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

Cramér-Rao bound and quantum parameter estimation with non-Hermitian systems

Jianning Li Center for Quantum Sciences and School of Physics, Northeast Normal University, Changchun 130024, China    Haodi Liu Center for Quantum Sciences and School of Physics, Northeast Normal University, Changchun 130024, China    Zhihai Wang Center for Quantum Sciences and School of Physics, Northeast Normal University, Changchun 130024, China    Xuexi Yi [email protected] Center for Quantum Sciences and School of Physics, Northeast Normal University, Changchun 130024, China Center for Advanced Optoelectronic Functional Materials Research, and Key Laboratory for UV-Emitting Materials and Technology of Ministry of Education, Northeast Normal University, Changchun 130024, China
Abstract

The quantum Fisher information constrains the achievable precision in parameter estimation via the quantum Cramér-Rao bound, which has attracted much attention in Hermitian systems since the 60s of the last century. However, less attention has been paid to non-Hermitian systems. In this Letter, working with different logarithmic operators, we derive two previously unknown expressions for quantum Fisher information, and two Cramér-Rao bounds lower than the well-known one are found for non-Hermitian systems. These lower bounds are due to the merit of non-Hermitian observable and it can be understood as a result of extended regimes of optimization. Two experimentally feasible examples are presented to illustrate the theory, saturation of these bounds and estimation precisions beyond the Heisenberg limit are predicted and discussed. A setup to measure non-Hermitian observable is also proposed.

Introduction.—Quantum parameter estimation helstrom67 ; helstrom76 ; holevo82 ; degen17 ; pezze18 aims at measuring the value of a continuous parameter θ\theta encoded in the state ρ(θ)\rho(\theta) of a quantum system. It gives better precision than the same measurement performed in a classical framework and plays a crucial role in the advancement of physics. The estimation process generally consists of two steps. In the first step, the state ρ(θ)\rho({\theta}) is prepared and measured. A simple way to prepare the state ρ(θ)\rho(\theta) is to evolve a reference initial state ρ0\rho_{0} under a signal Hamiltonian HθH_{\theta} that encodes the parameter θ\theta, ρt(θ)=eiHθtρ0eiHθt.\rho_{t}(\theta)=e^{-iH_{\theta}t}\rho_{0}e^{iH_{\theta}t}. In the second step, the estimation of θ\theta is obtained by data processing the measurement outcomes, aiming at the smallest uncertainty Δθ\Delta\theta given finite resources such as time and number of particles. The uncertainty is bounded by the Cramér-Rao bound (CRB), which expresses a lower limit on the variance of unbiased estimations, stating that the variance of any such estimation is at least as high as the inverse of the Fisher information fisher25 ; matson06 .

The quantum Cramér-Rao bound works in general with Hermitian quantum systems frieden90 ; braunstein94 ; liu20 in order to meet the requirement of quantum mechanics. However, recent studies found that non-Hermitian systems with unbroken PT symmetries also possess a real spectrum bender98 . In fact, non-Hermiticity is ubiquitous in the quantum world bender07 ; guo09 , including systems with gain and/or loss konotop16m ; feng17 ; ganainy19 ; miri19 ; ozdemir19 , many-body localization and dynamical stability of non-Hermitian systems hamazaki19 , as well as sensors designed with non-Hermitian systems berry04 ; heiss12 .

These facts give rise to a question that in non-Hermitian quantum systems moiseyev11 what is the Cramér-Rao bound? This question was answered about 5 decades ago by Yuen and Lax in Ref. yuen73 , where the authors proposed an idea to estimate a complex parameter by measuring a non-Hermitian observable. Unfortunately, this theory was connected the complex parameter with the estimator by assuming that the average of the estimator is exactly the estimate of the parameter. This assumption would lead to a flaw that the error propagation function is unity, resulting in a less regime of non-Hermitian observables for optimization and experiments. Consider that the study of non-Hermitian system and their unique properties have attracted fast growing interest in the last two decades, revisiting the quantum Cramér bound and its consequent estimation theory with current technologies is highly desired for non-Hermitian systems. We should address that there are estimation protocols (or sensors) based on non-Hermitian system recently wiersig14 ; liu16 ; chen17 ; hodaei17 ; djorwe19 ; mao20 , but all analyses so far are based on either the Fisher information resulting from the Hermitian quantum Cramér-Rao bound, or the properties of the exceptional points.

In this Letter, we first develop an uncertainty relation for non-Hermitian operators, then we derive two previously unknown expressions for quantum Fisher information with different logarithmic derivatives. Two quantum Cramér-Rao bounds for non-Hermitian systems are defined. We found that the Fisher information is significantly increased due to the use of non-Hermitian operators, in particular for systems in mixed states. This is in contrary to the results of Fisher information with Hermitian symmetric logarithmic derivatives. Saturation of the two bounds is analyzed and the optimal measurement to attain the bounds is derived. We elucidate the feature of non-Hermitian quantum Fisher information with GHZ states of trapped ions pezze18 ; leibried05 and the phase estimation setup with Mach-Zehnder interferometer. Comparison with the situation of non-Hermitian signal Hamiltonian is also presented and discussed.

Non-Hermitian uncertainty relation and Quantum Cramér bound.— Higher estimation precision demands more resources. The trade-off between the precision and the resources required is determined by the uncertainty principle, which constrains to what extent complementary variables maintain their averaged values and leads to the Heisenberg limit caves81 ; chin12 ; jarzyna15 ; giovannetti11 ; luis17 ; bai19 in quantum parameter estimation. One might wonder whether this is the case and how the uncertainty relation changes in non-Hermitian quantum mechanics moiseyev11 .

To explore the possible change of the uncertainty relation and introduce the variance for non-Hermitian systems, we introduce two operators AA and BB, which are linear but not necessarily Hermitian. Defining ΔA=AA\Delta A=A-\langle A\rangle, ΔB=BB\Delta B=B-\langle B\rangle and O=χΔA+iΔBO=\chi\Delta A+i\Delta B with χ\chi a real parameter, i2=1i^{2}=-1 and Z\langle Z\rangle being the expectation value of operator ZZ, we have OO0.\langle O^{\dagger}O\rangle\geq 0. Simple algebra yields supp , ΔAΔAΔBΔB14|C|2,\langle\Delta A^{\dagger}\Delta A\rangle\langle\Delta B^{\dagger}\Delta B\rangle\geq\frac{1}{4}|\langle C\rangle|^{2}, where CC was defined by C=i(ABBA)i[A,B]a.C=i(A^{\dagger}B-B^{\dagger}A)\equiv i[A,B]_{a}. Operator CC is a Hermitian operator regardless of AA and BB being Hermitian or not. [X,Y]a[X,Y]_{a} defines an abnormal commutation for operator XX and YY, which can be applied to determine whether a time-independent operator is a constant of motion for systems governed by non-Hermitian Hamiltonian rivero20 .

We define σA2ΔAΔA\sigma_{A}^{2}\equiv\langle\Delta A^{\dagger}\Delta A\rangle as the variance for non-Hermitian operator AA, which reduces to the traditional variance when AA is Hermitian. This uncertainty relation as well as that we will present in the following also hold for unitary operators AA and BB bong18 ; yu19 . In non-Hermitian quantum systems, the expectation value of dynamical variable AA might take complex values, A=|A|eiα\langle A\rangle=|\langle A\rangle|e^{i\alpha}. The absolute value |A||\langle A\rangle| and its phase α\alpha are both measurable quantities. For example, in scattering experiments the peaks in the cross sections are obtained when the projectiles have energy which is equal to the absolute value of the energy, not the real part of the energy moiseyev11 . This complex expectation of non-Hermitian operator is equal to the weak value of the positive-semidefinite part of the operator multiplied by a known complex number pati15 ; nirala19 ; sahoo20 .

A stronger uncertainty relations for non-Hermitian operators AA and BB follows from the Schwarz inequality, F|FG|G|F|G|2\langle F|F\rangle\langle G|G\rangle\geq|\langle F|G\rangle|^{2} with |F=ΔA|Ψ|F\rangle=\Delta A|\Psi\rangle and |G=ΔB|Ψ|G\rangle=\Delta B|\Psi\rangle and |Ψ|\Psi\rangle being an arbitrary state of system,

ΔAΔAΔBΔB|ABAB|2.\displaystyle\langle\Delta A^{\dagger}\Delta A\rangle\langle\Delta B^{\dagger}\Delta B\rangle\geq|\langle A^{\dagger}B\rangle-\langle A^{\dagger}\rangle\langle B\rangle|^{2}. (1)

This is a slightly stronger inequality for the variance of operators AA and BB, which together with error propagation supp

(Δθ)2=ΔAΔAA/θA/θ(\Delta\theta)^{2}=\frac{\langle\Delta A^{\dagger}\Delta A\rangle}{\partial\langle A^{\dagger}\rangle/\partial\theta\cdot\partial\langle A\rangle/\partial\theta} (2)

leads to a quantum Cramér-Rao bound for non-Hermitian system,

(Δθ)21FnH(1),(\Delta\theta)^{2}\geq\frac{1}{F_{nH}^{(1)}}, (3)

where FnH(1)=Tr(ρL(1)L(1))F_{nH}^{(1)}=\text{Tr}(\rho L^{(1)\dagger}L^{(1)}) is defined as non-Hermitian quantum Fisher information with helstrom73ijtp ; supp

ρθ=12(L(1)ρ+ρL(1)),\frac{\partial\rho}{\partial\theta}=\frac{1}{2}(L^{(1)}\rho+\rho L^{(1)\dagger}), (4)

where θ\theta is the real parameter to be estimated based on density matrix ρ(θ)\rho(\theta) and Hermitian operator AA. The slightly stronger uncertainty relation Eq.(1) reduces to the Robertson-Schrödinger uncertainty relation when both AA and BB are Hermitian and returns to the unitary uncertainty relation bong18 ; yu19 with both AA and BB being unitary. Eq.(S15) was derived with an assumption that A=AA^{\dagger}=A, but there is no requirement for LL. This indicates that Eq.(3) holds for both Hermitian and non-Hermitian LL. In fact, when LL is Hermitian, the result reduces to the widely used quantum Fisher information (denoted by FHF_{H} hereafter) in the literature, while non-Hermitian LL would lead to an enhanced precision for parameter estimations as shown below.

For non-Hermitian operator AA, we obtain the other quantum Cramér-Rao bound leading to non-Hermitian quantum information FnH(2)=Tr(ρL(2)L(2)),F_{nH}^{(2)}=\text{Tr}(\rho L^{(2)\dagger}L^{(2)}), which has the same expression as in Eq.(3) but with L(2)L^{(2)} instead of L(1)L^{(1)} helstrom73ijtp ; supp ,

ρθ=L(2)ρ=ρL(2).\frac{\partial\rho}{\partial\theta}=L^{(2)}\rho=\rho L^{(2)\dagger}. (5)

There are two fundamental differences between Hermitian and non-Hermitian system in the quantum Cramér-Rao bounds defined by FnH(1)F_{nH}^{(1)} and FnH(2)F_{nH}^{(2)}, which deserve to be outlined. Firstly, in Fisher information FnH(1)F_{nH}^{(1)}, the observable AA is Hermitian but the so-call symmetric logarithmic derivative LL might be not. While both AA and LL in FnH(2)F_{nH}^{(2)} are not Hermitian. The state ρ(θ)\rho(\theta) that encodes the parameter is Hermitian, however the signal Hamiltonian HθH_{\theta} might not be Hermitian, i.e., HθHθH_{\theta}\neq H_{\theta}^{\dagger}. Secondly, in non-Hermitian system, the Fisher information is given by FnH=Tr(ρL(y)L(y)),y=1,2F_{nH}=\text{Tr}(\rho L^{(y)\dagger}L^{(y)}),\,y=1,2 with ρθ=L(2)ρ=ρL(2)\frac{\partial\rho}{\partial\theta}=L^{(2)}\rho=\rho L^{(2)\dagger} and ρθ=12(L(1)ρ+ρL(1))\frac{\partial\rho}{\partial\theta}=\frac{1}{2}(L^{(1)}\rho+\rho L^{(1)\dagger}) instead of ρθ=12(Lρ+ρL)\frac{\partial\rho}{\partial\theta}=\frac{1}{2}(L\rho+\rho L) in Hermitian system. As we will show later, FnH(1)F_{nH}^{(1)} recovers the well-known expression of quantum Fisher information FHF_{H} for Hermitian system while FnH(2)F_{nH}^{(2)} can not. Whereas the Cramér-Rao bound defined by FnH(2)F_{nH}^{(2)} can be saturated but that by FnH(1)F_{nH}^{(1)} could not supp .

Fisher information of non-Hermitian system.— One of the main quests in quantum parameter estimation is to find out the highest achievable precision with given resources and design schemes that attain that precision. In general, looking for the optimal resources and schemes is difficult since one needs to optimize over the input state that encodes the parameter, the measurement that is performed at the output, and the estimator that assigns a parameter value to a given measurement outcome. One of the popular ways to obtain useful bounds in quantum parameter estimations, without the need for cumbersome optimization, is to use the quantum Fisher information. In the following, we will give an explicit expression for the non-Hermitian quantum Fisher information FnH(1)F_{nH}^{(1)} and FnH(2)F_{nH}^{(2)} in terms of the eigenstates and eigenvalues of the encoding density matrix ρ(θ)\rho(\theta).

Consider a NN-dimensional quantum system. The state of the system ρ(θ)\rho(\theta) is parameterized by the parameter θ\theta under estimation. Without loss of generality, we assume the density matrix may not be of full rank and the spectral decomposition of the density operator ρ(θ)\rho(\theta) is,

ρ(θ)=iMpi(θ)|ϕi(θ)ϕi(θ)|,MN.\rho(\theta)=\sum_{i}^{M}p_{i}(\theta)|\phi_{i}(\theta)\rangle\langle\phi_{i}(\theta)|,\,M\leq N. (6)

Here pi(θ),i=1,2,,Mp_{i}(\theta),i=1,2,...,M are required to be positive for any value of θ\theta and iMpi(θ)=1\sum_{i}^{M}p_{i}(\theta)=1, since pi(θ)p_{i}(\theta) is the probability of the system in state |ϕi(θ)|\phi_{i}(\theta)\rangle. The following normalization conditions ϕi(θ)|ϕj(θ)=δij\langle\phi_{i}(\theta)|\phi_{j}(\theta)\rangle=\delta_{ij} for all ii and jj are imposed. From the definition of Fisher information, we can find that,

FnH(y)=i=1Mj=1Npi(L(y))ijLji(y),y=1,2.F_{nH}^{(y)}=\sum_{i=1}^{M}\sum_{j=1}^{N}p_{i}(L^{(y)\dagger})_{ij}L^{(y)}_{ji},\,y=1,2.

Noticing the derivation for the two Fisher information FnH(1)F_{nH}^{(1)} and FnH(2)F_{nH}^{(2)} is different, we discuss it in the following separately.

We start our derivation of FnH(1)F_{nH}^{(1)} by considering a special case that L(1)=eiβL(1)L^{(1)\dagger}=e^{i\beta}L^{(1)} with β\beta a real parameter. This assumption together with Eq.(S15) leads to,

Lij(1)=ϕi|L(1)|ϕi=2(θρ)ijpj+pieiβL^{(1)}_{ij}=\langle\phi_{i}|L^{(1)}|\phi_{i}\rangle=\frac{2(\partial_{\theta}\rho)_{ij}}{p_{j}+p_{i}e^{i\beta}} (7)

and Lij(1)=eiβLij(1).L^{(1)\dagger}_{ij}=e^{i\beta}L^{(1)}_{ij}. A notation (θρ)ij=(ρ(θ)θ)ij(\partial_{\theta}\rho)_{ij}=\left(\frac{\partial\rho(\theta)}{\partial\theta}\right)_{ij} was used. Clearly, L(1)L(1)L^{(1)}\neq L^{(1)\dagger} as we expected. It is worth noticing that Lij(1)L^{(1)}_{ij} is in principle supported by the full Hilbert space, not only that spanned by the eigenstates of the density matrix. Thus the value of Lij(1)L^{(1)}_{ij} for i,j>Ni,j>N can not be established by the above equations. However, these terms play an important role in the quantum Fisher information. This problem can be solved by the use of completeness relation, i=1N|ϕiϕi|=1\sum_{i=1}^{N}|\phi_{i}\rangle\langle\phi_{i}|=1, i.e., i=M+1N|ϕiϕi|=1i=1M|ϕiϕi|\sum_{i=M+1}^{N}|\phi_{i}\rangle\langle\phi_{i}|=1-\sum_{i=1}^{M}|\phi_{i}\rangle\langle\phi_{i}|. We finally arrive at supp ,

FnH(1)=i=1M2(θpi)2pi(1+cosβ)+i=1M4piθϕi|θϕi+i=1Mj=1M(4pi(pipj)2pi2+pj2+2pipjcosβ4pi)θϕi|ϕjϕj|θϕi,\displaystyle F_{nH}^{(1)}=\sum_{i=1}^{M}\frac{2(\partial_{\theta}p_{i})^{2}}{p_{i}(1+\cos\beta)}+\sum_{i=1}^{M}4p_{i}\langle\partial_{\theta}\phi_{i}|\partial_{\theta}\phi_{i}\rangle+\sum_{i=1}^{M}\sum_{j=1}^{M}\left(\frac{4p_{i}(p_{i}-p_{j})^{2}}{p_{i}^{2}+p_{j}^{2}+2p_{i}p_{j}\cos\beta}-4p_{i}\right)\langle\partial_{\theta}\phi_{i}|\phi_{j}\rangle\langle\phi_{j}|\partial_{\theta}\phi_{i}\rangle, (8)

where |θϕ|ϕθ.|\partial_{\theta}\phi\rangle\equiv\frac{\partial|\phi\rangle}{\partial\theta}. This is one of the main results of this Letter. For pure state, FnH(1)F_{nH}^{(1)} reduce to FnH(1)=FH=4θϕ|θϕ4θϕ|ϕϕ|θϕ.F_{nH}^{(1)}=F_{H}=4\langle\partial_{\theta}\phi|\partial_{\theta}\phi\rangle-4\langle\partial_{\theta}\phi|\phi\rangle\langle\phi|\partial_{\theta}\phi\rangle. This is the widely used quantum fisher information in literature. For mixed states with β0\beta\rightarrow 0, FnH(1)F_{nH}^{(1)} reduces to the Hermitian Fisher information, whereas for mixed states with βπ\beta\rightarrow\pi,

FnH(1)i=1M4piθϕi|θϕi+Fc,F_{nH}^{(1)}\simeq\sum_{i=1}^{M}4p_{i}\langle\partial_{\theta}\phi_{i}|\partial_{\theta}\phi_{i}\rangle+F_{c}, (9)

where Fc=i=1M2(θpi)2pi(1+cosβ)|βπ.F_{c}=\sum_{i=1}^{M}\frac{2(\partial_{\theta}p_{i})^{2}}{p_{i}(1+\cos\beta)}|_{\beta\rightarrow\pi}. FcF_{c} approaches to infinity in the limit of βπ\beta\rightarrow\pi. This is not physical, since LijL_{ij} in Eq. (7) can not be established when βπ\beta\rightarrow\pi and i=ji=j. In the following discussion, we will focus on the case that pip_{i} are θ\theta-independent, such that Fc=0.F_{c}=0.

The quantum information FnH(2)F_{nH}^{(2)} can be calculated by the same technique. Different from the case of FnH(1)F_{nH}^{(1)}, here (L(2))ij=(θρ)ijpi(L^{(2)\dagger})_{ij}=\frac{(\partial_{\theta}\rho)_{ij}}{p_{i}}, Lji(2)=(θρ)jipiL^{(2)}_{ji}=\frac{(\partial_{\theta}\rho)_{ji}}{p_{i}} for pi0.p_{i}\neq 0. Although both AA and LL are not Hermitian, the Fisher information is real and given by supp ,

FnH(2)\displaystyle F_{nH}^{(2)} =\displaystyle= i=1M(θpi)2pi+i=1Mpiθϕi|θϕi\displaystyle\sum_{i=1}^{M}\frac{(\partial_{\theta}p_{i})^{2}}{p_{i}}+\sum_{i=1}^{M}p_{i}\langle\partial_{\theta}\phi_{i}|\partial_{\theta}\phi_{i}\rangle
+\displaystyle+ i=1Mj=1Mpj(pj2pi)piθϕi|ϕjϕj|θϕi.\displaystyle\sum_{i=1}^{M}\sum_{j=1}^{M}\frac{p_{j}(p_{j}-2p_{i})}{p_{i}}\langle\partial_{\theta}\phi_{i}|\phi_{j}\rangle\langle\phi_{j}|\partial_{\theta}\phi_{i}\rangle.

This is the other main result of this Letter. The quantum Fisher information FnH(2)F_{nH}^{(2)} for pure state |ϕ|\phi\rangle reduces to

FnH(2)=θϕ|θϕθϕ|ϕϕ|θϕ,F_{nH}^{(2)}=\langle\partial_{\theta}\phi|\partial_{\theta}\phi\rangle-\langle\partial_{\theta}\phi|\phi\rangle\langle\phi|\partial_{\theta}\phi\rangle, (11)

which is 1/41/4 times smaller than FHF_{H} in Hermitian system for pure stats. However, this is not the case for mixed states as shown in the examples below. The saturation of these bounds and the experimental measurement of non-Hermitian operator will be discussed briefly at the end of this Letter.

Examples.— Considering the Hermitian quantum Fisher information FHF_{H} and the non-Hermitian quantum information FnH(1)F_{nH}^{(1)} and FnH(2)F_{nH}^{(2)} are almost same for pure stats, i.e., FH=FnH(1)F_{H}=F_{nH}^{(1)} and FH=4FnH(2)F_{H}=4F_{nH}^{(2)} for pure states, we focus on the case of mixed states in the undergoing examples. To simplify the expression, we consider a type of simplest mixed states—it is constructed by only two orthogonal pure states with weight p1p_{1} and p2,p_{2}, and p1+p2=1.p_{1}+p_{2}=1.

We first consider the maximally entangled states (or Schrödinger cat state) |ϕ1,2=12(|aN±|bN)|\phi_{1,2}\rangle=\frac{1}{\sqrt{2}}(|a\rangle^{\otimes N}\pm|b\rangle^{\otimes N}) as the input state, which has been created with up to N=6N=6 B9e+{}^{9}Be^{+} ions leibried05 and N=14N=14 C40a+{}^{40}Ca^{+} ions monz11 in a linear Paul trap. The encoding operator is a spin rotation defined by U(θ)=eiθJzU(\theta)=e^{-i\theta J_{z}} with Hermitian signal Hamiltonian Jz=j=1Nsz(j)J_{z}=\sum_{j=1}^{N}s_{z}^{(j)}. The case of non-Hermitian signal Hamiltonian will be discussed at the end of examples. The states that encode the parameter θ\theta to construct the mixed states are,

|ϕ1,2(θ)=12(ei2Nθ|aN±ei2Nθ|bN).|\phi_{1,2}(\theta)\rangle=\frac{1}{\sqrt{2}}(e^{\frac{i}{2}N\theta}|a\rangle^{\otimes N}\pm e^{-\frac{i}{2}N\theta}|b\rangle^{\otimes N}).

The mixed state is ρ=i=12pi|ϕi(θ)ϕi(θ)|.\rho=\sum_{i=1}^{2}p_{i}|\phi_{i}(\theta)\rangle\langle\phi_{i}(\theta)|. Notice that pi,i=1,2p_{i},\,i=1,2 are θ\theta-independent. The rotation U(θ)=eiθJzU(\theta)=e^{-i\theta J_{z}} generates a relative phase NθN\theta between states |aN|a\rangle^{\otimes N} and |bN|b\rangle^{\otimes N}. Straightforward calculation gives the Hermitian Fisher information FHF_{H} and the non-Hermitian Fisher information FnH(1)F_{nH}^{(1)} and FnH(2)F_{nH}^{(2)} (setting p1=pp_{1}=p and βπ\beta\rightarrow\pi) supp ,

FH\displaystyle F_{H} =\displaystyle= (14p(1p))N2,\displaystyle\left(1-4p(1-p)\right)N^{2},
FnH(1)\displaystyle F_{nH}^{(1)} =\displaystyle= N2,FnH(2)=(2p1)24p(1p)N2.\displaystyle N^{2},\quad F_{nH}^{(2)}=\frac{(2p-1)^{2}}{4p(1-p)}N^{2}. (12)

The results are shown in Fig.1. We find that the non-Hermitian Fisher information are always larger than the Hermitian Fisher information, except the points p=0,0.5,1.p=0,0.5,1.

Refer to caption
Figure 1: Quantum Fisher information FHF_{H}, FnH(1)F_{nH}^{(1)} and FnH(2)F_{nH}^{(2)}. The Fisher information was plotted in units of N2N^{2}. pp is the eigenvalue of the density matrix.

It is worth addressing that at points p=0,1p=0,1, the state is pure, so the Fisher information should be calculated by the formula of pure states, i.e., FH=FnH(1)=N2F_{H}=F_{nH}^{(1)}=N^{2}, and FnH(2)=0.25N2.F_{nH}^{(2)}=0.25N^{2}. As pp approach to 11 and 0, FnH(2)F_{nH}^{(2)} tends to infinity, manifesting itself as a witness of transition from mixed states to pure states. The bound defined by FnH(1)F_{nH}^{(1)} and FnH(2)F_{nH}^{(2)} can be saturated by carefully chosen measurements. For details, see Supplemental Material supp .

The second example we will show is the phase estimation interferometric schemes caves81 ; jarzyna12 , which works with coherent and squeezed vacuum states interfered at the beam-splitter of the Mach-Zehnder interferometer. Let aa and bb be the anihilation operators of the two modes. |α=exp(αbαb)|0b|\alpha\rangle=\exp(\alpha b^{\dagger}-\alpha^{*}b)|0\rangle_{b} is a coherent state of mode bb and |r=exp[12ra212r(a)2]|0a|r\rangle=\exp[\frac{1}{2}r^{*}a^{2}-\frac{1}{2}r(a^{\dagger})^{2}]|0\rangle_{a} is a squeezed vacuum state of mode aa with squeezing parameter rr. After the input state supp has evolved through a beam splitter defined by Bπ=exp[i2π(ab+ab)]B_{\pi}=\exp[-\frac{i}{2}\pi(ab^{\dagger}+a^{\dagger}b)], the estimated parameter θ\theta is encoded into the state by relative phase shift operator U(θ)=exp[iθH],U(\theta)=\exp[-i\theta H], where HH is the signal Hamiltonian of mode aa, H=aaH=a^{\dagger}a. For the other beam splitter, see Supplemental Materials supp . We consider the following states that encode the parameter θ\theta, ρ=i=12pi|ϕi(θ)ϕi(θ)|\rho=\sum_{i=1}^{2}p_{i}|\phi_{i}(\theta)\rangle\langle\phi_{i}(\theta)| with

|ϕ1,2(θ)\displaystyle|\phi_{1,2}(\theta)\rangle =\displaystyle= U(θ)BS(r)D(α)|0a|X1,2b,\displaystyle U(\theta)BS(r)D(\alpha)|0\rangle_{a}\otimes|X_{1,2}\rangle_{b}, (13)

where |X1b=|0b|X_{1}\rangle_{b}=|0\rangle_{b}, |X2b=b|0b,|X_{2}\rangle_{b}=b^{\dagger}|0\rangle_{b}, and p1+p2=1.p_{1}+p_{2}=1. Tedious but straightforward calculations show that (set p1=pp_{1}=p) supp ,

FH\displaystyle F_{H} =\displaystyle= 4|α|2(4p26p+3),\displaystyle 4|\alpha|^{2}(4p^{2}-6p+3),
FnH(1)\displaystyle F_{nH}^{(1)} =\displaystyle= 4|α|4+(2016p)|α|2+4(1p),\displaystyle 4|\alpha|^{4}+(20-16p)|\alpha|^{2}+4(1-p),
FnH(2)\displaystyle F_{nH}^{(2)} =\displaystyle= 2p32p2+1p(1p)|α|2.\displaystyle\frac{2p^{3}-2p^{2}+1}{p(1-p)}|\alpha|^{2}. (14)
Refer to caption
Figure 2: Quantum information as a function of photon number |α|2|\alpha|^{2} and the eigenvalue of density matrix pp.

We have performed numerical calculations for the Fisher information. The results are shown in Fig. 2. We find that the two non-Hermitian quantum Fisher information favor p0p\sim 0. This can be understood as that squeezed Fock states benefit the parameter estimation more than the squeezed vacuum state. FnH(1)F_{nH}^{(1)} with term |α|4|\alpha|^{4} bounds the estimation precision beyond the Heisenberg limit of |α|2|\alpha|^{2}.

One might wonder what is the difference between the non-Hermitian Fisher information and the Hermitian Fisher information with non-Hermitian signal Hamiltonian. To answer this question, we replace the Hermitian operator JzJ_{z} by Jz(1iγ)J_{z}(1-i\gamma) supp in the first example to calculate the Hermitian Fisher information FHF_{H}, and the result will be denoted by FHnhsF_{H}^{nhs}. We compare FHnhsF_{H}^{nhs} with the Hermitian Fisher information FHF_{H} with Hermitian signal Hamiltonian JzJ_{z}. In other words, the comparison is conducted between the Hermitian Fisher information with Hermitian signal and non-Hermitian signal Hamiltonian for pure states. This comparison together with what we had in the two examples would show the difference between the Hermitian and non-Hermitian Fisher information. Calculation supp finds,

FH\displaystyle F_{H} =\displaystyle= N2,\displaystyle N^{2},
FHnhs\displaystyle F_{H}^{nhs} =\displaystyle= |N¯|2[1tanh2(γNθ)],\displaystyle{|\bar{N}|^{2}}\left[\frac{}{}1-\tanh^{2}(\gamma N\theta)\right], (15)

where N¯=N(1iγ).\bar{N}=N(1-i\gamma). Numerical results are given in Fig.3. We find FHnhsF_{H}^{nhs} is smaller than FHF_{H} for almost all γ\gamma and θ\theta, except γ\gamma and θ\theta very close to zero. The difference between FHnhsF_{H}^{nhs} and FHF_{H} can be interpreted by a residual dependence of the state on the parameter under estimation apart from the statistical manifold of states. This happens, for example, when the eigenstates of the density matrix have a parameter-dependent normalization seveso17 .

Refer to caption
Figure 3: Hermitian quantum Fisher information FHF_{H} (yellow-mesh) and the non-Hermitian Fisher information FHnhsF_{H}^{nhs}(blue-surf) as a function of estimation parameter θ\theta and the loss/gain rate γ\gamma. θ\theta is chosen in units of π\pi, and N=4.N=4.

In the second example, we take H¯=H(1iγ)\bar{H}=H(1-i\gamma) instead of HH as the signal Hamiltonian to account the effect of decoherence. FHnhsF_{H}^{nhs} is calculated with H¯\bar{H}, while the Hermitian Fisher information FHF_{H} is calculated with HH,

FH\displaystyle F_{H} =\displaystyle= 4|α|2,\displaystyle 4|\alpha|^{2},
FHnhs\displaystyle F_{H}^{nhs} =\displaystyle= 4|α|2(1+γ2)e2γθ.\displaystyle 4|\alpha|^{2}(1+\gamma^{2})e^{-2\gamma\theta}. (16)

The scaling of Fisher information with the particle number remains unchanged, indicating there is no significant enhancement to parameter estimation in this example.

Saturation of the bounds and feasible experiments to measure an non-Hermitian variable.— To saturate the bounds, the optimal measurement AoptA^{opt} has to meet Aopt=γLA^{opt}=\gamma L and (Aopt)L=LAopt(A^{opt})^{\dagger}L=L^{\dagger}A^{opt} simultaneously supp . These requirements can not be satisfied for bound 1/FnH(1)1/F_{nH}^{(1)}, as AA is Hermitian but LL is not as we stated earlier. Even if we lift the constrain on the Hermiticity of AA, the bound 1/FnH(1)1/F_{nH}^{(1)} can not be saturated too due to the requirement of β=π\beta=\pi. This claim is confirmed by our numerical simulation, see Supplemental Material supp . The situation is different for bound 1/FnH(2)1/F_{nH}^{(2)}. Simple analyses show that the measurement variable

Aijoptϕi|Aopt|ϕj=γ(θρ)ijpj\displaystyle A^{opt}_{ij}\equiv\langle\phi_{i}|A^{opt}|\phi_{j}\rangle=\gamma\frac{(\partial_{\theta}\rho)_{ij}}{p_{j}} (17)

with real γ\gamma can saturate the bound supp . For the example with trapped ions, AoptA^{opt} can be measure in a single-shot experiment in an interferometer setup supp .

Conclusions.—The framework of quantum mechanics in which observable are associated with Hermitian operators is too narrow to discuss parameter estimation. Considering in the past two decades the non-Hermitian physics has attracted fast growing interest in various field of research, we first derived an uncertainty relation for non-Hermitian operators, then we deduce a previously unknown expression for non-Hermitian Fisher information. Two Cramér-Rao bounds that in some cases one of them, and sometimes both of them, are superior to the previous result are found. The saturation of these bounds is analysed and the optimal measurement to attain the bounds are given. The theory was illustrated with two experimentally feasible systems. The setup to measure non-Hermitian observables is also proposed.

We thank Xiaoming Lu for helpful discussions. This work was supported by the National Natural Science Foundation of China (NSFC) under Grants No. 11775048, and No. 12047566.

References

  • (1) C. Helstrom, Minimum mean-squared error of estimates in quantum statistics, Phys. Lett. A 25, 101 (1967).
  • (2) C. W. Helstrom, Quantum Detection and Estimation Theory, (Academic Press, New York, 1976).
  • (3) A. S. Holevo, Probabilistic and Statistical Aspect of Quantum Theory, (North-Holland, Amsterdam, 1982).
  • (4) C. L. Degen, F. Reinhard, and P. Cappellaro, Quantum sensing, Rev. Mod. Phys. 89, 035002(2017).
  • (5) Luca Pezzè, Augusto Smerzi, Markus K. Oberthaler, Roman Schmied and Philipp Treutlein, Quantum metrology with nonclassical states of atomic ensembles, Rev. Mod. Phys. 90, 035005 (2018).
  • (6) R. A. Fisher, Theory of Statistical Estimation, Proc. Camb. Soc. 22, 700 (1925).
  • (7) Charles L. Matson and Alim Haji, Biased Cramér-Rao lower bound calculations for inequality-constrained estimators, J Opt Soc Am A 23,2702 (2006).
  • (8) B. Roy Frieden, Fisher information, disorder, and the equilibrium distributions of physics, Phys. Rev. A 41, 4265(1990).
  • (9) S. L. Braunstein, and C. M. Caves, Statistical distance and the geometry of quantum states, Phys. Rev. Lett. 72, 3439(1994).
  • (10) Jing Liu, Haidong Yuan, Xiao-Ming Lu, Xiaoguang Wang, Quantum Fisher information matrix and multiparameter estimation, J. Phys. A: Math. Theor. 53, 023001 (2020).
  • (11) Carl M. Bender and Stefan Boettcher, Real Spectra in Non-Hermitian Hamiltonians Having PT Symmetry, Phys. Rev. Lett. 80, 5243 (1998).
  • (12) C. M. Bender, D. C. Brody, H. F. Jones, and B. K. Meister, Faster than Hermitian Quantum Mechanics, Phys. Rev. Lett. 98, 040403 (2007).
  • (13) A. Guo, G. J. Salamo, D. Duchesne, R. Morandotti, M. Volatier-Ravat, V. Aimez, G. A. Siviloglou, and D. N. Christodoulides, Observation of PT-Symmetry Breaking in Complex Optical Potentials, Phys. Rev. Lett. 103, 093902 (2009).
  • (14) V. V. Konotop, J. Yang, and D. A. Zezyulin, Nonlinear Waves in PT-Symmetric Systems, Rev. Mod. Phys. 88, 035002 (2016).
  • (15) L. Feng, R. El-Ganainy, and L. Ge, Non-Hermitian Photonics Based on Parity-Time Symmetry, Nat. Photon. 11, 752 (2017).
  • (16) R. El-Ganainy, K. G. Makris, M. Khajavikhan, Z. H. Musslimani, S. Rotter, and D. N. Christodoulides, Non-Hermitian Physics and PT Symmetry, Nat. Phys. 14, 11 (2018).
  • (17) M.-A. Miri and A. Alú, Exceptional Points in Optics and Photonics, Science 363, eaar7709 (2019).
  • (18) S. K. Özdemir, S. Rotter, F. Nori, and L. Yang, Parity-Time Symmetry and Exceptional Points in Photonics, Nat. Mater. 18, 783 (2019).
  • (19) R. Hamazaki, K. Kawabata, and M. Ueda, Non-Hermitian Many-Body Localization, Phys. Rev. Lett. 123, 090603 (2019).
  • (20) M. V. Berry, Physics of Nonhermitian Degeneracies, Czech. J. Phys. 54, 1039 (2004).
  • (21) W. D. Heiss, The Physics of Exceptional Points, J. Phys. A 45, 444016 (2012).
  • (22) N. Moiseyev, Non-Hermitian Quantum Mechanics, (Cambridge University Press, Cambridge, 2011).
  • (23) H. P. Yuen and M. Lax, Multiple-Parameter Quantum Estimation and Measurement of Nonselfadjoint Observables, IEEE Trans. Inf. Theory 19 740 (1973).
  • (24) J. Wiersig, Enhancing the Sensitivity of Frequency and Energy Splitting Detection by Using Exceptional Points: Application to Microcavity Sensors for Single-Particle Detection, Phys. Rev. Lett. 112, 203901 (2014).
  • (25) Z.-P. Liu, J. Zhang, S. K. Özdemir, B. Peng, H. Jing, X.-Y. Lü, C.-W. Li, L. Yang, F. Nori, and Y.-X. Liu, Metrology with PT-Symmetric Cavities: Enhanced Sensitivity near the PT-Phase Transition, Phys. Rev. Lett. 117, 110802 (2016).
  • (26) W. Chen, S. K. Özdemir, G. Zhao, J. Wiersig, and L. Yang, Exceptional points enhance sensing in an optical microcavity, Nature 548, 192 (2017).
  • (27) H. Hodaei, A. U. Hassan, S. Wittek, H. Garcia-Gracia, R. El-Ganainy, D. N. Christodoulides, and M. Khajavikhan, Enhanced sensitivity at higher-order exceptional points, Nature 548, 187 (2017).
  • (28) P. Djorwe, Y. Pennec, and B. Djafari-Rouhani, Exceptional Point Enhances Sensitivity of Optomechanical Mass Sensors, Phys. Rev. Applied 12, 024002(2019).
  • (29) Xuan Mao, Guo-Qing Qin, Hong Yang, Hao Zhang, Min Wang, and Gui-Lu Long, Enhanced sensitivity of optical gyroscope in a mechanical parity-time-symmetric system based on exceptional point, New J. Phys. 22, 093009 (2020).
  • (30) D. Leibfried, E. Knill, S. Seidelin, J. Britton, R. B. Blakestad, J. Chiaverini, D. B. Hume, W. M. Itano, J. D. Jost, C. Langer, R. Ozeri, R. Reichle and D. J. Wineland, Creation of a six-atom ‘Schrödinger cat’ state, Nature (London) 438, 639(2005).
  • (31) Carlton M. Caves, Quantum-mechanical noise in an interferometer, Phys. Rev. D 23, 1693(1981).
  • (32) A. W. Chin, S. F. Huelga, and M. B. Plenio, Quantum Metrology in Non-Markovian Environments, Phys. Rev. Lett. 109, 233601 (2012).
  • (33) R. Demkowicz-Dobrzański, M. Jarzyna, and J. Kolodyński, Quantum limits in optical interferometry, Progress in Optics 60, 345 (2015).
  • (34) V. Giovannetti, S. Lloyd, L. Maccone, Advances in quantum metrology, Nature Photonics 5, 222(2011).
  • (35) A. Luis, Breaking the weak Heisenberg limit, Phys. Rev. A 95, 032113(2017).
  • (36) Kai Bai, Zhen Peng, Hong-Gang Luo, Jun-Hong An, Retrieving ideal precision in noisy quantum optical metrology, Phys. Rev. Lett. 123, 040402 (2019)
  • (37) C. W. Helstrom, Cramér-Rao inequalities for operator-valued measures in quantum mechanics, Int. J. Theor. Phys. 8,361(1973).
  • (38) See Supplemental Material for detailed discussions on the deviation of the error propagation, the derivation of the Fisher information and Cramé-Rao bounds, comparison between the non-Hermitian Fisher information and the Hermitian Fisher information with non-Hermitian signal Hamiltonian, the saturation of Cramé-Rao bounds as well as the measurement of non-Hermitian operators.
  • (39) Jose D. H. Rivero, Li Ge, Pseudo-chirality: a manifestation of Noether’s theorem in non-Hermitian systems, Phys. Rev. Lett. 125, 083902 (2020).
  • (40) Kok-Wei Bong, Nora Tischler, Raj B. Patel, Sabine Wollmann, Geoff J. Pryde, and Michael J. W. Hall, Strong Unitary and Overlap Uncertainty Relations: Theory and Experiment, Phys. Rev. Lett. 120, 230402 (2018).
  • (41) Bing Yu, Naihuan Jing, and Xianqing Li-Jost, Strong unitary uncertainty relations, Phys. Rev. A 100, 022116 (2019).
  • (42) Arun Kumar Pati, Uttam Singh, and Urbasi Sinha, Measuring non-Hermitian operators via weak values, Phys. Rev. A 92, 052120 (2015).
  • (43) Gaurav Nirala, Surya Narayan Sahoo, Arun K. Pati, and Urbasi Sinha, Measuring average of non-Hermitian operator with weak value in a Mach-Zehnder interferometer, Phys. Rev. A 99, 022111 (2019).
  • (44) Surya Narayan Sahoo, Sanchari Chakraborti, Arun K. Pati, and Urbasi Sinha, Quantum State Interferography, Phys. Rev. Lett. 125, 123601(2020).
  • (45) Thomas Monz, Philipp Schindler, Julio T. Barreiro, Michael Chwalla, Daniel Nigg, William A. Coish, Maximilian Harlander, Wolfgang Hänsel, Markus Hennrich, and Rainer Blatt, 14-Qubit Entanglement: Creation and Coherence, Phys. Rev. Lett. 106, 130506 (2011).
  • (46) Marcin Jarzyna and Rafal Demkowicz-Dobrzański, Quantum interferometry with and without an external phase reference, Phys. Rev. A 85, 011801(R)(2012).
  • (47) Luigi Seveso, Matteo A. C. Rossi, and Matteo G. A. Paris, Quantum metrology beyond the quantum Cramér-Rao theorem Phys. Rev. A 95, 012111(2017).

Supplementary material for:
Cramér-Rao bound and quantum parameter estimation with non-Hermitian systems

This supplemental material provides detailed derivations and calculations for the results in the main text. We give numbers to the equations and figures here with ”S” in contrast with that in the main text, for example, Eq.(S1), Fig. S1. Numbers without ”S” refer to the items in the main text, e.g., Eq. (1), Fig. 1. This materials are organized as follows.

  • S1.

    We derive two Cramé-Rao bounds from uncertainty relations of non-Hermitian operators.

  • S2.

    Non-Hermitian Fisher information is defined and previously unknown expressions to calculate the Fisher information is given.

  • S3.

    Two examples to illustrate the non-Hermitian Fisher information are presented.

  • S4.

    We present discussion and comparison between the non-Hermitian Fisher information and the Hermitian Fisher information with non-Hermitian signal Hamiltonian.

  • S5.

    We discuss the second example with the other beam splitter.

  • S6.

    This section devotes to discussion of saturation of Cramé-Rao bounds defined by the non-Hermitian Fisher information.

  • S7.

    We focus on the measurement of non-Hermitian operators. Detailed experimental proposal is given.

S1 S1.  Uncertainly relation and Cramé-Rao bound for non-Hermitian system

The precision of parameter estimation is limited by the quantum Cramér-Rao bound (CRB), which was an extension of Cramér-Rao bound in classical statistics to quantum metrology. As an extension of classical method, CRB has a wide range of applications, meanwhile it lacks a direct physical picture. In this section, we would derive the quantum CRB from the uncertainty relations that root deeply in the quantum theory from the beginning of the last century.

Consider two linear operators AA, BB and a real parameter χ\chi, for any state |Ψ|\Psi\rangle, the following relation holds,

Ψ|(χAiB)(χA+iB)|Ψ0.\langle\Psi|(\chi A^{\dagger}-iB^{\dagger})(\chi A+iB)|\Psi\rangle\geq 0. (S1)

Simple algebra yields,

χ2Ψ|AA|Ψ+χΨ|i(ABBA)|Ψ+Ψ|BB|Ψ0.\chi^{2}\langle\Psi|A^{\dagger}A|\Psi\rangle+\chi\langle\Psi|i(A^{\dagger}B-B^{\dagger}A)|\Psi\rangle+\langle\Psi|B^{\dagger}B|\Psi\rangle\geq 0. (S2)

Noticing Ψ|AA|Ψ0\langle\Psi|A^{\dagger}A|\Psi\rangle\geq 0 regardless of AA being Hermitian or not, and Ψ|i(ABBA)|Ψ\langle\Psi|i(A^{\dagger}B-B^{\dagger}A)|\Psi\rangle is real as Ci(ABBA)C\equiv i(A^{\dagger}B-B^{\dagger}A) is Hermitian, we obtain

Ψ|AA|ΨΨ|BB|Ψ14|Ψ|C|Ψ|2.\langle\Psi|A^{\dagger}A|\Psi\rangle\langle\Psi|B^{\dagger}B|\Psi\rangle\geq\frac{1}{4}|\langle\Psi|C|\Psi\rangle|^{2}. (S3)

Define ΔA=AΨ|A|Ψ\Delta A=A-\langle\Psi|A|\Psi\rangle and the same for ΔB\Delta B, and note the commutation relation, [A,B]=[ΔA,ΔB],[A,B]=[\Delta A,\Delta B], we have

Ψ|ΔAΔA|ΨΨ|ΔBΔB|Ψ14|Ψ|C|Ψ|2.\langle\Psi|\Delta A^{\dagger}\Delta A|\Psi\rangle\langle\Psi|\Delta B^{\dagger}\Delta B|\Psi\rangle\geq\frac{1}{4}|\langle\Psi|C|\Psi\rangle|^{2}. (S4)

This is the equation (not numbered) in the main text.

Although the proof is conducted for pure states, the uncertainty relation Eq.(S4) holds for mixed states. We prove this by introducing an ancilla aa, such that a mixed state ρ=jqj|ψjψj|\rho=\sum_{j}q_{j}|\psi_{j}\rangle\langle\psi_{j}| can be purified to be

|Ψ=jqj|ψj|ϕja,|\Psi^{\prime}\rangle=\sum_{j}\sqrt{q_{j}}|\psi_{j}\rangle\otimes|\phi_{j}\rangle_{a}, (S5)

and the state of the system is obtained by tracing |Ψ|\Psi^{\prime}\rangle over the ancilla, ρ=Tra|ΨΨ|.\rho=\text{Tr}_{a}|\Psi^{\prime}\rangle\langle\Psi^{\prime}|. With this consideration, Eq. (S4) can be straightforwardly extended to the composite system consisting of the system and the ancilla,

Ψ|ΔAIaΔAIa|ΨΨ|ΔBIaΔBIa|Ψ14|Ψ|CIa|Ψ|2.\langle\Psi^{\prime}|\Delta A^{\dagger}\otimes I_{a}\Delta A\otimes I_{a}|\Psi^{\prime}\rangle\langle\Psi^{\prime}|\Delta B^{\dagger}\otimes I_{a}\Delta B\otimes I_{a}|\Psi^{\prime}\rangle\geq\frac{1}{4}|\langle\Psi^{\prime}|C\otimes I_{a}|\Psi^{\prime}\rangle|^{2}. (S6)

Here, IaI_{a} is the identity operator of ancilla aa. Noticing Ψ|ΔAIaΔAIa|Ψ=Tr(ρΔAΔA)\langle\Psi^{\prime}|\Delta A^{\dagger}\otimes I_{a}\Delta A\otimes I_{a}|\Psi^{\prime}\rangle=\text{Tr}(\rho\Delta A^{\dagger}\Delta A) and denoting Tr(ρΔAΔA)=ΔAΔA\text{Tr}(\rho\Delta A^{\dagger}\Delta A)=\langle\Delta A^{\dagger}\Delta A\rangle with Tr representing the trace over the system, we finish the proof of the weaker uncertainly relation in the main text. The slightly stronger uncertainty relation in Eq.(1) with mixed states can be proved in a similar way.

In order to develop a Cramér-Rao bound for non-Hermitian systems, we first have to extend error propagation from Hermitian to non-Hermitian systems, see Eq.(2). Assume a quantum state ρ=ρ(θ)\rho=\rho(\theta) depends on parameter θ\theta, the expectation value of operator AA^{\dagger} and its conjugate AA would then depend on the parameter. Let us discretize the parameter θ\theta by θ1,θ2,θ3,.\theta_{1},\theta_{2},\theta_{3},.... The fluctuation is then,

ΔAΔA=jpj[A(θj)A(θ¯)][A(θj)A(θ¯)],\displaystyle\langle\Delta A^{\dagger}\Delta A\rangle=\sum_{j}p_{j}\left[\langle A^{\dagger}\rangle(\theta_{j})-\langle A^{\dagger}\rangle(\bar{\theta})\right]\left[\langle A\rangle(\theta_{j})-\langle A\rangle(\bar{\theta})\right], (S7)

where θ¯=jpjθj\bar{\theta}=\sum_{j}p_{j}\theta_{j}, and pjp_{j} (j=1,2,3,j=1,2,3,...) stand for probabilities. Expanding A(θj)\langle A\rangle(\theta_{j}) around θ¯\bar{\theta} as

A(θj)=A(θ¯)+Aθ|θ¯(θiθ¯)+\langle A\rangle(\theta_{j})=\langle A\rangle(\bar{\theta})+\frac{\partial\langle A\rangle}{\partial\theta}|_{\bar{\theta}}(\theta_{i}-\bar{\theta})+...

and approximating ΔAΔA\langle\Delta A^{\dagger}\Delta A\rangle up to the second order in (θjθ¯\theta_{j}-\bar{\theta}), we arrive at

ΔAΔA=Aθ|θ¯Aθ|θ¯(Δθ)2.\langle\Delta A^{\dagger}\Delta A\rangle=\frac{\partial\langle A^{\dagger}\rangle}{\partial\theta}|_{\bar{\theta}}\frac{\partial\langle A\rangle}{\partial\theta}|_{\bar{\theta}}(\Delta\theta)^{2}. (S8)

Here, (Δθ)2=jpj(θjθ¯)2(\Delta\theta)^{2}=\sum_{j}p_{j}(\theta_{j}-\bar{\theta})^{2}. This is the error propagation in Eq.(2) of the main text.

With this error propagation, we now derive the non-Hermitian Cramér-Rao bound in Eq.(3). Introducing an operator LL, which is not required to be Hermitian, we have

(Δθ)2AθAθΔLΔL\displaystyle(\Delta\theta)^{2}\frac{\partial\langle A^{\dagger}\rangle}{\partial\theta}\frac{\partial\langle A\rangle}{\partial\theta}\langle\Delta L^{\dagger}\Delta L\rangle =\displaystyle= ΔAΔAΔLΔL\displaystyle\langle\Delta A^{\dagger}\Delta A\rangle\langle\Delta L^{\dagger}\Delta L\rangle (S9)
\displaystyle\geq |ALAL|2.\displaystyle|\langle A^{\dagger}L\rangle-\langle A^{\dagger}\rangle\langle L\rangle|^{2}.

There are few differences between the present study and the earlier publications in LL, which deserve to be outlined. First, in most publications so far, LL is the so-called symmetric logarithmic derivative, which is required to be Hermitian. This requirement is due to the fact that the uncertainty relation used in the publications till now is for Hermitian operators. Whereas in our case LL is not required to be Hermitian. In fact, as we show in the main text an anti-Hermitian L=LL=-L^{\dagger} can maximize the Fisher information. Second, the uncertainty relation used to derive the Cramér bound is different, and our derivation can recover the results in the literature.

Until now, the operator AA is not required to be Hermitian. As we will show later, a non-Hermitian AA will lead a different Cramér-Rao bound.

Let us start to discuss a special case that AA is Hermitian. For Hermitian AA, the last line of Eq. (S9) becomes,

|ALAL|2\displaystyle|\langle A^{\dagger}L\rangle-\langle A^{\dagger}\rangle\langle L\rangle|^{2} =\displaystyle= |12AL+LAAL+12ALLA|2\displaystyle|\frac{1}{2}\langle AL+L^{\dagger}A\rangle-\langle A\rangle\langle L\rangle+\frac{1}{2}\langle AL-L^{\dagger}A\rangle|^{2} (S10)
\displaystyle\geq (AL+LA2)2.\displaystyle\left(\frac{\langle AL+L^{\dagger}A\rangle}{2}\right)^{2}.

To have the last inequality, we have ignored 12ALLA\frac{1}{2}\langle AL-L^{\dagger}A\rangle, which is pure imaginary, while term AL+LA\langle AL+L^{\dagger}A\rangle is real. By carefully choosing LL, we can have L=0.\langle L\rangle=0. Eq. (S9) reduces to Eq.(3) in the main text, where LL satisfies Eq.(4).

When AA is not Hermitian but with L=0\langle L\rangle=0, we still obtain Eq. (3), but in this case, LL satisfies Eq. (5). The proof is almost the same as that for Eq. (4).

Discussions are in order. Either

ρθ=12(L(1)ρ+ρL(1)),\frac{\partial\rho}{\partial\theta}=\frac{1}{2}(L^{(1)}\rho+\rho L^{(1)\dagger}), (S11)

or

ρθ=L(2)ρ=ρL(2).\frac{\partial\rho}{\partial\theta}=L^{(2)}\rho=\rho L^{(2)\dagger}. (S12)

guarantee L=0.\langle L\rangle=0. For Eq.(S11), as we will discuss below, L=LeiβL^{\dagger}=Le^{i\beta} is assumed. This assumption together with Tr(ρθ)=0\text{Tr}\left(\frac{\partial\rho}{\partial\theta}\right)=0 leads to L=0.\langle L\rangle=0. While for Eq.(S12), it is obvious that L=0\langle L\rangle=0 and L=0,\langle L^{\dagger}\rangle=0, since Tr(ρθ)=0\text{Tr}\left(\frac{\partial\rho}{\partial\theta}\right)=0.

S2 S2.  non-Hermitian quantum Fisher information

We present a detailed calculations for the Fisher information FnH(1)F_{nH}^{(1)} and FnH(2)F_{nH}^{(2)} defined in Eq. (3). We first calculate FnH(1).F_{nH}^{(1)}. Without lose of generality, we assume the dimension of the Hilbert space is NN, while the state of the system ρ\rho may be not of full rank. It has positive eigenvalues pjp_{j} required by quantum theory, and the corresponding eigenstates are denoted by |ϕj|\phi_{j}\rangle, i.e.,

ρ=j=1Mpj|ϕjϕj|.\rho=\sum_{j=1}^{M}p_{j}|\phi_{j}\rangle\langle\phi_{j}|. (S13)

Here we assume jj runs from j=1j=1 to j=Mj=M, MNM\leq N. With these notations, the Fishier information defined in Eq.(3) takes,

FnH(1)\displaystyle F_{nH}^{(1)} =\displaystyle= L(1)L(1)eiβ=i=1Mpiϕi|L(1)L(1)|ϕieiβ=j=1Ni=1Mpiϕi|L(1)|ϕjϕj|L(1)|ϕieiβ\displaystyle\langle L^{(1)}L^{(1)}\rangle e^{i\beta}=\sum_{i=1}^{M}p_{i}\langle\phi_{i}|L^{(1)}L^{(1)}|\phi_{i}\rangle e^{i\beta}=\sum_{j=1}^{N}\sum_{i=1}^{M}p_{i}\langle\phi_{i}|L^{(1)}|\phi_{j}\rangle\langle\phi_{j}|L^{(1)}|\phi_{i}\rangle e^{i\beta} (S14)
=\displaystyle= j=1Ni=1Mpi(L(1))ij(L(1))jieiβ\displaystyle\sum_{j=1}^{N}\sum_{i=1}^{M}p_{i}(L^{(1)})_{ij}(L^{(1)})_{ji}e^{i\beta}

Recalling Eq. (4) in the main text that,

ρθ=12(L(1)ρ+ρL(1)),\frac{\partial\rho}{\partial\theta}=\frac{1}{2}(L^{(1)}\rho+\rho L^{(1)\dagger}), (S15)

and assuming L(1)=L(1)eiβL^{(1)\dagger}=L^{(1)}e^{i\beta}, we have

Lij(1)=ϕi|L(1)|ϕj=2(θρ)ijpj+pieiβ.L_{ij}^{(1)}=\langle\phi_{i}|L^{(1)}|\phi_{j}\rangle=2\frac{(\partial_{\theta}\rho)_{ij}}{p_{j}+p_{i}e^{i\beta}}. (S16)

As addressed in the main text, L(1)L^{(1)} should not be limited to the space spanned by the eigenstates of ρ\rho, however for j>Mj>M or i>Mi>M, Lij(1)L_{ij}^{(1)} can not established by Eq. (S16). We will return to this point later. To compute Lij(1)L_{ij}^{(1)}, we need to know (θρ)ij(\partial_{\theta}\rho)_{ij}. Starting from Eq.(S13), we have

θρ=iθpi|ϕiϕi|+ipi|θϕiϕi|+ipi|ϕiθϕi|,\displaystyle\partial_{\theta}\rho=\sum_{i}\partial_{\theta}p_{i}|\phi_{i}\rangle\langle\phi_{i}|+\sum_{i}p_{i}|\partial_{\theta}\phi_{i}\rangle\langle\phi_{i}|+\sum_{i}p_{i}|\phi_{i}\rangle\langle\partial_{\theta}\phi_{i}|, (S17)

leading to

(θρ)ij\displaystyle(\partial_{\theta}\rho)_{ij} =\displaystyle= θpiδij+piθϕi|ϕj+pjϕi|θϕj\displaystyle\partial_{\theta}p_{i}\delta_{ij}+p_{i}\langle\partial_{\theta}\phi_{i}|\phi_{j}\rangle+p_{j}\langle\phi_{i}|\partial_{\theta}\phi_{j}\rangle (S18)
=\displaystyle= θpiδij+(pipj)θϕi|ϕj.\displaystyle\partial_{\theta}p_{i}\delta_{ij}+(p_{i}-p_{j})\langle\partial_{\theta}\phi_{i}|\phi_{j}\rangle.

Here, θϕi|ϕj+ϕi|θϕj=0\langle\partial_{\theta}\phi_{i}|\phi_{j}\rangle+\langle\phi_{i}|\partial_{\theta}\phi_{j}\rangle=0 has been applied in the derivation. Substituting Eq.(S18) and Eq. (S16) into Eq. (S14), we have

FnH(1)\displaystyle F_{nH}^{(1)} =\displaystyle= i=1Mj=1N4pi(θpi)2δijpi2+pj2+2pipjcosβ+i=1Mj=1N4pi(pipj)2θϕi|ϕjϕj|θϕipi2+pj2+2pipjcosβ\displaystyle\sum_{i=1}^{M}\sum_{j=1}^{N}4p_{i}\frac{(\partial_{\theta}p_{i})^{2}\delta_{ij}}{p_{i}^{2}+p_{j}^{2}+2p_{i}p_{j}\cos\beta}+\sum_{i=1}^{M}\sum_{j=1}^{N}4p_{i}\frac{(p_{i}-p_{j})^{2}\langle\partial_{\theta}\phi_{i}|\phi_{j}\rangle\langle\phi_{j}|\partial_{\theta}\phi_{i}\rangle}{p_{i}^{2}+p_{j}^{2}+2p_{i}p_{j}\cos\beta} (S19)
=\displaystyle= i=1M2(θpi)2pi(1+cosβ)+i=1Mj=1M4pi(pipj)2θϕi|ϕjϕj|θϕipi2+pj2+2pipjcosβ+i=1Mj=M+1N4piθϕi|ϕjϕj|θϕi\displaystyle\sum_{i=1}^{M}\frac{2(\partial_{\theta}p_{i})^{2}}{p_{i}(1+\cos\beta)}+\sum_{i=1}^{M}\sum_{j=1}^{M}4p_{i}\frac{(p_{i}-p_{j})^{2}\langle\partial_{\theta}\phi_{i}|\phi_{j}\rangle\langle\phi_{j}|\partial_{\theta}\phi_{i}\rangle}{p_{i}^{2}+p_{j}^{2}+2p_{i}p_{j}\cos\beta}+\sum_{i=1}^{M}\sum_{j=M+1}^{N}4p_{i}\langle\partial_{\theta}\phi_{i}|\phi_{j}\rangle\langle\phi_{j}|\partial_{\theta}\phi_{i}\rangle
=\displaystyle= i=1M2(θpi)2pi(1+cosβ)+i=1Mj=1M4pi(pipj)2θϕi|ϕjϕj|θϕipi2+pj2+2pipjcosβ\displaystyle\sum_{i=1}^{M}\frac{2(\partial_{\theta}p_{i})^{2}}{p_{i}(1+\cos\beta)}+\sum_{i=1}^{M}\sum_{j=1}^{M}4p_{i}\frac{(p_{i}-p_{j})^{2}\langle\partial_{\theta}\phi_{i}|\phi_{j}\rangle\langle\phi_{j}|\partial_{\theta}\phi_{i}\rangle}{p_{i}^{2}+p_{j}^{2}+2p_{i}p_{j}\cos\beta}
+\displaystyle+ i=1Mj=1M4piθϕi|(I|ϕjϕj|)|θϕi,\displaystyle\sum_{i=1}^{M}\sum_{j=1}^{M}4p_{i}\langle\partial_{\theta}\phi_{i}|\left(\frac{}{}I-|\phi_{j}\rangle\langle\phi_{j}|\frac{}{}\right)|\partial_{\theta}\phi_{i}\rangle,

where II is the identity operator of the Hilbert space, i.e., I=i=1N|ϕiϕi|I=\sum_{i=1}^{N}|\phi_{i}\rangle\langle\phi_{i}|. Note that the sum runs from 1 to NN. Finally, we obtain Eq. (8) in the main text,

FnH(1)=i=1M2(θpi)2pi(1+cosβ)+i=1M4piθϕi|θϕi+i=1Mj=1M(4pi(pipj)2pi2+pj2+2pipjcosβ4pi)θϕi|ϕjϕj|θϕi.\displaystyle F_{nH}^{(1)}=\sum_{i=1}^{M}\frac{2(\partial_{\theta}p_{i})^{2}}{p_{i}(1+\cos\beta)}+\sum_{i=1}^{M}4p_{i}\langle\partial_{\theta}\phi_{i}|\partial_{\theta}\phi_{i}\rangle+\sum_{i=1}^{M}\sum_{j=1}^{M}\left(\frac{4p_{i}(p_{i}-p_{j})^{2}}{p_{i}^{2}+p_{j}^{2}+2p_{i}p_{j}\cos\beta}-4p_{i}\right)\langle\partial_{\theta}\phi_{i}|\phi_{j}\rangle\langle\phi_{j}|\partial_{\theta}\phi_{i}\rangle.

FnH(2)F_{nH}^{(2)} can be derived by the same technique as employed in the derivation of FnH(1)F_{nH}^{(1)}. Note that Lij(2)=(θρ)ijpiL^{(2)\dagger}_{ij}=\frac{(\partial_{\theta}\rho)_{ij}}{p_{i}}, and Lji(2)=(θρ)jipiL_{ji}^{(2)}=\frac{(\partial_{\theta}\rho)_{ji}}{p_{i}} for FnH(2)F_{nH}^{(2)}.

S3 S3.  examples

Detailed calculation for the non-Hermitian quantum Fisher information FnH(1)F_{nH}^{(1)} and FnH(2)F_{nH}^{(2)} will be presented below. For mixed states of form ρ=i=12pi|ϕiϕi|\rho=\sum_{i=1}^{2}p_{i}|\phi_{i}\rangle\langle\phi_{i}| with θ\theta-independent pip_{i}, the Fisher information FHF_{H}, FnH(1)F_{nH}^{(1)} and FnH(2)F_{nH}^{(2)} reduce to,

FH\displaystyle F_{H} =\displaystyle= i=12(4piθϕi|θϕi4pi|ϕi|θϕi|2)16p1p2|ϕ1|θϕ2|2,\displaystyle\sum_{i=1}^{2}\left(4p_{i}\langle\partial_{\theta}\phi_{i}|\partial_{\theta}\phi_{i}\rangle-4p_{i}|\langle\phi_{i}|\partial_{\theta}\phi_{i}\rangle|^{2}\right)-16p_{1}p_{2}|\langle\phi_{1}|\partial_{\theta}\phi_{2}\rangle|^{2},
FnH(1)\displaystyle F_{nH}^{(1)} =\displaystyle= i=124piθϕi|θϕi,(β=π),\displaystyle\sum_{i=1}^{2}4p_{i}\langle\partial_{\theta}\phi_{i}|\partial_{\theta}\phi_{i}\rangle,\,(\beta=\pi),
FnH(2)\displaystyle F_{nH}^{(2)} =\displaystyle= i=12(piθϕi|θϕipi|ϕi|θϕi|2)+(p2(13p1)p1+p1(13p2)p2)|ϕ1|θϕ2|2.\displaystyle\sum_{i=1}^{2}\left(p_{i}\langle\partial_{\theta}\phi_{i}|\partial_{\theta}\phi_{i}\rangle-p_{i}|\langle\phi_{i}|\partial_{\theta}\phi_{i}\rangle|^{2}\right)+\left(\frac{p_{2}(1-3p_{1})}{p_{1}}+\frac{p_{1}(1-3p_{2})}{p_{2}}\right)|\langle\phi_{1}|\partial_{\theta}\phi_{2}\rangle|^{2}. (S21)

In the first example, the state before encoding is

ρ\displaystyle\rho =\displaystyle= i=12pi|ϕiϕi|,\displaystyle\sum_{i=1}^{2}p_{i}|\phi_{i}\rangle\langle\phi_{i}|,
|ϕ1,2\displaystyle|\phi_{1,2}\rangle =\displaystyle= 12(|aN±|bN).\displaystyle\frac{1}{\sqrt{2}}(|a\rangle^{\otimes N}\pm|b\rangle^{\otimes N}). (S22)

The signal Hamiltonian used to encode the parameter into the state is JzJ_{z}, i.e., the encoding operation is given by U(θ)=eiθJzU(\theta)=e^{-i\theta J_{z}}. The states that carry the information of the parameter is then

ρ(θ)\displaystyle\rho(\theta) =\displaystyle= i=12pi|ϕi(θ)ϕi(θ)|,\displaystyle\sum_{i=1}^{2}p_{i}|\phi_{i}(\theta)\rangle\langle\phi_{i}(\theta)|,
|ϕ1,2(θ)\displaystyle|\phi_{1,2}(\theta)\rangle =\displaystyle= 12(ei2Nθ|aN±ei2Nθ|bN).\displaystyle\frac{1}{\sqrt{2}}(e^{\frac{i}{2}N\theta}|a\rangle^{\otimes N}\pm e^{-\frac{i}{2}N\theta}|b\rangle^{\otimes N}). (S23)

It is easy to find that

θϕ1|θϕ1\displaystyle\langle\partial_{\theta}\phi_{1}|\partial_{\theta}\phi_{1}\rangle =\displaystyle= 14N2,\displaystyle\frac{1}{4}N^{2},
θϕ2|θϕ2\displaystyle\langle\partial_{\theta}\phi_{2}|\partial_{\theta}\phi_{2}\rangle =\displaystyle= 14N2,\displaystyle\frac{1}{4}N^{2},
θϕ1|ϕ1=θϕ2|ϕ2\displaystyle\langle\partial_{\theta}\phi_{1}|\phi_{1}\rangle=\langle\partial_{\theta}\phi_{2}|\phi_{2}\rangle =\displaystyle= 0,\displaystyle 0,
|θϕ1|ϕ2|2\displaystyle|\langle\partial_{\theta}\phi_{1}|\phi_{2}\rangle|^{2} =\displaystyle= 14N2.\displaystyle\frac{1}{4}N^{2}. (S24)

Finally, we obtain Eq.(12) of the main text.

In the second example, the state before encoding is

ρ\displaystyle\rho =\displaystyle= i=12pi|ϕiϕi|,\displaystyle\sum_{i=1}^{2}p_{i}|\phi_{i}\rangle\langle\phi_{i}|,
|ϕ1\displaystyle|\phi_{1}\rangle =\displaystyle= |rα00S(r)D(α)|00,\displaystyle|r\alpha\rangle_{00}\equiv S(r)\otimes D(\alpha)|00\rangle,
|ϕ2\displaystyle|\phi_{2}\rangle =\displaystyle= |rα01S(r)D(α)b|00,\displaystyle|r\alpha\rangle_{01}\equiv S(r)\otimes D(\alpha)b^{\dagger}|00\rangle, (S25)

where |00|00\rangle denotes that both modes aa and bb are in its vacuum state. S(r)S(r) and D(α)D(\alpha) was defined in the main text, i.e., S(r)=exp(r2a2r2(a)2)S(r)=\exp(\frac{r^{*}}{2}a^{2}-\frac{r}{2}(a^{\dagger})^{2}), D(α)=exp(αaαa)D(\alpha)=\exp(\alpha a^{\dagger}-\alpha^{*}a), and we will set r=|r|ei2ϕr=|r|e^{i2\phi}. The parameter is encoded into the state after passing through a beam splitter represented by BπB_{\pi}. The state encoding the parameter θ\theta is then

ρ(θ)\displaystyle\rho(\theta) =\displaystyle= i=12pi|ϕi(θ)ϕi(θ)|,\displaystyle\sum_{i=1}^{2}p_{i}|\phi_{i}(\theta)\rangle\langle\phi_{i}(\theta)|,
|ϕ1,2(θ)\displaystyle|\phi_{1,2}(\theta)\rangle =\displaystyle= U(θ)BS(r)D(α)|0a|X1,2b,\displaystyle U(\theta)BS(r)D(\alpha)|0\rangle_{a}\otimes|X_{1,2}\rangle_{b}, (S26)

where |X1b=|0b|X_{1}\rangle_{b}=|0\rangle_{b}, |X2b=b|0b,|X_{2}\rangle_{b}=b^{\dagger}|0\rangle_{b}, and p1+p2=1.p_{1}+p_{2}=1. Further, we find

θϕ1|θϕ1\displaystyle\langle\partial_{\theta}\phi_{1}|\partial_{\theta}\phi_{1}\rangle =\displaystyle= |α|4+|α|2,\displaystyle|\alpha|^{4}+|\alpha|^{2},
θϕ2|θϕ2\displaystyle\langle\partial_{\theta}\phi_{2}|\partial_{\theta}\phi_{2}\rangle =\displaystyle= |α|4+5|α|2+1,\displaystyle|\alpha|^{4}+5|\alpha|^{2}+1,
|θϕ1|ϕ1|2\displaystyle|\langle\partial_{\theta}\phi_{1}|\phi_{1}\rangle|^{2} =\displaystyle= |α|4,\displaystyle|\alpha|^{4},
|θϕ2|ϕ2|2\displaystyle|\langle\partial_{\theta}\phi_{2}|\phi_{2}\rangle|^{2} =\displaystyle= (1+|α|2)2,\displaystyle(1+|\alpha|^{2})^{2},
|θϕ1|ϕ2|2\displaystyle|\langle\partial_{\theta}\phi_{1}|\phi_{2}\rangle|^{2} =\displaystyle= |α|2.\displaystyle|\alpha|^{2}. (S27)

Collecting these results, we arrive at

FH\displaystyle F_{H} =\displaystyle= 4|α|2(4p26p+3),\displaystyle 4|\alpha|^{2}(4p^{2}-6p+3),
FnH(1)\displaystyle F_{nH}^{(1)} =\displaystyle= 4|α|4+(2016p)|α|2+4(1p),\displaystyle 4|\alpha|^{4}+(20-16p)|\alpha|^{2}+4(1-p),
FnH(2)\displaystyle F_{nH}^{(2)} =\displaystyle= 2p32p2+1p(1p)|α|2.\displaystyle\frac{2p^{3}-2p^{2}+1}{p(1-p)}|\alpha|^{2}. (S28)

This is Eq.(14) in the main text. The dependence of the Fisher information on the photon number is depicted in Fig.S1.

Refer to caption
Figure S1: Fisher information as a function of photon number. p=0.3p=0.3. This is the numerical simulation for Eq.(14) in the main text.

The Hermitian Fisher information for pure state is also shown for comparison with the others. We can clearly find that for large photon number, FnH(1)F_{nH}^{(1)} scales dominantly with |α|4=N2|\alpha|^{4}=N^{2}, indicating the breakdown of the Heisenberg limit.

S4 S4.  non-Hermitian signal Hamiltonian

In the following, we will consider a situation that the signal Hamiltonian is not Hermitian for FHF_{H}. Comparison is carried out between the Hermitian quantum Fisher information FHF_{H} with Hermitian signal Hamiltonian and the Hermitian quantum Fisher information with non-Hermitian signal Hamiltonian (will be denoted by FHnhsF_{H}^{nhs}), and only the case of pure states is considered.

For pure state |ϕ=|ϕ(θ)|\phi\rangle=|\phi(\theta)\rangle, FnH(1)=FHF_{nH}^{(1)}=F_{H}, so we do not take FnH(1)F_{nH}^{(1)} into comparison. For pure states, FHnhsF_{H}^{nhs} and FHF_{H} takes,

FH\displaystyle F_{H} =\displaystyle= 4θϕ|θϕ4θϕ|ϕϕ|θϕ,\displaystyle 4\langle\partial_{\theta}\phi|\partial_{\theta}\phi\rangle-4\langle\partial_{\theta}\phi|\phi\rangle\langle\phi|\partial_{\theta}\phi\rangle, (S29)
FHnhs\displaystyle F_{H}^{nhs} =\displaystyle= 4(M2θϕ|θϕ1M2(θM)2|θMM+M2ϕ|θϕ|2),\displaystyle 4\left(M^{2}\langle\partial_{\theta}\phi^{\prime}|\partial_{\theta}\phi^{\prime}\rangle-\frac{1}{M^{2}}(\partial_{\theta}M)^{2}-\left|\frac{\partial_{\theta}M}{M}+M^{2}\langle\phi^{\prime}|\partial_{\theta}\phi^{\prime}\rangle\right|^{2}\right), (S30)

where |ϕ|\phi^{\prime}\rangle denotes the unnormalized encoding state and MM is its normalization constant, namely, ϕ|ϕ=1M2\langle\phi^{\prime}|\phi^{\prime}\rangle=\frac{1}{M^{2}}. We might prove Eq.(S30) by substitution of |ϕ=M|ϕ|\phi\rangle=M|\phi^{\prime}\rangle into FH=4θϕ|θϕ4θϕ|ϕϕ|θϕF_{H}=4\langle\partial_{\theta}\phi|\partial_{\theta}\phi\rangle-4\langle\partial_{\theta}\phi|\phi\rangle\langle\phi|\partial_{\theta}\phi\rangle, and noticing

θϕ|θϕ\displaystyle\langle\partial_{\theta}\phi|\partial_{\theta}\phi\rangle =\displaystyle= M2θϕ|θϕ(θM)2M2,\displaystyle M^{2}\langle\partial_{\theta}\phi^{\prime}|\partial_{\theta}\phi^{\prime}\rangle-\frac{(\partial_{\theta}M)^{2}}{M^{2}},
ϕ|θϕ\displaystyle\langle\phi|\partial_{\theta}\phi\rangle =\displaystyle= θMM+M2ϕ|θϕ.\displaystyle\frac{\partial_{\theta}M}{M}+M^{2}\langle\phi^{\prime}|\partial_{\theta}\phi^{\prime}\rangle. (S31)

In the first example, the signal Hamiltonian JzJ_{z} is replaced with Jz(1iγ)J_{z}(1-i\gamma), the encoding state would have the same form as in Eq.(S3) but with N¯=N(1iγ)\bar{N}=N(1-i\gamma) instead of NN. With these knowledge, we have

FH\displaystyle F_{H} =\displaystyle= N2,\displaystyle N^{2},
FHnhs\displaystyle F_{H}^{nhs} =\displaystyle= |N¯|2(1tanh2(γNθ)),\displaystyle{|\bar{N}|^{2}}(1-\tanh^{2}(\gamma N\theta)), (S32)

which is Eq.(15) in main text.

Refer to caption
Figure S2: Fisher information as a function of ion number. This is the numerical simulation for Eq.(15) in the main text. The parameters chosen are θ=0.25π\theta=0.25\pi and γ=0.2\gamma=0.2.

One might wonder why the signal Hamiltonian takes Jz(1iγ)J_{z}(1-i\gamma) instead of JzJ_{z}. This can be understood as follows. Consider the rotation U(θ)=eiθJzU(\theta)=e^{-i\theta J_{z}} applied to encode the parameter θ\theta into the state. This rotation can be realized in trapped ionsmonz11 through an effective Hamiltonian HsignalH_{signal},

Hsignal=Ω2Jz,H_{signal}=\frac{\hbar\Omega}{2}J_{z}, (S33)

where Jz=lsz(l)J_{z}=\sum_{l}s_{z}^{(l)} is a collective spin of those trapped ions, Ω\Omega is the Rabi frequency of each ion or the magnetic field BzB_{z} to which the ion coupled. The time evolution operator U(T)=eiHsingalTeiθJzU(T)=e^{-\frac{i}{\hbar}H_{singal}T}\equiv e^{-i\theta J_{z}} plays the role of the rotation operator. Here the accumulated phase θ=Ω2T.\theta=\frac{\Omega}{2}T.

There is no quantum system being completely isolated from its surroundings including the trapped ions. A system that is coupled with its environment is called open system. A complete description of an open quantum system requires the inclusion of the environment. As a result of the interaction with the environment, the dynamics of the open system is governed by the master equationgardiner04 ,

ρt\displaystyle\frac{\partial\rho}{\partial t} =\displaystyle= i[Hsignal,ρ]+(ρ),\displaystyle-\frac{i}{\hbar}[H_{signal},\rho]+\cal{L}(\rho),
(ρ)\displaystyle\cal{L}(\rho) =\displaystyle= κ2l(2s(l)ρs+(l)ρs+(l)s(l)s+(l)s(l)ρ).\displaystyle\frac{\kappa}{2}\sum_{l}(2s_{-}^{(l)}\rho s_{+}^{(l)}-\rho s_{+}^{(l)}s_{-}^{(l)}-s_{+}^{(l)}s_{-}^{(l)}\rho). (S34)

Consider a very short encoding time, terms with s(l)ρs+(l)s^{(l)}_{-}\rho s^{(l)}_{+} can be ignored. And the dynamics of the open system is governed by an effective Hamiltonian

Heff=Hsignaliκ2ls+(l)s(l).H_{eff}=H_{signal}-\frac{i\hbar\kappa}{2}\sum_{l}s^{(l)}_{+}s^{(l)}_{-}. (S35)

Consider each ion being modelled by a two-level system, the terms ls+(l)s(l)\sum_{l}s^{(l)}_{+}s^{(l)}_{-} would contribute to HsignalH_{signal} as Jz\sim J_{z}. Namely, in this open system

H¯singalJz(1iγ).\bar{H}_{singal}\simeq J_{z}(1-i\gamma). (S36)

Here γκ\gamma\sim\kappa stands for the decay rate of collective spin JzJ_{z}.

For the example with Mach-Zehnder interferometer, we take H¯=aa(1iγ)\bar{H}=a^{\dagger}a(1-i\gamma) to replace HH in the Hermitian encoding case. The calculation is involved but straightforward, it shows that

M2\displaystyle M^{2} =\displaystyle= exp[|α|2(1e2γθ)],\displaystyle\exp\left[\frac{}{}|\alpha|^{2}(1-e^{-2\gamma\theta})\right],
θϕ|θϕ\displaystyle\langle\partial_{\theta}\phi^{\prime}|\partial_{\theta}\phi^{\prime}\rangle =\displaystyle= 1+γ2M2|α|2e2γθ(1+|α|2e2γθ),\displaystyle\frac{1+\gamma^{2}}{M^{2}}|\alpha|^{2}e^{-2\gamma\theta}(1+|\alpha|^{2}e^{-2\gamma\theta}),
θϕ|ϕ\displaystyle\langle\partial_{\theta}\phi^{\prime}|\phi^{\prime}\rangle =\displaystyle= (iγ)|α|2M2e2γθ.\displaystyle(i-\gamma)\frac{|\alpha|^{2}}{M^{2}}e^{-2\gamma\theta}. (S37)

Substituting Eq.(S4) into Eq.(S30), we obtain Eq.(16) in the main text.

S5 S5.  results with the other beam splitter

In Sec.S3, we employed a beam splitter, Bπ=ei2π(ab+ab)B_{\pi}=e^{-\frac{i}{2}\pi(ab^{\dagger}+a^{\dagger}b)}, in a standard Mach-Zehnder setup to study the Fisher information. With beam splitter BπB_{\pi}, the mode bb (or mode aa) is simply transmitted to another mode aa (or mode bb), so there is no interferometer at all, the Fisher information is however not zero jarzyna12 . In the following, we will employ the other beam splitter, Bπ2=ei4π(ab+ab)B_{\frac{\pi}{2}}=e^{-\frac{i}{4}\pi(ab^{\dagger}+a^{\dagger}b)}, to study the Fisher information. Involved but straightforward calculations show that,

θϕ2|θϕ2\displaystyle\langle\partial_{\theta}\phi_{2}|\partial_{\theta}\phi_{2}\rangle =\displaystyle= θϕ1|θϕ1+sinh2|r|+|α|2+12,\displaystyle\langle\partial_{\theta}\phi_{1}|\partial_{\theta}\phi_{1}\rangle+\sinh^{2}|r|+|\alpha|^{2}+\frac{1}{2},
θϕ1|θϕ1\displaystyle\langle\partial_{\theta}\phi_{1}|\partial_{\theta}\phi_{1}\rangle =\displaystyle= 14(|α|4+2|α|2+4|α|2sinh2|r|+sinh4|r|+2sinh2|r|cosh2|r|+sinh2|r|)\displaystyle\frac{1}{4}\left(|\alpha|^{4}+2|\alpha|^{2}+4|\alpha|^{2}\sinh^{2}|r|+\sinh^{4}|r|+2\sinh^{2}|r|\cosh^{2}|r|+\sinh^{2}|r|\right)
+\displaystyle+ 12(α2e2iϕ)cosh|r|sinh|r|,\displaystyle\frac{1}{2}\Re(\alpha^{2}e^{-2i\phi})\cosh|r|\sinh|r|,
|ϕ1|θϕ1|2\displaystyle|\langle\phi_{1}|\partial_{\theta}\phi_{1}\rangle|^{2} =\displaystyle= 14(sinh2|r|+|α|2)2,\displaystyle\frac{1}{4}(\sinh^{2}|r|+|\alpha|^{2})^{2},
|ϕ1|θϕ2|2\displaystyle|\langle\phi_{1}|\partial_{\theta}\phi_{2}\rangle|^{2} =\displaystyle= 14|α|2,\displaystyle\frac{1}{4}|\alpha|^{2},
|ϕ2|θϕ2|2\displaystyle|\langle\phi_{2}|\partial_{\theta}\phi_{2}\rangle|^{2} =\displaystyle= 14(sinh2|r|+|α|2+1)2.\displaystyle\frac{1}{4}(\sinh^{2}|r|+|\alpha|^{2}+1)^{2}. (S38)

Here ()\Re(...) stands for the real part of ()(...). Substituting Eq.(S5) into Eqs(S3), we obtain the Fisher information of states with beam splitter Bπ2B_{\frac{\pi}{2}}. Numerical results are presented in Fig. S3. We find that FnH(2)<FH<FnH(1)F_{nH}^{(2)}<F_{H}<F_{nH}^{(1)} except p0,1p\rightarrow 0,1. This is similar to the results in Fig. 2 of the main text.

Refer to caption
Figure S3: Fisher information as a function of |α|2|\alpha|^{2}. This is the numerical simulation for Fisher information with beam splitter Bπ2B_{\frac{\pi}{2}}. The parameters chosen are ϕ=0.25π\phi=0.25\pi and |r|=2|r|=2. The surf figure just below FHF_{H} (mesh figure) is for the Hermitian quantum information in pure state |ϕ1|\phi_{1}\rangle.

S6 S6.  Saturation of non-Hermitian quantum Cramér-Rao bounds 1/FnH(1)1/F_{nH}^{(1)} and 1/FnH(2)1/F_{nH}^{(2)}

The variance (Δθ)2(\Delta\theta)^{2} of the estimated parameter θ\theta is given by the error propagation Eq.(2) of the main text, it is bounded by the quantum Cramér-Rao bound (QCRB) defined through the quantum Fisher information as

(Δθ)21F,F=FH,FnH(1),FnH(2).(\Delta\theta)^{2}\geq\frac{1}{F},\quad F=F_{H},\,F_{nH}^{(1)},\,F_{nH}^{(2)}.

Given a signal Hamiltonian and initial state, the bounds can be saturated by carefully chosen measurements characterized by variable (operator) AA. In the following, we will show that the bound given by FnH(2)F_{nH}^{(2)} can be saturated, while FnH(1)F_{nH}^{(1)} can not. The optimal measurement AoptA^{opt} to saturate the bound 1FnH(2)\frac{1}{F_{nH}^{(2)}} is also given.

Recalling the Cauchy-Schwarz inequality, F|FG|G|F|G|2\langle F|F\rangle\langle G|G\rangle\geq|\langle F|G\rangle|^{2} and noticing the equality holds if and only if |F=C|G,|F\rangle=C|G\rangle, (CC is a constant) we claim that to saturate the bounds, Aopt=ΓLA^{opt}=\Gamma L (Γ\Gamma is a constant). In other words, to reach the lower Cramér-Rao bound, the optimal measurement variable AoptA^{opt} is required to be proportional to the symmetric logarithmic derivative LL.

In Eq. (S10), a term ALLA\langle A^{\dagger}L-L^{\dagger}A\rangle had been ignored in order to have the last inequality. For the equality to hold, (Aopt)L=LAopt(A^{opt})^{\dagger}L=L^{\dagger}A^{opt} is required, this together with Aopt=ΓLA^{opt}=\Gamma L require that Γ\Gamma is real. This means that to saturate the bounds, AoptA^{opt} and LL must be Hermitian or non-Hermitian simultaneously. This suggests that bound 1/FnH(1)1/F_{nH}^{(1)} can not be saturated, as AA is Hermitian but LL is not as we stated in the main text. In addition, β=π\beta=\pi can not be met as discussed in the main text.

As to the bound given by 1/FnH(2)1/F_{nH}^{(2)}, all requirements for saturation are met, and the optimal measurement AoptA^{opt} can be expressed in terms of the encoding state ρ(θ)\rho(\theta). Firstly, (Aopt)L=LAopt(A^{opt})^{\dagger}L=L^{\dagger}A^{opt} and Aopt=ΓLA^{opt}=\Gamma L can be met simultaneously, and there are no contradictions with the requirement in the main text. Secondly, from Aopt=ΓLA^{opt}=\Gamma L and Eq. (S12), we find that

Aijopt\displaystyle A^{opt}_{ij} \displaystyle\equiv ϕi|Aopt|ϕj=ΓLij,\displaystyle\langle\phi_{i}|A^{opt}|\phi_{j}\rangle=\Gamma L_{ij},
Lij\displaystyle L_{ij} =\displaystyle= (θρ)ijpj,\displaystyle\frac{(\partial_{\theta}\rho)_{ij}}{p_{j}}, (S39)

where, pip_{i} and |ϕi|\phi_{i}\rangle are the eigenvalues and its corresponding eigenstates of the encoding state ρ(θ)\rho(\theta), respectively.

To be specific, we assume pip_{i} to be θ\theta-independent as we did in the main text, then

(θρ)ij=piθϕi|ϕj+pjϕj|θϕi.\displaystyle(\partial_{\theta}\rho)_{ij}=p_{i}\langle\partial_{\theta}\phi_{i}|\phi_{j}\rangle+p_{j}\langle\phi_{j}|\partial_{\theta}\phi_{i}\rangle. (S40)

We apply these results to the first example, and find that the operators given in Eq. (S6) indeed saturate the bound 1/FnH(2)1/F_{nH}^{(2)}, see Fig. S4. The saturation of bound 1/FH1/F_{H} is also presented for comparison, see Fig. S5.

Refer to caption
Figure S4: Bounds 1/FH1/F_{H} (red-dashed), 1/FnH(1)1/F_{nH}^{(1)} (blue-dotted) and 1/FnH(2)1/F_{nH}^{(2)} (blue-solid) as well as the variance of the estimated parameter (Δθ)2(\Delta\theta)^{2} (green-circles) versus pp. pp is defined in Eq. (S3), by which we plot the bounds. The variance (green-circles) is calculated by randomly generating variable AA and computing (Δθ)2(\Delta\theta)^{2} with Eq. (S8).
Refer to caption
Figure S5: The same as Fig. S4. The difference is that the optimal AoptA^{opt} here is for the saturation of 1/FH1/F_{H}.

We would like to notice that for 0.4<p<0.60.4<p<0.6, the bounds 1/FnH(2)1/F_{nH}^{(2)} and 1/FH1/F_{H} are almost same, while 1/FnH(1)1/F_{nH}^{(1)} remains unchanged. See Fig. S6. For p>0.6p>0.6 the bounds and the variance behave similarly with respect the those in the region of p<0.4p<0.4.

Refer to caption
Figure S6: The same as Fig. S4. The bounds 1/FH1/F_{H} (red-solid) and 1/FnH(2)1/F_{nH}^{(2)} (blue-dashed) coincide, while 1/FnH(1)1/F_{nH}^{(1)} is very small with respect to 1/FH1/F_{H} and 1/FnH(2)1/F_{nH}^{(2)}.

From Figures S4, S5 and S6 we find that the bound 1/FnH(1)1/F_{nH}^{(1)} plays no role yet in constraint on the parameter estimation. To show the role that 1/FnH(1)1/F_{nH}^{(1)} plays, we examine the Mach-Zehnder setup with BπB_{\pi} splitter. The results are presented in Fig. S7. The circles in the region enclosed by the red-dashed, blue-solid and the blue-dashed lines (i.e., the region with dotted ellipse) break the constraint by 1/FH1/F_{H} and 1/FnH(2)1/F_{nH}^{(2)} while they are above the lower bound given by 1/FnH(1)1/F_{nH}^{(1)}.

Refer to caption
Figure S7: The bounds 1/FH1/F_{H} (red-dashed), 1/FnH(2)1/F_{nH}^{(2)} (blue-solid) and 1/FnH(1)1/F_{nH}^{(1)} (blue-dotted) as well as the variance (Δθ)2(\Delta\theta)^{2} (green-circles) as a function of pp. This figure shows the relation between the bounds and the variance in the second example, i.e., the Mach-Zehnder setup. We can find that bound 1/FnH(1)1/F_{nH}^{(1)} can not be saturated, however it really provides a lower bound for the variance (Δθ)2(\Delta\theta)^{2}.

S7 S7.  Measuring the average of a non-Hermitian operator

The key point of parameter estimation with non-Hermitian systems is to measure the average of non-Hermitian operator AA. In this section, we follow the proposal in Refs pati15 ; nirala19 ; abbott19 ; sahoo20 to show how to measure the non-Hermitian operator in our case.

Let us consider a non-Hermitian operator AA. The expectation value of AA in a quantum state |ϕin|\phi_{in}\rangle given by ϕin|A|ϕin\langle\phi_{in}|A|\phi_{in}\rangle is complex in general. This makes the non-Hermitian operator AA unobservable in experiments. Nevertheless, recent studies shown that this obstacle can be overcome with the help of polar decomposition hall15 , which states that given any operator AA, we can always decompose AA as A=URA=UR with UU being an unitary operator and RR a Hermitian semidefinite operator, R=AA.R=\sqrt{A^{\dagger}A}. This bridges the average of non-Hermitian operator AA and the weak value of Hermitian operator RR as follows,

ϕin|A|ϕin=ϕin|UR|ϕin=ϕ|R|ϕinϕ|ϕinϕ|ϕin,\displaystyle\langle\phi_{in}|A|\phi_{in}\rangle=\langle\phi_{in}|UR|\phi_{in}\rangle=\frac{\langle\phi|R|\phi_{in}\rangle}{\langle\phi|\phi_{in}\rangle}\langle\phi|\phi_{in}\rangle, (S41)

where ϕ|ϕin|U.\langle\phi|\equiv\langle\phi_{in}|U. It is Well-known that ϕ|R|ϕinϕ|ϕin\frac{\langle\phi|R|\phi_{in}\rangle}{\langle\phi|\phi_{in}\rangle} is a weak value of the positive-semidefinite operator RR, which can be measured directly in the weak measurement aharonov88 .

In fact, the weak measurement is not necessary for the weak value and the average of non-Hermitian operator AA—they can be given in a single-shot measurement with an interferometric technique reported in sahoo20 . Before going into details of such a technique, we first show how to decompose the non-Hermitian operator AA in our first model.

To simplify the representation, let us consider a single trapped ion. For such a system, any non-Hermitian operator AA can be written as,

A=aσ++bσ+cσz+d,A=a\sigma_{+}+b\sigma_{-}+c\sigma_{z}+d, (S42)

where a,b,c,da,b,c,d are complex parameters in general, and σ+=|ab|=(σ),\sigma_{+}=|a\rangle\langle b|=(\sigma_{-})^{\dagger}, σz=|aa||bb|.\sigma_{z}=|a\rangle\langle a|-|b\rangle\langle b|. It is easy to find that

AA=Axσx+Ayσy+Azσz+A0,A^{\dagger}A=A_{x}\sigma_{x}+A_{y}\sigma_{y}+A_{z}\sigma_{z}+A_{0}, (S43)

where Aj,j=x,y,z,0A_{j},\,j=x,y,z,0 can be written in terms of a,b,c,da,b,c,d and are all real since AAA^{\dagger}A is Hermitian. Namely, Ax=(bdbc+ac+ad),A_{x}=\Re(b^{*}d-b^{*}c+ac^{*}+ad^{*}), Ay=(bdbc+ac+ad),A_{y}=-\Im(b^{*}d-b^{*}c+ac^{*}+ad^{*}), Az=2(cd)+12(|b|2|a|2),A_{z}=2{\cal R}(c^{*}d)+\frac{1}{2}(|b|^{2}-|a|^{2}), A0=12(|a|2+|b|2)+|c|2+|d|2.A_{0}=\frac{1}{2}(|a|^{2}+|b|^{2})+|c|^{2}+|d|^{2}. ()\Re(...) and ()\Im(...) denote the real and imaginary part of ().(...). Simple algebra yields the eigenstates and the corresponding eigenvalues of AAA^{\dagger}A,

|E+A\displaystyle|E^{A}_{+}\rangle =\displaystyle= cosΘ2eiΞ|a+sinΘ2|b,\displaystyle\cos\frac{\Theta}{2}e^{-i\Xi}|a\rangle+\sin\frac{\Theta}{2}|b\rangle,
|EA\displaystyle|E^{A}_{-}\rangle =\displaystyle= sinΘ2eiΞ|acosΘ2|b,\displaystyle\sin\frac{\Theta}{2}e^{-i\Xi}|a\rangle-\cos\frac{\Theta}{2}|b\rangle,
E±A\displaystyle E^{A}_{\pm} =\displaystyle= A0±Ax2+Ay2+Az2.\displaystyle A_{0}\pm\sqrt{A_{x}^{2}+A_{y}^{2}+A_{z}^{2}}. (S44)

Here, tanΞ=AyAx,\tan\Xi=\frac{A_{y}}{A_{x}}, and cosΘ=AzAx2+Ay2+Az2.\cos\Theta=\frac{A_{z}}{\sqrt{A_{x}^{2}+A_{y}^{2}+A_{z}^{2}}}. Then we have, AA=j=+,EjA|EjAEjA|,A^{\dagger}A=\sum_{j=+,-}E^{A}_{j}|E^{A}_{j}\rangle\langle E^{A}_{j}|, leading to the decomposition,

R=AA=j=+,EjA|EjAEjA|.R=\sqrt{A^{\dagger}A}=\sum_{j=+,-}\sqrt{E^{A}_{j}}|E^{A}_{j}\rangle\langle E^{A}_{j}|. (S45)

Note that E±A0E^{A}_{\pm}\geq 0 as AAA^{\dagger}A is Hermitian and positive. This observation together with A=URA=UR leads to

U=Aj=+,1EjA|EjAEjA|.U=A\sum_{j=+,-}\frac{1}{\sqrt{E^{A}_{j}}}|E^{A}_{j}\rangle\langle E^{A}_{j}|. (S46)

With the decomposition A=URA=UR, Eq. (S45) and Eq. (S46), we now go to the measurement of non-Hermitian operator AA. The schematic setup is illustrated in Fig. S8. The average of AA in state |ϕin|\phi_{in}\rangle, ϕin|A|ϕin|A|eiζ\langle\phi_{in}|A|\phi_{in}\rangle\equiv|A|e^{i\zeta} can be readout from the average intensity measured by the detector.

Refer to caption
Figure S8: Schematic setup for measurement of non-Hermitian operator A=UR.A=UR. B1B_{1} and B2B_{2} are 50:50 Hadamard-type beam splitters, which split the spatial modes representing by |1|1\rangle and |2|2\rangle. eiχe^{i\chi} is a phase shifter that introduces a relative phase χ\chi between the two arms, and we measure the intensity at the detector as a function of χ\chi.

One point deserved to be addressed is that this set up can work for both polarized photons and two-level atoms. So, we would not specify which interferometer it is (Mach-Zehnder or Ramsey). In bases spanned by |a,1,|b,1,|a,2,|b,2|a,1\rangle,|b,1\rangle,|a,2\rangle,|b,2\rangle (|x,n|x|n,x=a,b;n=1,2|x,n\rangle\equiv|x\rangle\otimes|n\rangle,x=a,b;n=1,2), these operation Bi(i=1,2),R,U,eiχB_{i}(i=1,2),R,U^{\dagger},e^{i\chi} can be written as

B1,2\displaystyle B_{1,2} =\displaystyle= 12(10i0010ii0100i01),R=(RaaRab00RbaRbb0000100001),\displaystyle\frac{1}{\sqrt{2}}\left(\begin{array}[]{cccc}1&0&i&0\\ 0&1&0&i\\ i&0&1&0\\ 0&i&0&1\\ \end{array}\right),\quad R=\left(\begin{array}[]{cccc}R_{aa}&R_{ab}&0&0\\ R_{ba}&R_{bb}&0&0\\ 0&0&1&0\\ 0&0&0&1\\ \end{array}\right), (S55)
eiχ\displaystyle e^{i\chi} =\displaystyle= (1000010000eiχ0000eiχ),U=(1000010000UaaUab00UbaUbb),\displaystyle\left(\begin{array}[]{cccc}1&0&0&0\\ 0&1&0&0\\ 0&0&e^{i\chi}&0\\ 0&0&0&e^{i\chi}\\ \end{array}\right),\quad U^{\dagger}=\left(\begin{array}[]{cccc}1&0&0&0\\ 0&1&0&0\\ 0&0&U^{\dagger}_{aa}&U^{\dagger}_{ab}\\ 0&0&U^{\dagger}_{ba}&U^{\dagger}_{bb}\\ \end{array}\right), (S64)

where Rab=a|R|bR_{ab}=\langle a|R|b\rangle, and similar denotations hold for Raa,Rbb,RbaR_{aa},R_{bb},R_{ba} and Uxy,x,y=a,b.U^{\dagger}_{xy},x,y=a,b. With the same bases, the input state that only occupies spatial state |1|1\rangle takes,

|ϕin\displaystyle|\phi_{in}\rangle =\displaystyle= (ϕaϕb00),ϕaa|ϕin,ϕbb|ϕin.\displaystyle\left(\begin{array}[]{c}\phi_{a}\\ \phi_{b}\\ 0\\ 0\\ \end{array}\right),\phi_{a}\equiv\langle a|\phi_{in}\rangle,\,\phi_{b}\equiv\langle b|\phi_{in}\rangle. (S69)

The output state reads

|ϕout=B2eiχURB1|ϕin,|\phi_{out}\rangle=B_{2}e^{i\chi}U^{\dagger}RB_{1}|\phi_{in}\rangle, (S70)

simple algebra yields,

|ϕout\displaystyle|\phi_{out}\rangle =12\displaystyle=\frac{1}{2} (Raaϕa+Rabϕb(Uaaϕa+Uabϕb)eiχRbaϕa+Rbbϕb(Ubaϕa+Ubbϕb)eiχi[Raaϕa+Rabϕb+(Uaaϕa+Uabϕb)eiχ]i[Rbaϕa+Rbbϕb+(Ubaϕa+Ubbϕb)eiχ]).\displaystyle\left(\begin{array}[]{c}R_{aa}\phi_{a}+R_{ab}\phi_{b}-(U^{\dagger}_{aa}\phi_{a}+U^{\dagger}_{ab}\phi_{b})e^{i\chi}\\ R_{ba}\phi_{a}+R_{bb}\phi_{b}-(U^{\dagger}_{ba}\phi_{a}+U^{\dagger}_{bb}\phi_{b})e^{i\chi}\\ i[R_{aa}\phi_{a}+R_{ab}\phi_{b}+(U^{\dagger}_{aa}\phi_{a}+U^{\dagger}_{ab}\phi_{b})e^{i\chi}]\\ i[R_{ba}\phi_{a}+R_{bb}\phi_{b}+(U^{\dagger}_{ba}\phi_{a}+U^{\dagger}_{bb}\phi_{b})e^{i\chi}]\\ \end{array}\right). (S75)

The intensity the detector measures can be represented by

I(χ)=|ϕD|ϕD|2,|ϕDD|ϕout,I(\chi)=|\langle\phi_{D}|\phi_{D}\rangle|^{2},\quad|\phi_{D}\rangle\equiv D|\phi_{out}\rangle, (S76)

with

D=(0000000000100001),D=\left(\begin{array}[]{cccc}0&0&0&0\\ 0&0&0&0\\ 0&0&1&0\\ 0&0&0&1\\ \end{array}\right), (S77)

leading to

I(χ)=14(1+ϕin|R2|ϕin+2|ϕin|A|ϕin|cos(χζ)),I(\chi)=\frac{1}{4}\left(\frac{}{}1+\langle\phi_{in}|R^{2}|\phi_{in}\rangle+2|\langle\phi_{in}|A|\phi_{in}\rangle|\cos(\chi-\zeta)\right), (S78)

where ζ\zeta is the argument of ϕin|A|ϕin\langle\phi_{in}|A|\phi_{in}\rangle, i.e., ζ=argϕin|Aϕin.\zeta=\arg\langle\phi_{in}|A|\phi_{in}\rangle. In experiment, the intensity I(χ)I(\chi) together with ϕin|R2|ϕin\langle\phi_{in}|R^{2}|\phi_{in}\rangle which is the average of Hermitian operator R2R^{2} can determine the average of non-Hermitian operator A,A, as both |ϕin|A|ϕin||\langle\phi_{in}|A|\phi_{in}\rangle| and ζ\zeta can be inferred from the intensity I(χ)I(\chi).

References

  • (1) Thomas Monz, Philipp Schindler, Julio T. Barreiro, Michael Chwalla, Daniel Nigg, William A. Coish, Maximilian Harlander, Wolfgang Hänsel, Markus Hennrich, and Rainer Blatt, 14-Qubit Entanglement: Creation and Coherence, Phys. Rev. Lett. 106, 130506 (2011).
  • (2) C. W. Gardiner, and P. Zoller, Quantum Noise, (Berlin Heidelberg, Springer, 2004).
  • (3) Marcin Jarzyna and Rafal Demkowicz-Dobrzański, Quantum interferometry with and without an external phase reference, Phys. Rev. A 85, 011801(R)(2012).
  • (4) Arun Kumar Pati, Uttam Singh, and Urbasi Sinha, Measuring non-Hermitian operators via weak values, Phys. Rev. A 92, 052120 (2015).
  • (5) Gaurav Nirala, Surya Narayan Sahoo, Arun K. Pati, and Urbasi Sinha, Measuring average of non-Hermitian operator with weak value in a Mach-Zehnder interferometer, Phys. Rev. A 99, 022111 (2019).
  • (6) A. A. Abbott, R. Silva, J. Wechs, N. Brunner, and C. Branciard, Anomalous weak values without post-selection, Quantum 3, 194 (2019).
  • (7) Surya Narayan Sahoo, Sanchari Chakraborti, Arun K. Pati, and Urbasi Sinha, Quantum State Interferography, Phys. Rev. Lett. 125, 123601(2020).
  • (8) B. C. Hall, Lie Groups, Lie Algebras, and Representations: An Elementary Introduction (Springer, New York, 2015).
  • (9) Yakir Aharonov, David Z. Albert, and Lev Vaidman, How the result of a measurement of a component of the spin of a spin-1/2 particle can turn out to be 100, Phys. Rev. Lett. 60, 1351(1988).