{X}/\partial X\big{|}\simeq 1 near p40p_{4}\simeq 0, which however does not hold as p4p_{4} deviates more significantly from zero. Finally, the bottom right shows the ratio between the mean-square deviation and the Cramér-Rao lower bound (CRLB). This ratio decreases as ν\nu increases: The MLE is known to make the mean-square deviation (δXest)2X\langle(\delta X_{\text{est}})^{2}\rangle_{X} converge to the CRLB as ν\nu\rightarrow\infty according to the central limit theorem, which is thus confirmed.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure S2: Additional results on maximum likelihood estimation, where “moe” is the mean of estimates (XestX\langle X_{\text{est}}\rangle_{X}), “msd” is the mean-square deviation ((δXest)2X\langle(\delta X_{\text{est}})^{2}\rangle_{X}), and “domoe” stands for the absolute value of the derivative of the mean of estimates (|XestX/X|\big{|}\partial\langle X_{\text{est}}\rangle_{X}/\partial X\big{|}).

I.5 V. Convergence of results for increasing number of modes: M=2M=2, M=3M=3, and M=4M=4

In the main text, it is shown that even for a weakly interacting gas, studying the CFI can exhibit a discrepancy between two-mode interferometry and self-consistent metrology. By monitoring ρtm\rho_{\text{tm}}, the validity of the two-mode approximation, M=2M=2, had been confirmed in the time interval of interest, and accordingly, two modes are used in the self-consistent framework, too. In the weakly interacting regime, two modes explain the primary quantum dynamics of the system and thus, the quantum Fisher information which is governed by the final state can be predicted suffiiciently accurate by two-mode metrology. Since interparticle interaction is weak, i.e., gN=0.1gN=0.1, M=2M=2 is thus used for the self-consistent results.

Refer to caption
Refer to caption
Figure S3: Monitoring the two-mode truncation after p4p_{4} is turned on, by verifying whether ρtm=(ρ1+ρ2)/N1\rho_{\rm tm}=(\rho_{1}+\rho_{2})/N\lesssim 1 (we put N=10N=10), with gNgN values as indicated. Cat state is on the left and spin-coherent state on the right.

To verify the validity of the two-mode approximation, we compare here with the M=3M=3 and M=4M=4 cases in a similar regimes of parameters, namely, gN=0.1gN=0.1, N=2N=2, and t=04t=0\sim 4. The validity of two-mode approximation is checked in Fig. S3. For M=3M=3 and M=4M=4, it is evident that ρtm1\rho_{\text{tm}}\lesssim 1, implying that the two-mode approximation works; the additional modes are rarely occupied during this time period. The impact of the additional modes is thus not crucial.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure S4: First and second row display quantum (𝔉\mathfrak{F}) and classical (F) Fisher information, respectively, plotted versus time tt, for cat and coherent (coh) states, respectively.

This also can be seen in the QFI with respect to the parameter p4p_{4}; see upper row in Fig. S4. For the weakly interacting gas, a small number of modes, e.g. here M=2M=2, is sufficient to describe the dynamics of the system and thus non-self-consistent two-mode interferometry calculates the QFI rather precisely. The self-consistent results with two (yellow solid), three (orange dash-dotted), and four (red dotted) modes are almost identical, and are not significantly different from the value of two-mode interferometry (black solid).

We have seen in the main text that the CFI for p4p_{4} can significantly differ when evaluated in the non-self-consistent and self-consistent frameworks. The lower row in Fig. S4 shows that the inclusion of more orbitals reproduces (and thus also further emphasizes) the different predictions made for M=2M=2 in the main text, when comparing the CFI in a self-consistent metrological framework relative to that in a non-self-consistent one. In the bottom left of Fig. S4, with a cat distribution of the coefficients of the initial state, the CFI for a certain measurement, i.e., measuring the number of particles in the left and right well, remains almost zero. On the other hand, in the bottom right of Fig. S4, with a spin-coherent distribution of the coefficients of the initial state, the CFI is nonvanishing as time passes by. This is confirmed for M=2M=2 (yellow solid), M=3M=3 (orange dash-dot), and M=4M=4 (red dotted). The nonvanishing CFI cannot be predicted in the conventional framework of non-self-consistent two-mode interferometry. The self-consistent approach thus makes it possible at all to accurately evaluate the precision limit for estimating the parameter p4p_{4}.