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

Protocol for nonlinear state discrimination in rotating condensate

Michael R. Geller Department of Physics and Astronomy, University of Georgia, Athens, Georgia 30602, USA
(June 21, 2024)
Abstract


Abstract

Nonlinear mean field dynamics enables quantum information processing operations that are impossible in linear one-particle quantum mechanics. In this approach, a register of bosonic qubits (such as neutral atoms or polaritons) is initialized into a symmetric product state |ψn|\psi\rangle^{\!\otimes n} through condensation, then subsequently controlled by varying the qubit-qubit interaction. We propose an experimental implementation of quantum state discrimination, an important subroutine in quantum computation, with a toroidal Bose-Einstein condensate. The condensed bosons here are atoms, each in the same superposition of angular momenta 0 and \hbar, encoding a qubit. A nice feature of the protocol is that only readout of individual quantized circulation states (not superpositions) is required.

I Introduction

A variety of atomtronic architectures have been proposed for quantum computing and quantum technology applications Amico et al. (2021, 2022). Two main Bose-Einstein condensate (BEC) types have been considered for realizing qubits: multi-component condensates and multi-mode condensates. Multi-component and spinor condensate approaches Cirac et al. (1998); Tian and Zoller (2003); Byrnes et al. (2012, 2015); Luiz et al. (2015); Großardt encode a single qubit in two (or more) metastable atomic states, such as spin or hyperfine levels, with all atoms in the same translational mode (for example the motional ground state). Multi-mode approaches Solenov and Mozyrsky (2011); Amico et al. (2014); Aghamalyan et al. (2016) encode a single qubit using two (or more) translational modes in a scalar condensate, such as a BEC in a double-well trapping potential. In the limit where there are a large number of condensed bosons in each well, the system becomes equivalent to two (or more) BECs, each with a well defined phase, connected by tunneling barriers that act as Josephson junctions Solenov and Mozyrsky (2011); Amico et al. (2014); Aghamalyan et al. (2016). Arrays of such BECs can be produced in optical lattices and are described by the Bose-Hubbard model Amico et al. (2014); Aghamalyan et al. (2016). Another multi-mode approach, which we adopt here, uses circulating states in a ring geometry Ryu et al. (2020); Kapale and Dowling (2005); Ryu et al. (2007); Ramanathan et al. (2011); Eckel et al. (2014); Kim et al. (2018); Pezzè et al. for the translational modes.

Given the demonstrated high performance and scalability of trapped ion qubits Moses et al. (2023), superconducting qubits Krinner et al. (2022); Google Quantum AI (2023); McKay et al. , and of neutral atom arrays Bluvstein et al. (2022); Graham et al. (2022), what does a BEC qubit offer? We argue that it offers a platform for an alternative approach to quantum information processing that leverages the special properties of condensates. In this approach, a BEC is used to prepare a register of qubits in a product state |ψn|\psi\rangle^{\!\otimes n} and control its subsequent evolution. From a quantum computing perspective, having multiple identical copies of an unknown input is already a useful resource (whereas classical information is freely cloned). In standard circuit-model quantum computation, illustrated in Fig. 1a, initialized qubits are subsequently entangled using two-qubit gates. Here we do the opposite and try to suppress entanglement, Fig. 1b. This is achieved by making nn large, interactions weak, and by preserving permutation symmetry. In this limit entanglement monogamy Coffman et al. (2000); Zong et al. bounds the pairwise concurrence to zero, and the BEC is exactly described by a nonlinear Schrödinger equation (the Gross–Pitaevskii equation Gross (1961); Pitaevskii (1961)), enabling novel dynamics Abrams and Lloyd (1998); Wu and Niu (2000); Riedel et al. (2010); Meyer and Wong (2013); Amico et al. (2014); Childs and Young (2016); Aghamalyan et al. (2016); Ryu et al. (2020); Xu et al. (2022); Deffner (2022); Geller (2023a). The theory is developed in a large nn limit with a rigorous bound on the error resulting from the mean field approximation. The nonlinear approach trades exponential time complexity for space complexity, requiring nn to be large.

Refer to caption
Figure 1: Nonlinear quantum information processing with a BEC. (a) In circuit-model quantum computation, a register of qubits is initialized to a product state (such as |0n|0\rangle^{\!\otimes n}), after which gates are applied, entangling the qubits. (b) In the nonlinear approach, the qubits ideally remain in a product state |ψ(t)n|\psi(t)\rangle^{\!\otimes n} throughout the computation. The BEC simulates a single nonlinear qubit.

Does this mean that nn has to be exponentially large? Actually the requirements on nn are not that bad. This is because the BEC is assumed to be initialized in a product state, and it takes time tentt_{\rm ent} for the atomic collisions to produce entanglement. Ideally, the whole experiment is performed in a short-time regime. We measure the accuracy of mean field theory by ϵ:=ρeff(t)ρ1(t)1\epsilon:=\|\rho_{\rm eff}(t)-\rho_{1}(t)\|_{1}, and call this the model error. Here ρeff\rho_{\rm eff} is the mean field state, ρ1\rho_{1} is the exact state traced over all atoms but one, tt is the gate duration, and 1\|\cdot\|_{1} is the trace norm. In a large family of condensate models Erdős and Schlein (2009); Geller (2023b)

ϵcet/tent1n,\displaystyle\epsilon\leq c\frac{e^{t/t_{\rm ent}}-1}{n}, (1)

where cc and tentt_{\rm ent} are positive constants (model-dependent quantities independent of tt and nn). Although the error might grow exponentially in time, there is always a short-time window t<tentt<t_{\rm ent} where the required number of condensed atoms nct/tentϵn\approx ct/t_{\rm ent}\epsilon is sub-exponential in tt.

We propose a demonstration of quantum information processing using this nonlinearity. In the remainder of this section we discuss the qubit encoding and state discrimination subroutine. The protocol is explained in Sec. II. Conclusions are given in Sec. III, and additional information about the BEC model and large nn limit are provided in the Appendix.

I.1 Qubit encoding

BEC-based qubits necessarily encode a small number of parameters (ψ0,1\psi_{0,1}\in{\mathbb{C}}) into a large number of degrees of freedom and the map is not unique. However two encodings can often be considered: Let ala_{l}^{\dagger} create an atom in BEC component l{0,1}l\in\{0,1\} (in a two-component condensate) or in translational mode l{0,1}l\in\{0,1\} (in a two-mode condensate), and let ψ0,1\psi_{0,1} be complex coordinates satisfying |ψ0|2+|ψ1|2=1|\psi_{0}|^{2}+|\psi_{1}|^{2}=1. One encoding that is interesting from a quantum foundations perspective is

|CATn:=ψ0(a0)n+ψ1(a1)nn!|vac,CATn|CATn=1,n1,\displaystyle|{\rm CAT}_{n}\rangle:=\frac{\psi_{0}(a_{0}^{\dagger})^{n}+\psi_{1}(a_{1}^{\dagger})^{n}}{\sqrt{n!}}|{\rm vac}\rangle,\ \ \langle{\rm CAT}_{n}|{\rm CAT}_{n}\rangle=1,\ \ n\geq 1, (2)

but this is a superposition of two macroscopically distinct BECs (a Schrödinger cat state) which would be highly susceptible to decoherence Cirac et al. (1998). Instead we use the encoding

|Fn:=(ψ0a0+ψ1a1)nn!|vac,Fn|Fn=1,n1,\displaystyle|F_{n}\rangle:=\frac{(\psi_{0}\,a_{0}^{\dagger}+\psi_{1}\,a_{1}^{\dagger})^{n}}{\sqrt{n!}}|{\rm vac}\rangle,\ \ \langle F_{n}|F_{n}\rangle=1,\ \ n\geq 1, (3)

which is a condensate of nn bosons ψ0a0+ψ1a1\psi_{0}\,a_{0}^{\dagger}+\psi_{1}\,a_{1}^{\dagger}, each a single atom in a superposition of components or modes. (While |CATn|{\rm CAT}_{n}\rangle and |Fn|{F}_{n}\rangle depend on both nn and ψ0,1\psi_{0,1}, the latter dependence is suppressed.) The encoding (3) was originally proposed by Cirac et al. Cirac et al. (1998) and by Byrnes et al. Byrnes et al. (2012, 2015) for two-component condensates; in that case |Fn|F_{n}\rangle is a pseudospin coherent state Byrnes et al. (2015). But our a0a^{\dagger}_{0} and a1a^{\dagger}_{1} create atoms in circulating states of orbital angular momentum 0 and \hbar, respectively, and it is better to regard |Fn|F_{n}\rangle as a coherent state of atoms in angular momenta superpositions. The states (3) are mean field states since the atoms are not entangled. They satisfy

al|Fn=ψln|Fn1andalal|Fn=ψlψln(n1)|Fn2.\displaystyle a_{l}|F_{n}\rangle=\psi_{l}\sqrt{n}|F_{n-1}\rangle\ \ \ {\rm and}\ \ \ a_{l}a_{l^{\prime}}|F_{n}\rangle=\psi_{l}\psi_{l^{\prime}}\sqrt{n(n-1)}|F_{n-2}\rangle. (4)

Because each atom in (3) carries a copy of the qubit state |ψ=ψ0|0+ψ1|1|\psi\rangle=\psi_{0}|0\rangle+\psi_{1}|1\rangle, the state |Fn|F_{n}\rangle exhibits a bosonic orthogonality catastrophe Guenther et al. (2021) in the large nn limit, meaning that close qubit states |ψ|\psi\rangle and |ψ|\psi^{\prime}\rangle encode to orthogonal |Fn|F_{n}\rangle and |Fn|F^{\prime}_{n}\rangle as nn\rightarrow\infty (the semiclassical limit in the spin coherent state picture Byrnes et al. (2015)). Furthermore, due to the polynomial encoding in |Fn|F_{n}\rangle, the single-particle superposition principle with respect to ψ0,1\psi_{0,1} is violated (see below).

In the atomtronic implementation we assume a toroidal BEC operated in a regime supporting metastable quantized circulation states with ll trapped vortices

|Φln=(al)nn|vac,\displaystyle|\Phi_{l}^{n}\rangle=\frac{(a_{l}^{\dagger})^{n}}{\sqrt{n}}|\rm vac\rangle, (5)

where ala_{l}^{\dagger} creates an atom in the ring with angular momentum ll\in{\mathbb{Z}}. These states are stabilized by the repulsive atomic interactions Mueller (2002); Baharian and Baym (2013). An atom with mass mm and l=1l=1 has velocity /mR\hbar/mR and circles the ring with angular velocity Ω0=/mR2\Omega_{0}=\hbar/mR^{2}. We construct a low-energy effective description for the BEC within the manifold of states (3). This is possible because they are selected out by the path integral in the large nn limit, due to their diverging contribution to the action. The action in the subspace spanned by these states is

Seff[ψ¯l,ψl]=𝑑tFn|itHrot|Fn,\displaystyle S_{\rm eff}[{\bar{\psi}}_{l},\psi_{l}]=\int\!dt\,\langle F_{n}|i\partial_{t}-H_{\rm rot}|F_{n}\rangle, (6)

where HrotH_{\rm rot} is the BEC Hamiltonian in the rotating frame. The BEC is rotated with frequency ΩΩ0/2\Omega\approx\Omega_{0}/2 to bring the 0-vortex (no circulation) state |Φ0n|\Phi_{0}^{n}\rangle and the 1-vortex state |Φ1n|\Phi_{1}^{n}\rangle close in energy. Higher energy ll are then neglected, leading to a two-mode model. In the large nn limit (see Appendix) the saddle point equations are

ddt(ψ0ψ1)=iHeff(ψ0ψ1),Heff=V01σx+Bzσz+g(|ψ0|2|ψ1|2)σz.\displaystyle\frac{d}{dt}\!\begin{pmatrix}\psi_{0}\\ \psi_{1}\end{pmatrix}\!=\!-iH_{\rm eff}\!\begin{pmatrix}\psi_{0}\\ \psi_{1}\end{pmatrix}\!,\ \ H_{\rm eff}=V_{01}\sigma^{x}+B_{z}\sigma^{z}+g(|\psi_{0}|^{2}\!-\!|\psi_{1}|^{2})\,\sigma^{z}\!. (7)

The first two terms in HeffH_{\rm eff} generate rigid xx and zz rotations of the Bloch sphere. Rotations about xx couple l=0l=0 and l=1l=1 angular momenta and are produced by breaking rotational symmetry. Here V01V_{01} is a matrix element for an applied potential energy barrier. The parameter BzB_{z} is controlled by the frequency Ω\Omega of the BEC rotation discussed above. The nonlinear term describes a zz rotation with a rate that increases with increasing Bloch sphere coordinate tr(ρσz)=|ψ0|2|ψ1|2{\rm tr}(\rho\sigma^{z})=|\psi_{0}|^{2}\!-\!|\psi_{1}|^{2}, vanishes on the equator, and reverses direction for tr(ρσz)<0{\rm tr}(\rho\sigma^{z})<0. This zz-axis torsion Mielnik (1980) (1-axis twisting Kitagawa and Ueda (1993)) of the Bloch sphere is the key to fast state discrimination, but is prohibited in ordinary single-particle quantum mechanics. Although the qubit here is informational and not associated with any physical 2-state system, we can define a logical basis {|0,|1}\{|0\rangle,|1\rangle\} and treat it like any other qubit:

|ψ=ψ0|0+ψ1|1=(ψ0ψ1),|0:=(10)=|Φ0n,|1:=(01)=|Φ1n.\displaystyle|\psi\rangle=\psi_{0}|0\rangle+\psi_{1}|1\rangle=\begin{pmatrix}\psi_{0}\\ \psi_{1}\end{pmatrix}\!,\ \ |0\rangle:=\begin{pmatrix}1\\ 0\end{pmatrix}=|\Phi_{0}^{n}\rangle,\ \ |1\rangle:=\begin{pmatrix}0\\ 1\end{pmatrix}=|\Phi_{1}^{n}\rangle. (8)

It should be emphasized that (3) is the physical state of the quantum gas, not (8). However the basis states |0,|1|0\rangle,|1\rangle are the quantized circulation states |Φ0,1n|\Phi_{0,1}^{n}\rangle, which is important for the readout step.

I.2 Single-input state discrimination

As an application, we consider the problem of quantum state discrimination Barnet and Croke ; Bae and Hwang (2013); Bae and Kwek (2015); Rouhbakhsh and Ghoreishi , a basic task in quantum information science. In the two-state variant considered here, a quantum state |ψ{|a,|b}|\psi\rangle\in\{|a\rangle,|b\rangle\} is input to a processor, which knows the values of |a|a\rangle and |b|b\rangle ahead of time and tries determine which was provided (with a bounded failure probability). This is easy if |a|a\rangle and |b|b\rangle are orthogonal: For a qubit, a single unitary Uread=|0a|+|1b|U_{\rm read}=|0\rangle\langle a|+|1\rangle\langle b| rotates α|a+β|b\alpha|a\rangle+\beta|b\rangle to α|0+β|1\alpha|0\rangle+\beta|1\rangle, which is then measured in the standard basis. The challenging case is when |a|a\rangle and |b|b\rangle are similar, |a|b|2=12kwithk1|\langle a|b\rangle|^{2}=1-2^{-k}\ {\rm with}\ k\gg 1, where n>2kn>2^{k} identical copies of the input are required Helstrom (1976). In minimum-error discrimination, the subroutine selects |a|a\rangle or |b|b\rangle, each with some probability of error, and the objective is to minimize the average error. In unambiguous state discrimination, the subroutine identifies |a|a\rangle or |b|b\rangle perfectly, but has the possibility of abstaining, returning an inconclusive result. State discrimination can be used to solve NP-complete (and harder) problems Abrams and Lloyd (1998); Aaronson ; Childs and Young (2016), at the expense of 2k2^{k} input copies and exponential runtime. This cost reflects the limited information gained from measurement.

Abrams and Lloyd Abrams and Lloyd (1998) showed that certain nonlinearity in the Schrödinger equation would bypass this exponential cost, allowing NP-complete problems to be solved efficiently (in an idealized setting with no errors or decoherence). But the presence of such nonlinearity would constitute a fundamental modification of quantum mechanics that is not supported by experiment Bollinger et al. (1989); Chupp and Hoare (1990); Walsworth et al. (1990); Majumder et al. (1990). In a condensate, the nonlinearity is not fundamental, but effective. Although we can realize nonlinear gates, this doesn’t constitute a complexity violation, due to the large nn requirement of mean field theory.

II Protocol

Refer to caption
Figure 2: State discrimination channel. Here a|b0\langle a|b\rangle\neq 0 but 0|1=0\langle 0|1\rangle=0, so the channel must be nonlinear. Note that the output is always a basis state, |0|0\rangle or |1|1\rangle, simplifying readout.

The process is illustrated in Fig. 2. A single state |ψ{|a,|b}|\psi\rangle\in\{|a\rangle,|b\rangle\} is input to the discriminator, which ideally returns output |0|0\rangle if |ψ=|a|\psi\rangle\!=\!|a\rangle, or returns |1|1\rangle if |ψ=|b|\psi\rangle\!=\!|b\rangle. The single-input discriminator regarded as a channel must be nonunitary, because the overlap a|b\langle a|b\rangle is not preserved. Equivalently, the distance ρaρb1\|\rho_{a}-\rho_{b}\|_{1} between their density matrices in trace norm is not preserved in time (here X1:=trXX\|X\|_{1}:={\rm tr}\sqrt{X^{\dagger}X}). For pure states, ρaρb1=2|sin(θab/2)|\|\rho_{a}-\rho_{b}\|_{1}=2|\sin(\theta_{ab}/2)|, where θab\theta_{ab} is the angle between their Block vectors. Linear completely-positive trace preserving (CPTP) channels satisfy ddtρaρb10\frac{d}{dt}\|\rho_{a}-\rho_{b}\|_{1}\leq 0; they are either distance preserving or strictly contractive on the inputs Ruskai (1994). Because the discriminator orthogonalizes the potential inputs, it is expansive on those inputs: ddtρaρb1>0\frac{d}{dt}\|\rho_{a}-\rho_{b}\|_{1}>0. Thus, the discriminator is described by a nonlinear PTP channel Abrams and Lloyd (1998); Mielnik (1980); Geller (2023a).

The implementation proposed here does not discriminate an unknown input (produced by a previous computation), but instead uses a black box state preparation step to randomly prepare |a|a\rangle or |b|b\rangle, with a small Bloch vector angle θab0\theta_{ab}\geq 0 between them. Then |a|b|2=cos2(θab/2)1(θab/2)2|\langle a|b\rangle|^{2}=\cos^{2}(\theta_{ab}/2)\approx 1-(\theta_{ab}/2)^{2}. This can be accomplished by initializing in the |Φ0n|\Phi^{n}_{0}\rangle state and using V01V_{01} and BzB_{z} in (7) to apply xx and zz rotations. (Ideally, this step is hidden from the remainder of the experiment.) The discrimination gate itself follows Refs. Abrams and Lloyd (1998); Childs and Young (2016) and uses the zz-axis torsion to increase the angle between |a|a\rangle and |b|b\rangle. It’s clear that |a|a\rangle and |b|b\rangle should begin with equal and opposite zz components za=zbz_{a}=-z_{b} [here ra,bμ=tr(ρa,bσμ),μ{1,2,3}r^{\mu}_{a,b}={\rm tr}(\rho_{a,b}\sigma^{\mu}),\ \mu\in\{1,2,3\}]. Consider a simple option with ya,b=0y_{a,b}=0, namely

|a\displaystyle|a\rangle =\displaystyle= cos(πθab4)|0+sin(πθab4)|1,\displaystyle\cos\bigg{(}\frac{\pi-\theta_{ab}}{4}\bigg{)}|0\rangle+\sin\bigg{(}\frac{\pi-\theta_{ab}}{4}\bigg{)}|1\rangle, (9)
|b\displaystyle|b\rangle =\displaystyle= cos(π+θab4)|0+sin(π+θab4)|1,\displaystyle\cos\bigg{(}\frac{\pi+\theta_{ab}}{4}\bigg{)}|0\rangle+\sin\bigg{(}\frac{\pi+\theta_{ab}}{4}\bigg{)}|1\rangle, (10)

which has

xa=xb=|cos(θab2)|,ya=yb=0,za=sin(θab2),zb=sin(θab2).\displaystyle x_{a}=x_{b}=\bigg{|}\!\cos\bigg{(}\frac{\theta_{ab}}{2}\bigg{)}\bigg{|},\ \ y_{a}=y_{b}=0,\ \ z_{a}=\sin\bigg{(}\frac{\theta_{ab}}{2}\bigg{)},\ \ z_{b}=-\sin\bigg{(}\frac{\theta_{ab}}{2}\bigg{)}. (11)

After switching on gg, the two input options evolve as Rz(±gtθab)R_{z}(\pm gt\theta_{ab}) and orthogonalize after a time tπ/gθabt\approx\pi/g\theta_{ab}. However this implementation does not have a favorable scaling with θab\theta_{ab}. The optimal protocol for nonlinear discrimination was derived by Childs and Young (CY) in Childs and Young (2016). Instead of (11), the CY gate begins with

xa=xb=|cos(θab2)|,ya=za=sin(θab2)2,yb=zb=sin(θab2)2,\displaystyle x_{a}=x_{b}=\bigg{|}\!\cos\bigg{(}\frac{\theta_{ab}}{2}\bigg{)}\bigg{|},\ \ y_{a}=z_{a}=\frac{\sin(\frac{\theta_{ab}}{2})}{\sqrt{2}},\ \ y_{b}=z_{b}=-\frac{\sin(\frac{\theta_{ab}}{2})}{\sqrt{2}}, (12)

and applies xx rotations to hold ya,b=za,by_{a,b}=z_{a,b} during the subsequent evolution in order to reach antipodal points on the Bloch sphere. The options orthogonalize in a time t=O(log1θab)t=O(\log\frac{1}{\theta_{ab}}), after which a readout gate UreadU_{\rm read} (defined with respect to time-evolved |a,|b|a\rangle,|b\rangle) transforms them to circulation states |Φ0n|\Phi_{0}^{n}\rangle or |Φ1n|\Phi_{1}^{n}\rangle, which are then measured via time-of-flight Amico et al. (2005); Moulder et al. (2012).

In an idealized context where (7) is regarded as exact, and where there are no control errors, readout errors, decoherence errors, or noise, the nonlinear discriminator works perfectly every time. We refer to this idealization as a single-input discriminator to distinguish it from the more familiar minimum error and unambiguous discriminators based on linear CPTP channels Barnet and Croke ; Bae and Hwang (2013); Bae and Kwek (2015); Rouhbakhsh and Ghoreishi . Of course any actual atomtronic realization is likely to suffer from all such errors, and may fail to give the correct answer or return an answer at all. Although the theoretically achievable performance depends sensitively on the system and device details, and is beyond the scope of this work, we note that combining torsion with non-CP dissipation is predicted to implement an autonomous discriminator Geller (2023a), whose control sequence and operation is (mostly) independent of |a|a\rangle and |b|b\rangle. In this implementation, the nonlinearity and dissipation create two basins of attraction with a shared boundary in the Bloch ball, one with an attracting fixed point near |0|0\rangle and the other with an attracting fixed point near |1|1\rangle, giving the discriminator a degree of intrinsic fault-tolerance.

Refer to caption
Figure 3: Linear versus nonlinear qubit evolution. (a) Unitary evolution. Vectors show the isometric flow of states on the Bloch sphere generated by linear Hamiltonian H=σzH=\sigma^{z}. (b) Torsion dynamics generated by nonlinear Hamiltonian H=ψ|σz|ψσzH=\langle\psi|\sigma^{z}|\psi\rangle\,\sigma^{z}.

The presence of channel nonlinearity indicates a breakdown of the superposition principle. Figure 3 illustrates a nice example of this effect: In Fig. 3a, the evolution of a superposition ψ0|0+ψ1|1\psi_{0}|0\rangle+\psi_{1}|1\rangle is given by a superposition of evolved basis states eit|0e^{-it}|0\rangle and eit|1e^{it}|1\rangle, shown as a velocity field. However in Fig. 3b, the evolved states eit|0e^{-it}|0\rangle and eit|1e^{-it}|1\rangle are now static (phase factors are a global phase), whereas the actual dynamics is not, except on the equatorial plane.

III Conclusions

We have discussed an approach to quantum information processing that leverages the special properties of condensates, including their nonlinearity, and proposed an atomtronic implementation of a “nonlinear” qubit. An experimental demonstration of nonlinear state discrimination, while striking, would not by itself constitute a computation, because the qubit isn’t coupled to anything. To implement a useful computation, the BEC qubit must be entangled with other qubits (for example trapped ions) in a scalable circuit-model quantum computer, which is not addressed here.

The standard models of quantum computation assume gates and errors based on linear CPTP channels. Physical hardware, however, might admit initial correlation and be better described by more general maps Dominy et al. (2016); Dominy and Lidar (2016). It is therefore interesting to investigate any additional computational power enabled by quantum channels beyond the linear CPTP paradigm, as we did here. Another example was investigated by Chen et al. Chen et al. , who experimentally demonstrated unambiguous state discrimination in a linear but non-Hermitian optical system. After completing this work, Großardt posted a preprint Großardt proposing the use of a two-component BEC coupled to a neutral atom computer to simulate a large family of nonlinear Schrödinger equations. Given their potential for fast quantum state discrimination and simulation of the nonlinear Schrödinger equation, the non-Hermitian and nonlinear approaches to quantum information processing deserve further exploration.

Acknowledgements.
This work was partly supported by the NSF under grant no. DGE-2152159.

Appendix A BEC model

Here we derive the qubit equation of motion (7). We consider a toroidal BEC (thin circular ring with radius RR) with rotating tunneling barriers that act as Josephson junctions Amico et al. (2014); Eckel et al. (2014); Aghamalyan et al. (2016); Ryu et al. (2020). Thin means the dynamics is quasi-1d in the azimuthal direction, θ\theta. This requires the energy, temperature, and effective interaction strength to be below an energy scale Δϵ\Delta\epsilon determined by the confining potential. The shape of the potential (without barriers) is mostly arbitrary as long as it is invariant under rotations about the axis threading the ring, which we call the zz axis. The angular momentum eigenfunctions on the ring are φl(θ)=eilθ/2πR,\varphi_{l}(\theta)=e^{il\theta}/\sqrt{2\pi R}, with ll\in{\mathbb{Z}} the angular momentum. In the absence of the tunnel barriers and interaction, these are stationary states. The condensate consists of nn weakly interacting bosonic atoms of mass mm, each in their electronic ground state |Ψ0|\Psi_{0}\rangle. At sufficiently low energy and densitiy, the atomic collisions are elastic, and the Hamiltonian is

H(t)=Vold3r{2ϕϕ2m+U2ϕϕϕϕ+Vϕϕ},[ϕ(𝐫),ϕ(𝐫)]=δ(𝐫𝐫).\displaystyle H(t)=\int_{{\rm Vol}}\!\!d^{3}r\,\bigg{\{}\frac{\hbar^{2}\nabla\phi^{\dagger}\!\cdot\!\nabla\phi}{2m}+\frac{U}{2}\,\phi^{\dagger}\phi^{\dagger}\phi\,\phi+V\phi^{\dagger}\phi\bigg{\}},\ \ [\phi({\bf r}),\phi^{\dagger}({\bf r}^{\prime})]=\delta({\bf r}-{\bf r}^{\prime}). (13)

Here Vol{\rm Vol} is the volume of the ring, U=4π2as/mU=4\pi\hbar^{2}a_{\rm s}/m is a short-range interaction strength (proportional to the s-wave scattering length asa_{\rm s}), and V(𝐫,t)V({\bf r},t) is a confining potential, including the rotating barriers. Acting on the vacuum, ϕ(𝐫)\phi^{\dagger}({\bf r}) creates a bosonic atom in state |Ψ0|\Psi_{0}\rangle at point 𝐫{\bf r}. We assume a tunable repulsive interaction with as0a_{\rm s}\geq 0. We also assume zero temperature, no dissipation, and no disorder.

Two rotating tunnel barriers are used to implement an atomtronic quantum interference device Aghamalyan et al. (2016); Ryu et al. (2020). When the barriers are turned on, the Hamiltonian (13) is time dependent. Assuming the barriers are rigidly rotated about the zz axis with frequency Ω\Omega, we have V(𝐫,t)=eiΩtLz/V(𝐫,0)eiΩtLz/V({\bf r},t)=e^{-i\Omega tL_{z}/\hbar}\,V({\bf r},0)\,e^{i\Omega tL_{z}/\hbar}, where LzL_{z} is the angular momentum.

However, we can transform to a noninertial reference frame in which the Hamiltonian, HrotH_{\rm rot}, is time independent. Decomposing the time-evolution operator in the lab frame as

Ulab=Tei0tH𝑑t=eiΩtLz/Urot,\displaystyle U_{\rm lab}=Te^{-\frac{i}{\hbar}\int_{0}^{t}Hdt^{\prime}}=e^{-i\Omega tL_{z}/\hbar}U_{\rm rot}, (14)

we obtain

dUrotdt=iHrotUrot,Hrot=eiΩtLz/(HΩLz)eiΩtLz/=H(0)ΩLz.\displaystyle\frac{dU_{\rm rot}}{dt}=-\frac{i}{\hbar}H_{\rm rot}U_{\rm rot},\ \ H_{\rm rot}=e^{i\Omega tL_{z}/\hbar}(H-\Omega L_{z})e^{-i\Omega tL_{z}/\hbar}=H(0)-\Omega L_{z}. (15)

Next we discuss the two-mode limit: In the low energy, thin ring limit, we can expand the field operators and angular momentum as

ϕ(𝐫)=leilθVolal,Lz=llalal,[al,al]=δll,\displaystyle\phi({\bf r})=\sum_{l}\frac{e^{il\theta}}{\sqrt{{\rm Vol}}}\,a_{l},\ \ L_{z}=\sum_{l}\hbar la^{\dagger}_{l}a_{l},\ \ [a_{l},a_{l^{\prime}}^{\dagger}]=\delta_{ll^{\prime}}, (16)

which leads to

Hrot\displaystyle H_{\rm rot} =\displaystyle= Ω02ll2alal+U2Voll1,l2,l3al1+l3al2l3al2al1+l1,l2Vl1l2al1al2Ωllalal,\displaystyle\frac{\hbar\Omega_{0}}{2}\sum_{l}l^{2}a_{l}^{\dagger}a_{l}+\frac{U}{2\,{\rm Vol}}\sum_{l_{1},l_{2},l_{3}}a_{l_{1}+l_{3}}^{\dagger}a_{l_{2}-l_{3}}^{\dagger}a_{l_{2}}a_{l_{1}}+\sum_{l_{1},l_{2}}\,V_{l_{1}l_{2}}\,a_{l_{1}}^{\dagger}a_{l_{2}}-\hbar\Omega\sum_{l}la^{\dagger}_{l}a_{l},\ \ \ (17)

where

Vl1l2\displaystyle V_{l_{1}l_{2}} =\displaystyle= dθ2πV(θ,t=0)ei(l1l2)θ.\displaystyle\oint\!\frac{d\theta}{2\pi}\,V(\theta,t\!=\!0)\,e^{-i(l_{1}-l_{2})\theta}. (18)

Nonzero Vl1l2V_{l_{1}l_{2}} induce transitions between angular momentum states. Then we have

Hrot=lωlalal+U2Voll1,l2,l3al1+l3al2l3al2al1+l1,l2Vl1l2al1al2nΩ22Ω0,\displaystyle H_{\rm rot}=\sum_{l}\hbar\omega_{l}a_{l}^{\dagger}a_{l}+\frac{U}{2\,{\rm Vol}}\sum_{l_{1},l_{2},l_{3}}a_{l_{1}+l_{3}}^{\dagger}a_{l_{2}-l_{3}}^{\dagger}a_{l_{2}}a_{l_{1}}+\sum_{l_{1},l_{2}}\,V_{l_{1}l_{2}}\,a_{l_{1}}^{\dagger}a_{l_{2}}-\frac{n\hbar\Omega^{2}}{2\Omega_{0}}, (19)

where

ωl=(ΩlΩ0)22Ω0,ω0=Ω22Ω0,ω1=(ΩΩ0)22Ω0.\displaystyle\omega_{l}=\frac{(\Omega-l\Omega_{0})^{2}}{2\Omega_{0}},\ \ \omega_{0}=\frac{\Omega^{2}}{2\Omega_{0}},\ \ \omega_{1}=\frac{(\Omega-\Omega_{0})^{2}}{2\Omega_{0}}. (20)

As explained above, the BEC is rotated with frequency ΩΩ0/2\Omega\approx\Omega_{0}/2 to bring the l=0l=0 and l=1l=1 states close in energy. We restrict (19) to angular momenta l=0,1l=0,1 neglecting the others on the basis of their higher energy. Then

l1,l2,l3al1+l3al2l3al2al1\displaystyle\sum_{l_{1},l_{2},l_{3}}a_{l_{1}+l_{3}}^{\dagger}a_{l_{2}-l_{3}}^{\dagger}a_{l_{2}}a_{l_{1}} =\displaystyle= l{alala0a0+al+1ala0a1+ala1la1a0+a1+la1la1a1}\displaystyle\sum_{l\in{\mathbb{Z}}}\bigg{\{}a_{l}^{\dagger}a_{-l}^{\dagger}a_{0}a_{0}+a_{l+1}^{\dagger}a_{-l}^{\dagger}a_{0}a_{1}+a_{l}^{\dagger}a_{1-l}^{\dagger}a_{1}a_{0}+a_{1+l}^{\dagger}a_{1-l}^{\dagger}a_{1}a_{1}\bigg{\}}\ \ \ \ (21)
=\displaystyle= a0a0a0a0+a1a1a1a1+4a0a1a1a0\displaystyle a_{0}^{\dagger}a_{0}^{\dagger}a_{0}a_{0}+a_{1}^{\dagger}a_{1}^{\dagger}a_{1}a_{1}+4a_{0}^{\dagger}a_{1}^{\dagger}a_{1}a_{0} (22)
=\displaystyle= (a0a0)2a0a0+(a1a1)2a1a1+4a0a0a1a1.\displaystyle(a_{0}^{\dagger}a_{0})^{2}-a_{0}^{\dagger}a_{0}+(a_{1}^{\dagger}a_{1})^{2}-a_{1}^{\dagger}a_{1}+4a_{0}^{\dagger}a_{0}a_{1}^{\dagger}a_{1}. (23)

This leads to a two-mode model

Hrot=l=0,1(ωl+Vll+γalalγ)alal+γa0a0a1a1+(V01a0a1+V¯01a1a0)nΩ22Ω0,\displaystyle H_{\rm rot}\!=\!\sum_{l=0,1}\bigg{(}\!\hbar\omega_{l}+V_{ll}+\gamma a_{l}^{\dagger}a_{l}-\gamma\!\bigg{)}a_{l}^{\dagger}a_{l}+\gamma^{\prime}a_{0}^{\dagger}a_{0}a_{1}^{\dagger}a_{1}+\bigg{(}\!V_{01}a_{0}^{\dagger}a_{1}\!+\!{\bar{V}_{01}}a_{1}^{\dagger}a_{0}\!\bigg{)}-\frac{n\hbar\Omega^{2}}{2\Omega_{0}},\ \ \ \ \ \ \ (24)

where

γ=U2Vol,γ=4γ=2UVol.\displaystyle\gamma=\frac{U}{2\,{\rm Vol}},\ \ \gamma^{\prime}=4\gamma=\frac{2U}{{\rm Vol}}. (25)

In what follows we will treat γ,γ0\gamma,\gamma^{\prime}\geq 0 as independent parameters, allowing (24) to apply to other systems as well. The last term in (24) subtracts the classical kinetic energy of the spinning ring: nΩ2/2Ω0=12IringΩ2,Iring=nmR2.n\hbar\Omega^{2}\!/2\Omega_{0}=\frac{1}{2}I_{\rm ring}\Omega^{2},\ I_{\rm ring}=nmR^{2}\!.

Finally we discuss the large nn limit: Condensates feature an enhanced two-particle interaction aaaan(n1)\langle a^{\dagger}a^{\dagger}aa\rangle\approx n(n-1) caused by the effectively infinite-ranged interaction between condensed atoms. This makes a naive large nn limit unphysical, because the energy per particle diverges Lieb et al. (2000), and our low-energy assumptions would be violated. The framework discussed here is instead based on a modified large nn limit where the interaction simultaneously weakens as 1/n1/n (a standard assumption in rigorous studies of mean field theory Lieb et al. (2000); Benedikter et al. (2016)). This allows for a rigorous study of the large nn limit including bounds on the accuracy of mean field theory. For real V01=dθ2πV(θ)eiθV_{01}\!=\!\oint\!\frac{d\theta}{2\pi}\,V(\theta)\,e^{i\theta}, and after dropping the classical kinetic energy term (which does not affect the qubit dynamics) we obtain

Hrot=l=0,1((ωl+Vll)alal+γalalalalγalal)+γa0a0a1a1+V01(a0a1+a1a0).\displaystyle H_{\rm rot}=\sum_{l=0,1}\bigg{(}\!(\!\hbar\omega_{l}+V_{ll})a_{l}^{\dagger}a_{l}+\gamma a_{l}^{\dagger}a_{l}^{\dagger}a_{l}a_{l}-\gamma a_{l}^{\dagger}a_{l}\!\bigg{)}+\gamma^{\prime}a_{0}^{\dagger}a_{0}a_{1}^{\dagger}a_{1}+V_{01}\bigg{(}\!a_{0}^{\dagger}a_{1}\!+\!a_{1}^{\dagger}a_{0}\!\bigg{)}.\ \ \ (26)

Evaluating (6) and assuming n1n\gg 1 leads to (setting =1\hbar=1)

Seff=n𝑑t{l=0,1(ψ¯litψl(ωl+Vll)|ψl|2nγ|ψl|4)nγ|ψ0ψ1|2V01(ψ¯0ψ1+ψ¯1ψ0)}.\displaystyle S_{\rm eff}\!=\!n\!\!\int\!\!dt\bigg{\{}\!\sum_{l=0,1}\!\bigg{(}\!{\bar{\psi}}_{l}i\partial_{t}\psi_{l}\!-\!(\!\omega_{l}+V_{ll})|\psi_{l}|^{2}\!-\!n\gamma\,|\psi_{l}|^{4}\!\bigg{)}\!-\!n\gamma^{\prime}|\psi_{0}\psi_{1}|^{2}\!-\!V_{01}({\bar{\psi}}_{0}\psi_{1}\!+\!{\bar{\psi}}_{1}\psi_{0})\!\bigg{\}}.\ \ \ \ \ \ \ \ (27)

Here z¯{\bar{z}} denotes complex conjugation. Due to the O(n2)O(n^{2}) interaction energies in (26), we cannot take the nn\rightarrow\infty limit in (27) without violating our low-energy assumptions. Instead we consider a modified limit defined by the simultaneous limits γ,γ0\gamma,\gamma^{\prime}\rightarrow 0, nn\rightarrow\infty, and low energy. To evaluate (27) in this limit we assume that the interaction strengths decrease with nn as γ=K/n\gamma=K/n and γ=K/n\gamma^{\prime}=K^{\prime}/n, where KK and KK^{\prime} are now fixed coupling constants. Then

Seff=n𝑑t{l=0,1(ψ¯litψl(ωl+Vll)|ψl|2K|ψl|4)K|ψ0ψ1|2V01(ψ¯0ψ1+ψ¯1ψ0)}.\displaystyle S_{\rm eff}\!=\!n\!\!\int\!\!dt\bigg{\{}\!\sum_{l=0,1}\!\bigg{(}\!{\bar{\psi}}_{l}i\partial_{t}\psi_{l}\!-\!(\!\hbar\omega_{l}+V_{ll})|\psi_{l}|^{2}\!-\!K|\psi_{l}|^{4}\bigg{)}\!-\!K^{\prime}|\psi_{0}\psi_{1}|^{2}\!-\!V_{01}({\bar{\psi}}_{0}\psi_{1}\!+\!{\bar{\psi}}_{1}\psi_{0})\!\bigg{\}}.\ \ \ \ \ \ \ \ (28)

To obtain (28) we have used the results summarized below in Table 1, which also gives the corresponding results for encoding (2). In the large nn limit the stationary phase approximation leads to (7) with

Heff\displaystyle H_{\rm eff} =\displaystyle= (ω0+V00+2K|ψ0|2+K|ψ1|2V01V01ω1+V11+2K|ψ1|2+K|ψ0|2)\displaystyle\begin{pmatrix}\hbar\omega_{0}+V_{00}+2K|\psi_{0}|^{2}+K^{\prime}\,|\psi_{1}|^{2}&V_{01}\\ V_{01}&\hbar\omega_{1}+V_{11}+2K|\psi_{1}|^{2}+K^{\prime}\,|\psi_{0}|^{2}\end{pmatrix} (29)
=\displaystyle= V01σx+Bzσz+gtr(ρσz)σz+const.,\displaystyle V_{01}\sigma^{x}+B_{z}\sigma^{z}+g\,{\rm tr}(\rho\sigma^{z})\,\sigma^{z}+{\rm const.}, (30)

where

Bz:=ω0ω1+V00V112andg:=2KK2.\displaystyle B_{z}:=\frac{\hbar\omega_{0}-\hbar\omega_{1}+V_{00}-V_{11}}{2}\ \ {\rm and}\ \ g:=\frac{2K-K^{\prime}}{2}. (31)

This concludes the derivation of (7).

Table 1: One- and two-particle correlators versus encoding.
alal\langle a_{l}^{\dagger}a_{l^{\prime}}\rangle alalalal\langle a_{l}^{\dagger}a_{l}a_{l^{\prime}}^{\dagger}a_{l^{\prime}}\rangle
|CATn|{\rm CAT}_{n}\rangle n|ψl|2δlln\,|\psi_{l}|^{2}\delta_{ll^{\prime}} n(n1)|ψl|2δlln(n-1)\,|\psi_{l}|^{2}\delta_{ll^{\prime}}
|Fn|{F}_{n}\rangle nψlψln\,\psi_{l}^{*}\psi_{l^{\prime}} n(n1)|ψl|2|ψl|2n(n-1)\,|\psi_{l}|^{2}|\psi_{l^{\prime}}|^{2}

References

  • Amico et al. (2021) L. Amico, M. Boshier, G. Birkl, A. Minguzzi, C. Miniatura, L.-C. Kwek, D. Aghamalyan, V. Ahufinger, D. Anderson, N. Andrei, A. S. Arnold, M. Baker, T. A. Bell, T. Bland, J. P. Brantut, D. Cassettari, W. J. Chetcuti, F. Chevy, R. Citro, S. De Palo, R. Dumke, M. Edwards, R. Folman, J. Fortagh, S. A. Gardiner, B. M. Garraway, G. Gauthier, A. Günther, T. Haug, C. Hufnagel, M. Keil, P. Ireland, M. Lebrat, W. Li, L. Longchambon, J. Mompart, O. Morsch, P. Naldesi, T. W. Neely, M. Olshanii, E. Orignac, S. Pandey, A. Pérez-Obiol, H. Perrin, L. Piroli, J. Polo, A. L. Pritchard, N. P. Proukakis, C. Rylands, H. Rubinsztein-Dunlop, F. Scazza, S. Stringari, F. Tosto, A. Trombettoni, N. Victorin, W. von Klitzing, D. Wilkowski, K. Xhani,  and A. Yakimenko, “Roadmap on atomtronics: State of the art and perspective,” AVS Quantum Sci. 3, 039201 (2021).
  • Amico et al. (2022) L. Amico, D. Anderson, M. Boshier, J.-P. Brantut, L.-C. Kwek, A. Minguzzi,  and W. von Klitzing, “Atomtronic circuits: From many-body physics to quantum technologies,” Rev. Mod. Phys. 94, 041001 (2022).
  • Cirac et al. (1998) J. I. Cirac, M. Lewenstein, K. Mølmer,  and P. Zoller, “Quantum superposition states of Bose-Einstein condensates,” Phys. Rev. A 57, 1208 (1998).
  • Tian and Zoller (2003) L. Tian and P. Zoller, “Quantum computing with atomic Josephson junction arrays,” Phys. Rev. A 68, 042321 (2003).
  • Byrnes et al. (2012) T. Byrnes, K. Wen,  and Y. Yamamoto, “Macroscopic quantum computation using Bose-Einstein condensates,” Phys. Rev. A 85, 040306(R) (2012).
  • Byrnes et al. (2015) T. Byrnes, D. Rosseau, M. Khosla, A. Pyrkov, A. Thomasen, T. Mukai, S. Koyama, A. Abdelrahman,  and E. Ilo-Okeke, “Macroscopic quantum information processing using spin coherent states,” Optics Communications 337, 102 (2015).
  • Luiz et al. (2015) F. S. Luiz, E. I. Duzzioni,  and L. Sanz, “Implementation of quantum logic gates using coupled Bose-Einstein condensates,” Braz. J. Phys. 45, 550 (2015).
  • (8) A. Großardt, “Nonlinear-ancilla aided quantum algorithm for nonlinear Schrödinger equations,” arXiv: 2403.10102.
  • Solenov and Mozyrsky (2011) D. Solenov and D. Mozyrsky, “Cold atom qubits,” Journal of Computational and Theoretical Nanoscience 8, 481 (2011).
  • Amico et al. (2014) L. Amico, D. Aghamalyan, F. Auksztol, H. Crepaz, R. Dumke,  and L. C. Kwek, “Superfluid qubit systems with ring shaped optical lattices,” Sci. Rep. 4, 4298 (2014).
  • Aghamalyan et al. (2016) D. Aghamalyan, N. T. Nguyen, F. Auksztol, K. S. Gan, M. M. Valado, P. C. Condylis, L.-C. Kwek, R. Dumke,  and L. Amico, “An atomtronic flux qubit: a ring lattice of Bose–Einstein condensates interrupted by three weak links,” New J. Phys. 18, 075013 (2016).
  • Ryu et al. (2020) C. Ryu, E .C. Samson,  and M. G. Boshier, “Quantum interference of currents in an atomtronic SQUID,” Nat Commun. 11, 3338 (2020).
  • Kapale and Dowling (2005) K. T. Kapale and J. P. Dowling, “Vortex phase qubit: Generating arbitrary, counterrotating, coherent superpositions in Bose-Einstein condensates via optical angular momentum beams,” Phys. Rev. Lett. 95, 173601 (2005).
  • Ryu et al. (2007) C. Ryu, M. F. Andersen, P. Cladé, V. Natarajan, K. Helmerson,  and W. D. Phillips, “Observation of persistent flow of a Bose-Einstein condensate in a toroidal trap,” Phys. Rev. Lett. 99, 260401 (2007).
  • Ramanathan et al. (2011) A. Ramanathan, K. C. Wright, S. R. Muniz, M. Zelan, W. T. Hill III, C. J. Lobb, K. Helmerson, W. D. Phillips,  and K. C. Gampbell, “Superflow in a toroidal Bose-Einstein condensate: An atom circuit with a tunable weak link,” Phys. Rev. Lett. 106, 130401 (2011).
  • Eckel et al. (2014) S. Eckel, J. G. Lee, F. Jendrzejewski, N. Murray, C. W. Clark, C. J. Lobb, W. D. Phillips, M. Edwards,  and G. K. Campbell, “Hysteresis in a quantized superfluid atomtronic circuit,” Nature 506, 200 (2014).
  • Kim et al. (2018) H. Kim, G. Zhu, J. V. Porto,  and M. Hafezi, “Optical lattice with torus topology,” Phys. Rev. Lett. 121, 133002 (2018).
  • (18) L. Pezzè, K. Xhani, C. Daix, N. Grani, B. Donelli, F. Scazza, D. Hernandez-Rajkov, W. J. Kwon, G. Del Pace,  and G. Roati, “Stabilizing persistent currents in an atomtronic Josephson junction necklace,” arXiv: 2311.05523.
  • Moses et al. (2023) S. A. Moses, C. H. Baldwin, M. S. Allman, R. Ancona, L. Ascarrunz, C. Barnes, J. Bartolotta, B. Bjork, P. Blanchard, M. Bohn, J. G. Bohnet, N. C. Brown, N. Q. Burdick, W. C. Burton, S. L. Campbell, J. P. Campora, C. Carron, J. Chambers, J. W. Chan, Y. H. Chen, A. Chernoguzov, E. Chertkov, J. Colina, J. P. Curtis, R. Daniel, M. DeCross, D. Deen, C. Delaney, J. M. Dreiling, C. T. Ertsgaard, J. Esposito, B. Estey, M. Fabrikant, C. Figgatt, C. Foltz, M. Foss-Feig, D. Francois, J. P. Gaebler, T. M. Gatterman, C. N. Gilbreth, J. Giles, E. Glynn, A. Hall, A. M. Hankin, A. Hansen, D. Hayes, B. Higashi, I. M. Hoffman, B. Horning, J. J. Hout, R. Jacobs, J. Johansen, L. Jones, J. Karcz, T. Klein, P. Lauria, P. Lee, D. Liefer, S. T. Lu, D. Lucchetti, C. Lytle, A. Malm, M. Matheny, B. Mathewson, K. Mayer, D. B. Miller, M. Mills, B. Neyenhuis, L. Nugent, S. Olson, J. Parks, G. N. Price, Z. Price, M. Pugh, A. Ransford, A. P. Reed, C. Roman, M. Rowe, C. Ryan-Anderson, S. Sanders, J. Sedlacek, P. Shevchuk, P. Siegfried, T. Skripka, B. Spaun, R. T. Sprenkle, R. P. Stutz, M. Swallows, R. I. Tobey, A. Tran, T. Tran, E. Vogt, C. Volin, J. Walker, A. M. Zolot,  and J. M. Pino, “A race track trapped-ion quantum processor,” Phys. Rev. X 13, 041052 (2023).
  • Krinner et al. (2022) S. Krinner, N. Lacroix, A. Remm, A. Di Paolo, E. Genois, C. Leroux, C. Hellings, S. Lazar, F. Swiadek, J. Herrmann, G. J. Norris, C. K. Andersen, M. Müller, A. Blais, C. Eichler,  and A. Wallraff, “Realizing repeated quantum error correction in a distance-three surface code,” Nature 605, 669 (2022).
  • Google Quantum AI (2023) Google Quantum AI, “Suppressing quantum errors by scaling a surface code logical qubit,” Nature 614, 676 (2023).
  • (22) D. C. McKay, I. Hincks, E. J. Pritchett, M. Carroll, L. C. G. Govia,  and S. T. Merkel, “Benchmarking quantum processor performance at scale,” arXiv: 2311.05933.
  • Bluvstein et al. (2022) D. Bluvstein, H. Levine, G. Semeghini, T. T. Wang, S. Ebadi, M. Kalinowski, A. Keesling, N. Maskara, H. Pichler, M. Greiner, V. Vuletić,  and M. D. Lukin, “A quantum processor based on coherent transport of entangled atom arrays,” Nature 604, 451 (2022).
  • Graham et al. (2022) T. M. Graham, Y. Song, J. Scott, C. Poole, L. Phuttitarn, K. Jooya, P. Eichler, X. Jiang, A. Marra, B. Grinkemeyer, M. Kwon, M. Ebert, J. Cherek, M. T. Lichtman, M. Gillette, J. Gilbert, D. Bowman, T. Ballance, C. Campbell, E. D. Dahl, O. Crawford, N. S. Blunt, B. Rogers, T. Noel,  and M. Saffman, “Multi-qubit entanglement and algorithms on a neutral-atom quantum computer,” Nature 604, 457 (2022).
  • Coffman et al. (2000) V. Coffman, J. Kundu,  and W. K. Wootters, “Distributed entanglement,” Phys. Rev. A 61, 052306 (2000).
  • (26) X.-L. Zong, H.-H. Yin, W. Song,  and Z.-L. Cao, “Monogamy of quantum entanglement,” arXiv: 2201.00366.
  • Gross (1961) E. P. Gross, “Structure of a quantized vortex in boson systems,” Nuovo Cimento 20, 454 (1961).
  • Pitaevskii (1961) L. P. Pitaevskii, “Vortex lines in an imperfect Bose gas,” Sov. Phys. JETP 13, 451 (1961).
  • Abrams and Lloyd (1998) D. S. Abrams and S. Lloyd, “Nonlinear quantum mechanics implies polynomial-time solution for NP-Complete and #\#P problems,” Phys. Rev. Lett. 81, 3992 (1998).
  • Wu and Niu (2000) B. Wu and Q. Niu, “Nonlinear Landau-Zener tunneling,” Phys. Rev. A 62, 023402 (2000).
  • Riedel et al. (2010) 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).
  • Meyer and Wong (2013) D. A. Meyer and T. G. Wong, “Nonlinear quantum search using the Gross-Pitaevskii equation,” New J. Phys. 15, 063014 (2013).
  • Childs and Young (2016) A. M. Childs and J. Young, “Optimal state discrimination and unstructured search in nonlinear quantum mechanics,” Phys. Rev. A 93, 022314 (2016).
  • Xu et al. (2022) S. Xu, J. Schmiedmayer,  and B. C. Sanders, “Nonlinear quantum gates for a Bose-Einstein condensate,” Phys. Rev. Research 4, 023071 (2022).
  • Deffner (2022) S. Deffner, “Nonlinear speed-ups in ultracold quantum gases,” Europhys. Lett. 140, 48001 (2022).
  • Geller (2023a) M. R. Geller, “Fast quantum state discrimination with nonlinear PTP channels,” Adv. Quantum Technol. , 2200156 (2023a), arXiv: 2111.05977.
  • Erdős and Schlein (2009) L. Erdős and B. Schlein, “Quantum dynamics with mean field interactions: A new approach,” J. Stat. Phys. 134, 859 (2009).
  • Geller (2023b) M. R. Geller, “The universe as a nonlinear quantum simulation: Large n limit of the central spin model,” Phys. Rev. A 108, 042210 (2023b), arXiv: 2112.09005.
  • Guenther et al. (2021) N.-E. Guenther, R. Schmidt, G. M. Bruun, V. Gurarie,  and P. Massignan, “Mobile impurity in a Bose-Einstein condensate and the orthogonality catastrophe,” Phys. Rev. A 103, 013317 (2021).
  • Mueller (2002) E. J. Mueller, “Superfluidity and mean-field energy loops: Hysteretic behavior in Bose-Einstein condensates,” Phys. Rev. A 66, 063603 (2002).
  • Baharian and Baym (2013) S. Baharian and G. Baym, “Bose-Einstein condensates in toroidal traps: Instabilities, swallow-tail loops, and self-trapping,” Phys. Rev. A 87, 013619 (2013).
  • Mielnik (1980) B. Mielnik, “Mobility of nonlinear systems,” J. Math. Phys. 21, 44 (1980).
  • Kitagawa and Ueda (1993) M. Kitagawa and M. Ueda, “Squeezed spin states,” Phys. Rev. A 47, 5138 (1993).
  • (44) S. M. Barnet and S. Croke, “Quantum state discrimination,” arXiv: 0810.1970.
  • Bae and Hwang (2013) J. Bae and W.-Y. Hwang, “Minimum-error discrimination of qubit states: Methods, solutions, and properties,” Phys. Rev. A 87, 012334 (2013).
  • Bae and Kwek (2015) J. Bae and L.-C. Kwek, “Quantum state discrimination and its applications,” J. Phys. A: Math. Theor. 48, 083001 (2015).
  • (47) M. Rouhbakhsh and S. A. Ghoreishi, “Minimum-error discrimination of qubit states revisited,” arXiv: 2108.12299.
  • Helstrom (1976) C. W. Helstrom, Quantum Detection and Estimation Theory (Academic, 1976).
  • (49) S. Aaronson, “NP-complete problems and physical reality,” arXiv: quant-ph/0502072.
  • Bollinger et al. (1989) J. J. Bollinger, D. J. Heinzen, W. M. Itano, S. L. Gilbert,  and D. J. Wineland, “Test of the linearity of quantum mechanics by rf spectroscopy of the 9Be+ ground state,” Phys. Rev. Lett. 63, 1031 (1989).
  • Chupp and Hoare (1990) T. E. Chupp and R. J. Hoare, “Coherence in freely precessing Ne21{}^{21}\mathrm{Ne} and a test of linearity of quantum mechanics,” Phys. Rev. Lett. 64, 2261 (1990).
  • Walsworth et al. (1990) R. L. Walsworth, I. F. Silvera, E. M. Mattison,  and R. F. C. Vessot, “Test of the linearity of quantum mechanics in an atomic system with a hydrogen maser,” Phys. Rev. Lett. 64, 2599 (1990).
  • Majumder et al. (1990) P. K. Majumder, B. J. Venema, S. K. Lamoreaux, B. R. Heckel,  and E. N. Fortson, “Test of the linearity of quantum mechanics in optically pumped 201Hg,” Phys. Rev. Lett. 65, 2931 (1990).
  • Ruskai (1994) M. B. Ruskai, “Beyond strong subadditivity? Improved bounds on the contraction of generalized relative entropy,” Rev. Math. Phys. 6, 1147 (1994).
  • Amico et al. (2005) L. Amico, A. Osterloh,  and F. Cataliotti, “Quantum many particle systems in ring-shaped optical lattices,” Phys. Rev. Lett. 95, 063201 (2005).
  • Moulder et al. (2012) S. Moulder, S. Beattie, R. P. Smith, N. Tammuz,  and Z. Hadzibabic, “Quantized supercurrent decay in an annular Bose-Einstein condensate,” Phys. Rev. A 86, 013629 (2012).
  • Dominy et al. (2016) J. M. Dominy, A. Shabani,  and D. A. Lidar, “A general framework for complete positivity,” Quantum Inf. Process. 15, 465 (2016).
  • Dominy and Lidar (2016) J. M. Dominy and D. A. Lidar, “Beyond complete positivity,” Quantum Inf. Process. 15, 1349 (2016).
  • (59) D.-X. Chen, Y. Zhang, J.-L. Zhao, Q.-C. Wu, Y.-L. Fang, C.-P. Yang,  and F. Nori, “Quantum state discrimination in a PT-symmetric system,” arXiv: 2209.02481.
  • Lieb et al. (2000) E. H. Lieb, R. Seiringer,  and J. Yngvason, “Bosons in a trap: A rigorous derivation of the Gross-Pitaevskii energy functional,” Phys. Rev. A 61, 043602 (2000).
  • Benedikter et al. (2016) N. Benedikter, M. Porta,  and B. Schlein, Effective Evolution Equations from Quantum Dynamics (Springer, Berlin, 2016).