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

\floatsetup

[figure]style=plain,subcapbesideposition=top

Optimal Strategies of Quantum Metrology with a Strict Hierarchy

Qiushi Liu [email protected] QICI Quantum Information and Computation Initiative, Department of Computer Science, The University of Hong Kong, Pokfulam Road, Hong Kong, China    Zihao Hu [email protected] Department of Mechanical and Automation Engineering, The Chinese University of Hong Kong, Shatin, Hong Kong, China    Haidong Yuan [email protected] Department of Mechanical and Automation Engineering, The Chinese University of Hong Kong, Shatin, Hong Kong, China    Yuxiang Yang [email protected] QICI Quantum Information and Computation Initiative, Department of Computer Science, The University of Hong Kong, Pokfulam Road, Hong Kong, China
Abstract

One of the main quests in quantum metrology is to attain the ultimate precision limit with given resources, where the resources are not only of the number of queries, but more importantly of the allowed strategies. With the same number of queries, the restrictions on the strategies constrain the achievable precision. In this work, we establish a systematic framework to identify the ultimate precision limit of different families of strategies, including the parallel, the sequential, and the indefinite-causal-order strategies, and provide an efficient algorithm that determines an optimal strategy within the family of strategies under consideration. With our framework, we show there exists a strict hierarchy of the precision limits for different families of strategies.

preprint: APS/123-QED

Introduction.—Quantum metrology [1, 2] features a series of promising applications in the near future [3]. In the prototypical setting of quantum metrology, the goal is to estimate an unknown parameter carried by a quantum channel, given NN queries to it. A pivotal task is to design a strategy that utilizes these NN queries to generate a quantum state with as much information about the unknown parameter as possible. This often involves, for example, preparing a suitable input probe state [4, 5, 6] and applying intermediate quantum control [7, 8, 9, 10] as well as quantum error correction [11, 12, 13, 14].

In reality, the implementation of strategies is subject to physical restrictions. In particular, within the noisy and intermediate-scale quantum (NISQ) era [15], we have to adjust the strategy to accommodate the limitations on the system. For example, for systems with short coherence time it might be favorable to adopt the parallel strategy [Fig. 1LABEL:sub@subfig:parallel_strategy], where multiple queries of the unknown channel are applied simultaneously on a multipartite entangled state [4]. When the system has longer coherence time and can be better controlled, one could choose to query the channel sequentially [Fig. 1LABEL:sub@subfig:sequential_strategy], which may potentially enhance the precision. In addition to the parallel and sequential strategies, it was recently discovered that the quantum SWITCH [16], a primitive where the order of making queries to the unknown channel is in a quantum superposition [Fig. 1LABEL:sub@subfig:quantum_switch_strategy], can be employed to generate new strategies of quantum metrology [17, 18, 19] that may even break the Heisenberg limit [19]. Moreover, indefinite causal structures beyond the quantum SWITCH [16, 20, 21] [Figs. 1LABEL:sub@subfig:causal_superposition_strategy and LABEL:sub@subfig:general_strategy] have recently been shown to further boost the performance of certain information processing tasks [22, 23]. The ultimate performance of these strategies in quantum metrology, however, remains unknown. This is mainly due to the lack of a systematic method that optimizes the probe state, the control and other degrees of freedom in a strategy in a unified fashion which leads to the ultimate precision limit.

\sidesubfloat

[]Refer to caption \sidesubfloat[]Refer to caption
\sidesubfloat[]Refer to caption \sidesubfloat[]Refer to caption \sidesubfloat[]Refer to caption

Figure 1: Prototypical strategies of quantum metrology (for the N=2N=2 case). ϕ\mathcal{E}_{\phi} is a quantum channel carrying an unknown parameter ϕ\phi, and the blue shaded area represents a strategy. LABEL:sub@subfig:parallel_strategy A parallel strategy. LABEL:sub@subfig:sequential_strategy A sequential strategy, where UU is a control operation. LABEL:sub@subfig:quantum_switch_strategy A quantum SWITCH strategy. The blue and red lines, respectively, correspond to two different execution orders entangled with a control qubit. LABEL:sub@subfig:causal_superposition_strategy A causal superposition strategy. Two sequential strategies, plotted in blue and in red, respectively, are entangled with a control qubit (not shown in the figure) and the output will be measured with the control qubit collectively. LABEL:sub@subfig:general_strategy A general indefinite-causal-order strategy.

In this work, we develop a semidefinite programming (SDP) method of evaluating the optimal precision of single-parameter quantum metrology for finite NN (which we call the nonasymptotic regime) over a family of admissible strategies. With this method, we show a strict hierarchy (see Fig. 2) of the optimal performances under different families of strategies, which include the parallel, the sequential, and the indefinite-causal-order [16, 20, 21] ones (see Fig. 1). In conjunction, we design an algorithm to obtain an optimal strategy achieving the highest precision. For the strategy set that admits a symmetric structure, we develop a method of reducing the complexity of our algorithms by an exponential factor.

Quantum Fisher information.—The uncertainty δϕ^\delta\hat{\phi} of estimating an unknown parameter ϕ\phi encoded in a quantum state ρϕ\rho_{\phi}, for any unbiased estimator ϕ^\hat{\phi}, can be determined via the quantum Cramér-Rao bound (QCRB) as δϕ^1/νJQ(ρϕ)\delta\hat{\phi}\geq 1/\sqrt{\nu J_{Q}(\rho_{\phi})}[24, 25, 26], where JQ(ρϕ)J_{Q}(\rho_{\phi}) is the quantum Fisher information (QFI) of the state ρϕ\rho_{\phi} and ν\nu is the number of repeated measurements [27]. For single-parameter estimation, the QCRB is achievable, and the QFI thus quantifies the amount of information that can be extracted from the quantum state. One way to compute the QFI is [28, 29]

JQ(ρϕ)=4minTrA(|ΨϕΨϕ|)=ρϕΨ˙ϕ|Ψ˙ϕ,J_{Q}(\rho_{\phi})=4\min_{\Tr_{A}{(\outerproduct{\Psi_{\phi}}{\Psi_{\phi}})}=\rho_{\phi}}\innerproduct{\dot{\Psi}_{\phi}}{\dot{\Psi}_{\phi}}, (1)

where |Ψϕ\ket{\Psi_{\phi}} is the purification of ρϕ\rho_{\phi} with an ancillary space A\mathcal{H}_{A}, TrA\Tr_{A} denotes the partial trace over A\mathcal{H}_{A}, and Ψ˙:=Ψ/ϕ\dot{\Psi}:=\partial\Psi/\partial\phi. When the parameter is carried by a quantum channel ϕ\mathcal{E}_{\phi}, i.e., a completely positive trace-preserving (CPTP) map, the channel QFI can be defined as the maximal QFI of output states using the optimal input assisted by arbitrary ancillae [28, 30, 31, 32]: JQ(chan)(ϕ)=maxρin𝒮(SA)JQ[(ϕA)(ρin)]J_{Q}^{(\mathrm{chan})}(\mathcal{E}_{\phi})=\max_{\rho_{\mathrm{in}}\in\mathcal{S}(\mathcal{H}_{S}\otimes\mathcal{H}_{A})}J_{Q}[(\mathcal{E}_{\phi}\otimes\mathcal{I}_{A})(\rho_{\mathrm{in}})], where 𝒮()\mathcal{S}(\mathcal{H}) denotes the space of density operators on the Hilbert space \mathcal{H}, S/A\mathcal{H}_{S/A} denotes the Hilbert space of the system or ancillae, and A\mathcal{I}_{A} is the identity on A\mathcal{H}_{A}.

We denote by ()\mathcal{L}(\mathcal{H}) the set of linear operators on the finite-dimensional Hilbert space \mathcal{H}, and [(1),(2)]\mathcal{L}[\mathcal{L}(\mathcal{H}_{1}),\mathcal{L}(\mathcal{H}_{2})] denotes the set of linear maps from (1)\mathcal{L}(\mathcal{H}_{1}) to (2)\mathcal{L}(\mathcal{H}_{2}). By the Choi-Jamiołkowski (CJ) isomorphism, a parametrized quantum channel ϕ[(2i1),(2i)]\mathcal{E}_{\phi}\in\mathcal{L}[\mathcal{L}(\mathcal{H}_{2i-1}),\mathcal{L}(\mathcal{H}_{2i})] (for 1iN1\leq i\leq N) can be represented by a positive semidefinite operator (called the CJ operator) Eϕ=𝖢𝗁𝗈𝗂(ϕ)=ϕ(|I\rrangle\llangleI|)E_{\phi}=\mathsf{Choi}(\mathcal{E}_{\phi})=\mathcal{E}_{\phi}\otimes\mathcal{I}(\lvert I\rrangle\llangle I\rvert), where |I\rrangle=i|i|i\lvert I\rrangle=\sum_{i}\ket{i}\ket{i}. The CJ operator of NN identical quantum channels is Nϕ=EϕN(i=12Ni)N_{\phi}=E_{\phi}^{\otimes N}\in\mathcal{L}\left(\otimes_{i=1}^{2N}\mathcal{H}_{i}\right).

Strategy set in quantum metrology.—A strategy is an arrangement of physical processes (the blue shaded area in Fig. 1) which, when concatenated with given queries to ϕ\mathcal{E}_{\phi}, generates an output quantum state carrying the information about ϕ\phi. A strategy can be described by a CJ operator on (Fi=12Ni)\mathcal{L}\left(\mathcal{H}_{F}\otimes_{i=1}^{2N}\mathcal{H}_{i}\right), where F\mathcal{H}_{F} denotes the output Hilbert space of the concatenation, referred to as the global future space. The concatenation of two processes is characterized by the link product [33, 34] of two corresponding CJ operators A(a𝒜a)A\in\mathcal{L}\left(\otimes_{a\in\mathcal{A}}\mathcal{H}_{a}\right) and B(bb)B\in\mathcal{L}\left(\otimes_{b\in\mathcal{B}}\mathcal{H}_{b}\right) as

AB:=Tr𝒜[(𝟙\𝒜AT𝒜)(B𝟙𝒜\)],A*B:=\Tr_{\mathcal{A}\cap\mathcal{B}}\left[\left(\mathds{1}_{\mathcal{B}\backslash\mathcal{A}}\otimes A^{T_{\mathcal{A}\cap\mathcal{B}}}\right)\left(B\otimes\mathds{1}_{\mathcal{A}\backslash\mathcal{B}}\right)\right], (2)

where TiT_{i} denotes the partial transpose on i\mathcal{H}_{i}, and 𝒜/\mathcal{H}_{\mathcal{A}/\mathcal{B}} denotes i𝒜/i\otimes_{i\in\mathcal{A}/\mathcal{B}}\mathcal{H}_{i}. The output state lies in the global future FF, which should not affect any state in the past.

Following the above formalism, given a sufficiently large ancillary Hilbert space, a strategy set determined by the relevant causal constraints is described by a subset 𝖯\mathsf{P} of

𝖲𝗍𝗋𝖺𝗍:={P(Fi=12Ni)|P0,𝗋𝖺𝗇𝗄(P)=1}.\mathsf{Strat}:=\left\{P\in\mathcal{L}\left(\mathcal{H}_{F}\otimes_{i=1}^{2N}\mathcal{H}_{i}\right)\middle|P\geq 0,\mathsf{rank}(P)=1\right\}. (3)

Here, without loss of generality [35], we have restricted PP to pure processes (rank-1 operators) due to the monotonicity of QFI [36]. Our goal is to identify the ultimate precision limit of parameter estimation characterized by the QFI within such constraints:

Definition 1.

The QFI of NN quantum channels ϕ\mathcal{E}_{\phi} [37] given a strategy set 𝖯\mathsf{P} is

J(𝖯)(Nϕ):=maxP𝖯JQ(PNϕ),J^{(\mathsf{P})}(N_{\phi}):=\max_{P\in\mathsf{P}}J_{Q}(P*N_{\phi}), (4)

where JQ(ρ)J_{Q}(\rho) is the QFI of the state ρ\rho, and NϕN_{\phi} is the CJ operator of NN channels.

In general we can write the ensemble decomposition [28] of the CJ operator NϕN_{\phi} as Nϕ=i=1r|Nϕ,iNϕ,i|=𝐍ϕ𝐍ϕN_{\phi}=\sum_{i=1}^{r}\outerproduct{N_{\phi,i}}{N_{\phi,i}}=\mathbf{N}_{\phi}\mathbf{N}_{\phi}^{\dagger}, where 𝐍ϕ:=(|Nϕ,1,,|Nϕ,r)\mathbf{N}_{\phi}:=\left(\ket{N_{\phi,1}},\dots,\ket{N_{\phi,r}}\right) and r:=maxϕ𝗋𝖺𝗇𝗄(Nϕ)r:=\max_{\phi}\mathsf{rank}({N_{\phi}}). We also define 𝐍~˙ϕ:=𝐍˙ϕi𝐍ϕh\dot{\tilde{\mathbf{N}}}_{\phi}:=\dot{\mathbf{N}}_{\phi}-\mathrm{i}\mathbf{N}_{\phi}h for hrh\in\mathbb{H}_{r}, where r\mathbb{H}_{r} denotes the set of r×rr\times r Hermitian matrices, and the performance operator [38]

Ωϕ(h):=4(𝐍~˙ϕ𝐍~˙ϕ)T.\Omega_{\phi}(h):=4\left(\dot{\tilde{\mathbf{N}}}_{\phi}\dot{\tilde{\mathbf{N}}}_{\phi}^{\dagger}\right)^{T}. (5)

With these notions, we can show (see Appendix A, which is analogous to the approach in [38]) that the QFI admits the form

J(𝖯)(Nϕ)=maxP~𝖯~minhrTr[P~Ωϕ(h)],J^{(\mathsf{P})}(N_{\phi})=\max_{\tilde{P}\in\tilde{\mathsf{P}}}\min_{h\in\mathbb{H}_{r}}\Tr\left[\tilde{P}\Omega_{\phi}(h)\right], (6)

with

𝖯~:={P~=TrFP|P𝖯}.\tilde{\mathsf{P}}:=\left\{\tilde{P}=\Tr_{F}P\middle|P\in\mathsf{P}\right\}. (7)

To evaluate the QFI, we first exchange maxP~\max_{\tilde{P}} and minh\min_{h} without changing the optimal QFI, assured by the minimax theorem [39, 40] since the objective function is concave on P~\tilde{P} and convex on hh [41]. Hence, the problem is cast into

J(𝖯)(Nϕ)=minhmaxP~𝖯~Tr[P~Ωϕ(h)].J^{(\mathsf{P})}(N_{\phi})=\min_{h}\max_{\tilde{P}\in\tilde{\mathsf{P}}}\Tr\left[\tilde{P}\Omega_{\phi}(h)\right]. (8)

Then we fix hh and formulate the dual problem of maximization over the set 𝖯~\tilde{\mathsf{P}}. Finally we further optimize the value of hh. To simplify the calculation of QFI we require that

𝖯~=𝖢𝗈𝗇𝗏{i=1K{Si0|Si𝖲i}},\tilde{\mathsf{P}}=\mathsf{Conv}\left\{\bigcup_{i=1}^{K}\left\{S^{i}\geq 0\middle|S^{i}\in\mathsf{S}^{i}\right\}\right\}, (9)

where 𝖢𝗈𝗇𝗏{}\mathsf{Conv}\{\cdot\} denotes the convex hull, and each 𝖲i\mathsf{S}^{i} for i=1,,Ki=1,\dots,K is an affine space of Hermitian operators. Adopting the above-mentioned method, we get

Theorem 1.

Given an arbitrary strategy set 𝖯\mathsf{P} such that 𝖯~\tilde{\mathsf{P}} given by Eq. (7) satisfies the condition Eq. (9), the QFI of NN quantum channels ϕ\mathcal{E}_{\phi} can be expressed as the following optimization problem:

J(𝖯)(Nϕ)=\displaystyle J^{(\mathsf{P})}(N_{\phi})= minλ,Qi,hλ,\displaystyle\min_{\lambda,Q^{i},h}\lambda, (10)
s.t.\displaystyle\mathrm{s.t.} λQiΩϕ(h),Qi𝖲¯i,i=1,,K,\displaystyle\,\lambda Q^{i}\geq\Omega_{\phi}(h),\ Q^{i}\in\overline{\mathsf{S}}^{i},\ i=1,\dots,K,

where 𝖲¯i:={QisHermitian|Tr(QS)=1,S𝖲i}\overline{\mathsf{S}}^{i}:=\left\{Q\mathrm{\ is\ Hermitian}\middle|\Tr(QS)=1,S\in\mathsf{S}^{i}\right\} is the dual affine space of 𝖲i\mathsf{S}^{i}.

The proof can be found in Appendix B. We remark that similar optimization ideas have been applied to other tasks, such as quantum Bayesian estimation [42], quantum network optimization [43], non-Markovian quantum metrology [38], and quantum channel discrimination [23]. The minimization problem in Theorem 1 can be further written in the form of SDP and solved efficiently, with detailed numerically solvable forms given in Appendix E, where the constraints in Eq. (10) can be further simplified in some cases.

Optimal strategies.—By itself, the QFI does not reveal how to implement the optimal strategy achieving the highest precision. Here, in addition to Theorem 1, we design an algorithm that yields a strategy attaining the optimal QFI for any strategy set satisfying Eq. (9). The method, which generalizes the method of finding an optimal probe state for a single channel [44, 45], is summarized as Algorithm 1 (see Appendix C for its derivation).

Algorithm 1 Find an optimal strategy in the set 𝖯\mathsf{P}.
  1. (i)

    Given NϕN_{\phi} the CJ operator of NN channels, solve for an optimal value h=h(opt)h=h^{(\mathrm{opt})} in Eq. (10) of Theorem 1 via SDP.

  2. (ii)

    Fixing h=h(opt)h=h^{(\mathrm{opt})}, solve for an optimal value P~(opt)\tilde{P}^{(\mathrm{opt})} of P~𝖯~\tilde{P}\in\mathsf{\tilde{P}} in Eq. (6) via SDP such that

    Re{Tr{P~(opt)[i𝐍ϕ(𝐍˙ϕi𝐍ϕh(opt))]T}}\displaystyle\real\left\{\Tr\left\{\tilde{P}^{(\mathrm{opt})}\left[-\mathrm{i}\mathbf{N}_{\phi}\mathscr{H}\left(\dot{\mathbf{N}}_{\phi}-\mathrm{i}\mathbf{N}_{\phi}h^{(\mathrm{opt})}\right)^{\dagger}\right]^{T}\right\}\right\} (11)
    =0forallr,\displaystyle=0\ \mathrm{for}\ \mathrm{all}\ \mathscr{H}\in\mathbb{H}_{r},

    where 𝐍ϕ:=(|Nϕ,1,,|Nϕ,r)\mathbf{N}_{\phi}:=\left(\ket{N_{\phi,1}},\dots,\ket{N_{\phi,r}}\right). An optimal strategy P(opt)𝖯P^{(\mathrm{opt})}\in\mathsf{P} can be taken as a purification of P~(opt)\tilde{P}^{(\mathrm{opt})}.

By Algorithm 1 we obtain the CJ operator of a strategy that attains the optimal QFI. For strategies following definite causal order, there exists an operational method of mapping the CJ operator of the strategy to a probe state and a sequence of in-between control operations with minimal memory space [46]. For causal order superposition strategies (see the strategy set 𝖲𝗎𝗉\mathsf{Sup}), we show that they can always be implemented by controlling the order of operations in a circuit with a quantum SWITCH (see Appendix I). In this way, we obtain a systematic method to identify optimal sequential and causal superposition strategies, one of the key problems in quantum metrology.

Strategy sets.—We consider the evaluation of QFI for five different families of strategies. In all the following definitions the subscript ii of an operator denotes the Hilbert space i\mathcal{H}_{i} it acts on.

The family of parallel strategies [see Fig. 1LABEL:sub@subfig:parallel_strategy] is the first and one of the most successful examples of quantum-enhanced metrology, featuring the usage of entanglement to achieve precision beyond the classical limit [47]. By making parallel use of NN quantum channels together with ancillae, we can regard these NN channels as one single channel from (i=1N2i1)\mathcal{L}\left(\otimes_{i=1}^{N}\mathcal{H}_{2i-1}\right) to (i=1N2i)\mathcal{L}\left(\otimes_{i=1}^{N}\mathcal{H}_{2i}\right). A parallel strategy set 𝖯𝖺𝗋\mathsf{Par} is defined as the collection of P𝖲𝗍𝗋𝖺𝗍P\in\mathsf{Strat} such that [34]

TrFP=𝟙2,4,,2NP(1),TrP(1)=1.\Tr_{F}{P}=\mathds{1}_{2,4,\dots,2N}\otimes P^{(1)},\ \Tr P^{(1)}=1. (12)

Note that the optimal QFI of parallel strategies can also be evaluated using the method in [28, 30].

A more general protocol is to allow for sequential use of NN channels assisted by ancillae, where only the output of the former channel can affect the input of the latter channel, and any control gates can be inserted between channels [see Fig. 1LABEL:sub@subfig:sequential_strategy]. A sequential strategy set 𝖲𝖾𝗊\mathsf{Seq} is defined as the collection of P𝖲𝗍𝗋𝖺𝗍P\in\mathsf{Strat} such that [34]

TrFP=𝟙2NP(N),TrP(1)=1,\displaystyle\Tr_{F}P=\mathds{1}_{2N}\otimes P^{(N)},\ \Tr P^{(1)}=1, (13)
Tr2k1P(k)=𝟙2k2P(k1),k=2,,N.\displaystyle\Tr_{2k-1}P^{(k)}=\mathds{1}_{2k-2}\otimes P^{(k-1)},\ k=2,\dots,N.

Unlike the case of parallel strategies, there is no existing way of evaluating the exact QFI using sequential strategies.

We also consider families of strategies involving indefinite causal order. The first one, denoted by 𝖲𝖶𝖨\mathsf{SWI}, takes advantage of the (generalized) quantum SWITCH [48, 49], where the execution order of NN channels is entangled with the state of an N!N!-dimensional control system [see Fig. 1LABEL:sub@subfig:quantum_switch_strategy]. See Appendix E for the formal definition.

More generally, we consider the quantum superposition of multiple sequential orders, each with a unique order of querying the NN channels [see Fig. 1LABEL:sub@subfig:causal_superposition_strategy]. This can be implemented by entangling N!N! definite causal orders with a quantum control system [50]. If N=2N=2 and the control system is traced out, this notion is equivalent to causal separability [20, 21]. A causal superposition strategy set 𝖲𝗎𝗉\mathsf{Sup} is defined as the collection of P𝖲𝗍𝗋𝖺𝗍P\in\mathsf{Strat} such that

TrFP=πqπPπ,πSNqπ=1,\displaystyle\Tr_{F}P=\sum_{\pi}q^{\pi}P^{\pi},\ \sum_{\pi\in S_{N}}q^{\pi}=1, (14)
Pπ𝖲𝖾𝗊π,qπ0,πSN,\displaystyle\,P^{\pi}\in\mathsf{Seq}^{\pi},\ q^{\pi}\geq 0,\ \pi\in S_{N},

where each permutation π\pi is an element of the symmetric group SNS_{N} of degree NN, and each 𝖲𝖾𝗊π\mathsf{Seq}^{\pi} denotes a sequential strategy set whose execution order of NN channels is ϕπ(1)ϕπ(2)ϕπ(N)\mathcal{E}_{\phi}^{\pi(1)}\rightarrow\mathcal{E}_{\phi}^{\pi(2)}\rightarrow\cdots\rightarrow\mathcal{E}_{\phi}^{\pi(N)}, having denoted by ϕk\mathcal{E}_{\phi}^{k} the channel from (2k1)\mathcal{L}(\mathcal{H}_{2k-1}) to (2k)\mathcal{L}(\mathcal{H}_{2k}). Note that 𝖲𝖶𝖨\mathsf{SWI} is a subset of 𝖲𝗎𝗉\mathsf{Sup}, where the intermediate control is trivial. There are other strategies, such as quantum circuits with quantum controlled casual order (QC-QCs) and probabilistic QC-QCs [50, 51], which we will not discuss here.

Finally, we introduce the family of general indefinite-causal-order strategy [see Fig. 1LABEL:sub@subfig:general_strategy], which is the most general strategy set considered in this work. Here the only requirement is that the concatenation of the strategy PP with NN arbitrary channels results in a legitimate quantum state. The causal relations in this case [21] are a bit cumbersome, but for our purpose what matters is the dual affine space (see Theorem 1), which is simply the space of no-signaling channels [16, 43]. A general indefinite-causal-order strategy set 𝖨𝖢𝖮\mathsf{ICO} is defined as the collection of P𝖲𝗍𝗋𝖺𝗍P\in\mathsf{Strat} such that

ρF=P(j=1NEj),ρF0,TrρF=1,\rho_{F}=P*\left(\otimes_{j=1}^{N}E^{j}\right),\ \rho_{F}\geq 0,\ \Tr\rho_{F}=1, (15)

for any Ej(2j12jAj)E^{j}\in\mathcal{L}(\mathcal{H}_{2j-1}\otimes\mathcal{H}_{2j}\otimes\mathcal{H}_{A_{j}}) that denotes the CJ operator of an arbitrary quantum channel with an arbitrary ancillary space Aj\mathcal{H}_{A_{j}}.

We note that, unlike the previous strategies that can always be physically realized, the physical realization of the general 𝖨𝖢𝖮\mathsf{ICO} is untraceable [51, 50]. The optimal value obtained with general 𝖨𝖢𝖮\mathsf{ICO} nevertheless serves as a useful tool that can gauge the performances of different strategies. For example, as we will show, in some cases the optimal QFI J(𝖲𝗎𝗉)J^{(\mathsf{Sup})} and J(𝖨𝖢𝖮)J^{(\mathsf{ICO})} are equal or nearly equal. This then shows that the physically realizable strategy obtained from the set 𝖲𝗎𝗉\mathsf{Sup} is already optimal or nearly optimal among all possible strategies, which we will not be able to tell without J(𝖨𝖢𝖮)J^{(\mathsf{ICO})}.

Symmetry reduced programs for optimal metrology.—The complexity of the original optimization problems in Theorem 1 and Algorithm 1 can be reduced by exploiting the permutation symmetry. In Appendix D, we prove that we can choose a permutation-invariant matrix hh for Theorem 1 and solve for a permutation-invariant optimal strategy [52] by Algorithm 1 based on this choice, if any permutation πSN\pi\in S_{N} bijectively maps each affine space 𝖲i\mathsf{S}^{i} [in Eq. (9)] to some affine space 𝖲j\mathsf{S}^{j}. That is, for any πSN\pi\in S_{N} and any ii, there exists a jj such that the mapping SGπSGπS\mapsto G_{\pi}SG_{\pi}^{\dagger} on 𝖲i\mathsf{S}^{i} is a bijective function from 𝖲i\mathsf{S}^{i} to 𝖲j\mathsf{S}^{j}, where GπG_{\pi} is a unitary representation of π\pi. Furthermore, if each space 𝖲i\mathsf{S}^{i} itself is permutation invariant, we can restrict each Qi𝖲¯iQ^{i}\in\overline{\mathsf{S}}^{i} to be permutation invariant, further reducing the complexity of optimization. For both optimization problems we can apply the technique of group-invariant SDP to reduce the size as there exists an isomorphism which preserves positive semidefiniteness, from the permutation-invariant subspace to the space of block-diagonal matrices [53, Theorem 9.1]. Table 1 compares the number of variables involved in QFI evaluation with and without exploiting the symmetry (see Appendix E for its derivation as well as Appendix F for the complexity of Algorithm 1, where by group-invariant SDP we also numerically evaluate the growth of QFI J(𝖨𝖢𝖮)J^{(\mathsf{ICO})} up to N=5N=5).

Table 1: Complexity of QFI evaluation for each family of strategies (with repect to NN). The asymptotic numbers of variables in optimization are compared between the original (Ori.) and group-invariant (Inv.) SDP. We denote d:=𝖽𝗂𝗆(1)𝖽𝗂𝗆(2)d:=\mathsf{dim}(\mathcal{H}_{1})\mathsf{dim}(\mathcal{H}_{2}) and s:=maxϕ𝗋𝖺𝗇𝗄(Eϕ)ds:=\max_{\phi}\mathsf{rank}(E_{\phi})\leq d.
SDP 𝖯𝖺𝗋\mathsf{Par} 𝖲𝖾𝗊\mathsf{Seq} 𝖲𝖶𝖨\mathsf{SWI} 𝖲𝗎𝗉\mathsf{Sup} 𝖨𝖢𝖮\mathsf{ICO}
Ori. O(sN)O\left(s^{N}\right) O(dN)O\left(d^{N}\right) O(sN)O\left(s^{N}\right) O(N!dN)O\left(N!\,d^{N}\right) O(dN)O\left(d^{N}\right)
Inv. O(Nd21)O\left(N^{d^{2}-1}\right) O(dN)O\left(d^{N}\right) O(Ns21)O\left(N^{s^{2}-1}\right) O(dN)O\left(d^{N}\right) O(Nd21)O\left(N^{d^{2}-1}\right)

Hierarchy of strategies.—By substituting the definitions of different strategy sets into Theorem 1, we obtain the exact values of the optimal QFI. We find that a strict hierarchy of QFI exists quite prevalently. For demonstration purposes, here we show only the result for the amplitude damping channel for N=2N=2 and supplement our findings with bountiful numerical results in Appendix G. In this case, the process encoding ϕ\phi is a zz rotation Uz(ϕ)=eiϕtσz/2U_{z}(\phi)=e^{-\mathrm{i}\phi t\sigma_{z}/2}, where tt is the evolution time, followed by an amplitude damping channel described by two Kraus operators: K1(AD)=|00|+1p|11|K_{1}^{(\mathrm{AD})}=\outerproduct{0}{0}+\sqrt{1-p}\outerproduct{1}{1} and K2(AD)=p|01|K_{2}^{(\mathrm{AD})}=\sqrt{p}\outerproduct{0}{1}, with the decay parameter pp.

Refer to caption
Figure 2: Hierarchy of QFI using parallel, sequential, and indefinite-causal-order strategies. We take N=2N=2 and ϕ=1.0\phi=1.0, fix t=1.0t=1.0 and vary the decay parameter pp. The gaps can be seen more clearly by zooming in on the interval [0.35,0.45][0.35,0.45] of the value of pp. J(𝖲𝗎𝗉)=J(𝖨𝖢𝖮)J^{(\mathsf{Sup})}=J^{(\mathsf{ICO})} with an error tolerance of no more than 10810^{-8} in this case, but the gap between J(𝖲𝗎𝗉)J^{(\mathsf{Sup})} and J(𝖨𝖢𝖮)J^{(\mathsf{ICO})} could exist for larger NN or for other types of noise, which can be observed by randomly sampling noise channels.

In Fig. 2 we plot the QFI versus pp for the amplitude damping noise with all 5 strategy sets for N=2N=2. A strict hierarchy of 𝖯𝖺𝗋\mathsf{Par}, 𝖲𝖾𝗊\mathsf{Seq} and 𝖨𝖢𝖮\mathsf{ICO} holds if pp is neither 1 nor 0, i.e., J(𝖯𝖺𝗋)<J(𝖲𝖾𝗊)<J(𝖨𝖢𝖮)J^{(\mathsf{Par})}<J^{(\mathsf{Seq})}<J^{(\mathsf{ICO})}. This is in contrast to the asymptotic regime of NN\to\infty, where the relative difference between J(𝖲𝖾𝗊)J^{(\mathsf{Seq})} and J(𝖯𝖺𝗋)J^{(\mathsf{Par})} vanishes for this channel [45]. Besides, in this case general 𝖨𝖢𝖮\mathsf{ICO} cannot strictly outperform 𝖲𝗎𝗉\mathsf{Sup}, implying that causally superposing two sequential strategies is sufficient to achieve the general optimality in this particular scenario. The gap between J(𝖲𝗎𝗉)J^{(\mathsf{Sup})} and J(𝖨𝖢𝖮)J^{(\mathsf{ICO})}, however, could be observed for the same channel with larger NN or for other channels when N=2N=2 (see Appendix G). In fact, by randomly sampling noise channels from CPTP channel ensembles, we find that for 984 of 1000 random channels, a strict hierarchy J(𝖯𝖺𝗋)<J(𝖲𝖾𝗊)<J(𝖲𝗎𝗉)<J(𝖨𝖢𝖮)J^{(\mathsf{Par})}<J^{(\mathsf{Seq})}<J^{(\mathsf{Sup})}<J^{(\mathsf{ICO})} holds for N=2N=2, implying that there exist more powerful strategies than causal superposition strategies in these cases. We note that a strict hierarchy of strategies has been found for channel discrimination in [23], but much less is known in quantum metrology until our work.

Our method can also test the tightness of existing QFI bounds in the nonasymptotic regime, which has seldom been done until this work. Here we take the commonly used, asymptotically tight [45] upper bound for parallel strategies [see [28, Theorem 4 and Eq. (17)] or [30, Eq. (16)]]. For p=0.5p=0.5, our result shows that the exact parallel QFI J(𝖯𝖺𝗋)=1.795J^{(\mathsf{Par})}=1.795 is 32.7%32.7\% lower than the asymptotically tight parallel upper bound 2.6672.667, and even the exact sequential QFI J(𝖲𝖾𝗊)=2.179J^{(\mathsf{Seq})}=2.179 is 18.3%18.3\% lower than this parallel upper bound [54]. Similar phenomena are observed in other noise models and for different NN (see Appendix H).

With Algorithm 1 we can also construct strategies to achieve the optimal QFI. Remarkably, we find that a simple strategy of applying a quantum SWITCH using a control qubit |\plusC:=(|0C+|1C)/2\ket{\plus}_{C}:=\left(\ket{0}_{C}+\ket{1}_{C}\right)/\sqrt{2} (without any additional control operations on the probe) beats any sequential strategies (which can involve complex control) in certain cases (e.g. p<0.5p<0.5). To our best knowledge, this is the only instance of noisy quantum metrology so far, where the advantage of indefinite causal orders is established rigorously. In Appendix I we also present two explicit examples of implementing optimal sequential and causal superposition strategies, obtained by first applying Algorithm 1 and converting the CJ operators into quantum circuit consisting of single-qubit rotations and CNOT gates (as well as a quantum SWITCH for the case of 𝖲𝗎𝗉\mathsf{Sup}). For optimal causal superposition strategies, the permutation symmetry allows us to only control the execution order of channels while fixing state preparation and intermediate control [ρ=ρ\rho_{\uparrow}=\rho_{\downarrow} and U=UU_{\uparrow}=U_{\downarrow} in Fig. 1LABEL:sub@subfig:causal_superposition_strategy], which can be implemented by a (2N1)(2N-1)-quantum SWITCH of NN channels ϕ\mathcal{E}_{\phi} and N1N-1 intermediate operations.

Our result serves as a versatile tool for the demonstration of optimal quantum metrology and the design of optimal quantum sensors, especially in the context of control optimization [55, 56] and indefinite causal orders [20, 16, 21, 18, 22, 19, 17, 23].

The code accompanying the paper is openly available [57].

We thank Cyril Branciard, Alastair A. Abbott, and Raphael Mothe for helpful discussions and comments on our first manuscript. This work is supported by Guangdong Natural Science Fund—General Programme via Project 2022A1515010340, by HKU Seed Fund for Basic Research for New Staff via Project 202107185045, and by the Research Grants Council of Hong Kong through the Grant No. 14307420.

Note added.—Recently, it has been shown that neither sequential nor causal superposition strategies provide any advantage over parallel strategies asymptotically [58].

References

  • Giovannetti et al. [2011] V. Giovannetti, S. Lloyd, and L. Maccone, Advances in quantum metrology, Nat. Photonics 5, 222 (2011).
  • Degen et al. [2017] C. L. Degen, F. Reinhard, and P. Cappellaro, Quantum sensing, Rev. Mod. Phys. 89, 035002 (2017).
  • Martinis [2015] J. M. Martinis, Qubit metrology for building a fault-tolerant quantum computer, npj Quantum Inf. 1, 15005 (2015).
  • Lee et al. [2002] H. Lee, P. Kok, and J. P. Dowling, A quantum rosetta stone for interferometry, J. Mod. Opt. 49, 2325 (2002).
  • Bužek et al. [1999] V. Bužek, R. Derka, and S. Massar, Optimal Quantum Clocks, Phys. Rev. Lett. 82, 2207 (1999).
  • Kitagawa and Ueda [1993] M. Kitagawa and M. Ueda, Squeezed spin states, Phys. Rev. A 47, 5138 (1993).
  • Demkowicz-Dobrzański and Maccone [2014] R. Demkowicz-Dobrzański and L. Maccone, Using Entanglement Against Noise in Quantum Metrology, Phys. Rev. Lett. 113, 250801 (2014).
  • Yuan and Fung [2015] H. Yuan and C.-H. F. Fung, Optimal Feedback Scheme and Universal Time Scaling for Hamiltonian Parameter Estimation, Phys. Rev. Lett. 115, 110401 (2015).
  • Yuan [2016] H. Yuan, Sequential Feedback Scheme Outperforms the Parallel Scheme for Hamiltonian Parameter Estimation, Phys. Rev. Lett. 117, 160801 (2016).
  • Pang and Jordan [2017] S. Pang and A. N. Jordan, Optimal adaptive control for quantum metrology with time-dependent hamiltonians, Nat. Commun. 8, 14695 (2017).
  • Dür et al. [2014] W. Dür, M. Skotiniotis, F. Fröwis, and B. Kraus, Improved Quantum Metrology Using Quantum Error Correction, Phys. Rev. Lett. 112, 080801 (2014).
  • Kessler et al. [2014] E. M. Kessler, I. Lovchinsky, A. O. Sushkov, and M. D. Lukin, Quantum Error Correction for Metrology, Phys. Rev. Lett. 112, 150802 (2014).
  • Demkowicz-Dobrzański et al. [2017] R. Demkowicz-Dobrzański, J. Czajkowski, and P. Sekatski, Adaptive Quantum Metrology under General Markovian Noise, Phys. Rev. X 7, 041009 (2017).
  • Zhou et al. [2018] S. Zhou, M. Zhang, J. Preskill, and L. Jiang, Achieving the heisenberg limit in quantum metrology using quantum error correction, Nat. Commun. 9, 78 (2018).
  • Preskill [2018] J. Preskill, Quantum Computing in the NISQ era and beyond, Quantum 2, 79 (2018).
  • Chiribella et al. [2013] G. Chiribella, G. M. D’Ariano, P. Perinotti, and B. Valiron, Quantum computations without definite causal structure, Phys. Rev. A 88, 022318 (2013).
  • Chapeau-Blondeau [2021] F. Chapeau-Blondeau, Noisy quantum metrology with the assistance of indefinite causal order, Phys. Rev. A 103, 032615 (2021).
  • Mukhopadhyay et al. [2018] C. Mukhopadhyay, M. K. Gupta, and A. K. Pati, Superposition of causal order as a metrological resource for quantum thermometry (2018), arXiv:1812.07508 [quant-ph] .
  • Zhao et al. [2020] X. Zhao, Y. Yang, and G. Chiribella, Quantum Metrology with Indefinite Causal Order, Phys. Rev. Lett. 124, 190503 (2020).
  • Oreshkov et al. [2012] O. Oreshkov, F. Costa, and Č. Brukner, Quantum correlations with no causal order, Nat. Commun. 3, 1092 (2012).
  • Araújo et al. [2015] M. Araújo, C. Branciard, F. Costa, A. Feix, C. Giarmatzi, and Č. Brukner, Witnessing causal nonseparability, New J. Phys. 17, 102001 (2015).
  • Quintino et al. [2019] M. T. Quintino, Q. Dong, A. Shimbo, A. Soeda, and M. Murao, Reversing Unknown Quantum Transformations: Universal Quantum Circuit for Inverting General Unitary Operations, Phys. Rev. Lett. 123, 210502 (2019).
  • Bavaresco et al. [2021] J. Bavaresco, M. Murao, and M. T. Quintino, Strict Hierarchy between Parallel, Sequential, and Indefinite-Causal-Order Strategies for Channel Discrimination, Phys. Rev. Lett. 127, 200504 (2021).
  • Helstrom [1976] C. Helstrom, Quantum Detection and Estimation Theory (Academic Press, New York, 1976).
  • Holevo [2011] A. S. Holevo, Probabilistic and Statistical Aspects of Quantum Theory, Vol. 1 (Springer Science & Business Media, Berlin, 2011).
  • Braunstein and Caves [1994] S. L. Braunstein and C. M. Caves, Statistical distance and the geometry of quantum states, Phys. Rev. Lett. 72, 3439 (1994).
  • [27] For ν\nu\rightarrow\infty, the maximum likelihood estimator is unbiased and the quantum Cramér-Rao bound is achieved. This should not be confused with the nonasymptotic regime (where NN is finite).
  • Fujiwara and Imai [2008] A. Fujiwara and H. Imai, A fibre bundle over manifolds of quantum channels and its application to quantum statistics, J. Phys. A 41, 255304 (2008).
  • Escher et al. [2011] B. M. Escher, R. L. de Matos Filho, and L. Davidovich, General framework for estimating the ultimate precision limit in noisy quantum-enhanced metrology, Nat. Phys. 7, 406 (2011).
  • Demkowicz-Dobrzański et al. [2012] R. Demkowicz-Dobrzański, J. Kołodyński, and M. Guţă, The elusive heisenberg limit in quantum-enhanced metrology, Nat. Commun. 3, 1063 (2012).
  • Yuan and Fung [2017a] H. Yuan and C.-H. F. Fung, Fidelity and fisher information on quantum channels, New J. Phys. 19, 113039 (2017a).
  • Yuan and Fung [2017b] H. Yuan and C.-H. F. Fung, Quantum parameter estimation with general dynamics, npj Quantum Inf. 3, 14 (2017b).
  • Chiribella et al. [2008a] G. Chiribella, G. M. D’Ariano, and P. Perinotti, Quantum circuit architecture, Phys. Rev. Lett. 101, 060401 (2008a).
  • Chiribella et al. [2009] G. Chiribella, G. M. D’Ariano, and P. Perinotti, Theoretical framework for quantum networks, Phys. Rev. A 80, 022339 (2009).
  • [35] For any CJ operator PP of rank rPr_{P}, we can always choose an rPr_{P}-dimensional ancillary space to purify PP, which only extends the global future space F\mathcal{H}_{F} without changing the causal relations between queries to the channel ϕ\mathcal{E}_{\phi}. As QFI is monotonic under CPTP maps, the optimal QFI can be achieved by a purified process, and thus it is sufficient to only consider pure processes.
  • Petz [2008] D. Petz, Quantum Information Theory and Quantum Statistics (Springer, Berlin, 2008).
  • Yang [2019] Y. Yang, Memory Effects in Quantum Metrology, Phys. Rev. Lett. 123, 110501 (2019).
  • Altherr and Yang [2021] A. Altherr and Y. Yang, Quantum Metrology for Non-markovian Processes, Phys. Rev. Lett. 127, 060501 (2021).
  • Fan [1953] K. Fan, Minimax theorems, Proc. Natl. Acad. Sci. U.S.A. 39, 42 (1953).
  • Rockafellar [1970] R. T. Rockafellar, Convex Analysis (Princeton University Press, Princeton, NJ, 1970).
  • [41] We also suppose that the strategy set 𝖯\mathsf{P} is compact in order to apply Fan’s minimax theorem.
  • Chiribella [2012] G. Chiribella, Optimal networks for quantum metrology: Semidefinite programs and product rules, New J. Phys. 14, 125008 (2012).
  • Chiribella and Ebler [2016] G. Chiribella and D. Ebler, Optimal quantum networks and one-shot entropies, New J. Phys. 18, 093053 (2016).
  • Zhou and Jiang [2020] S. Zhou and L. Jiang, Optimal approximate quantum error correction for quantum metrology, Phys. Rev. Res. 2, 013235 (2020).
  • Zhou and Jiang [2021] S. Zhou and L. Jiang, Asymptotic theory of quantum channel estimation, PRX Quantum 2, 010343 (2021).
  • Bisio et al. [2011] A. Bisio, G. M. D’Ariano, P. Perinotti, and G. Chiribella, Minimal computational-space implementation of multiround quantum protocols, Phys. Rev. A 83, 022325 (2011).
  • Giovannetti et al. [2006] V. Giovannetti, S. Lloyd, and L. Maccone, Quantum Metrology, Phys. Rev. Lett. 96, 010401 (2006).
  • Colnaghi et al. [2012] T. Colnaghi, G. M. D’Ariano, S. Facchini, and P. Perinotti, Quantum computation with programmable connections between gates, Phys. Lett. A 376, 2940 (2012).
  • Araújo et al. [2014] M. Araújo, F. Costa, and v. Brukner, Computational Advantage from Quantum-Controlled Ordering of Gates, Phys. Rev. Lett. 113, 250402 (2014).
  • Wechs et al. [2021] J. Wechs, H. Dourdent, A. A. Abbott, and C. Branciard, Quantum circuits with classical versus quantum control of causal order, PRX Quantum 2, 030335 (2021).
  • Purves and Short [2021] T. Purves and A. J. Short, Quantum Theory Cannot Violate a Causal Inequality, Phys. Rev. Lett. 127, 110402 (2021).
  • [52] Concretely, P~(opt)=TrFP(opt)\tilde{P}^{(\mathrm{opt})}=\Tr_{F}P^{(\mathrm{opt})} is permutation invariant in this case.
  • Bachoc et al. [2012] C. Bachoc, D. C. Gijswijt, A. Schrijver, and F. Vallentin, Invariant semidefinite programs, in Handbook on Semidefinite, Conic and Polynomial Optimization, edited by M. F. Anjos and J. B. Lasserre (Springer US, Boston, MA, 2012) pp. 219–269.
  • [54] Note that the best existing upper bound on the QFI of sequential strategies [7, 45] is greater than or equal to that of parallel strategies and is thus omitted here. The sequential upper bound is also harder to evaluate numerically because it cannot be readily formulated as a convex optimization problem.
  • Hou et al. [2019] Z. Hou, R.-J. Wang, J.-F. Tang, H. Yuan, G.-Y. Xiang, C.-F. Li, and G.-C. Guo, Control-Enhanced Sequential Scheme for General Quantum Parameter Estimation at the Heisenberg Limit, Phys. Rev. Lett. 123, 040501 (2019).
  • Hou et al. [2021] Z. Hou, Y. Jin, H. Chen, J.-F. Tang, C.-J. Huang, H. Yuan, G.-Y. Xiang, C.-F. Li, and G.-C. Guo, “Super-Heisenberg” and Heisenberg Scalings Achieved Simultaneously in the Estimation of a Rotating Field, Phys. Rev. Lett. 126, 070503 (2021).
  • Liu et al. [2022] Q. Liu, Z. Hu, H. Yuan, and Y. Yang, Quantum channel estimation within constraints on strategies, https://github.com/qiushi-liu/strategies_in_metrology (2022).
  • Kurdzialek et al. [2022] S. Kurdzialek, W. Gorecki, F. Albarelli, and R. Demkowicz-Dobrzanski, Using adaptiveness and causal superpositions against noise in quantum metrology (2022), arXiv:2212.08106 [quant-ph] .
  • Watrous [2018] J. Watrous, The Theory of Quantum Information (Cambridge University Press, Cambridge, England, 2018).
  • Fulton and Harris [2013] W. Fulton and J. Harris, Representation Theory: A First Course, Graduate Texts in Mathematics (Springer New York, NY, 2013).
  • Schur [1901] I. Schur, Ueber eine Klasse von Matrizen, die sich einer gegebenen Matrix zuordnen lassen, Ph.D. dissertation, Friedrich-Wilhelms-Universität, Berlin (1901).
  • Montealegre-Mora et al. [2021] F. Montealegre-Mora, D. Rosset, J.-D. Bancal, and D. Gross, Certifying numerical decompositions of compact group representations (2021), arXiv:2101.12244 [math.RT] .
  • Rosset et al. [2021] D. Rosset, F. Montealegre-Mora, and J.-D. Bancal, Replab: A computational/numerical approach to representation theory, in Quantum Theory and Symmetries, edited by M. B. Paranjape, R. MacKenzie, Z. Thomova, P. Winternitz, and W. Witczak-Krempa (Springer International Publishing, Cham, 2021) pp. 643–653.
  • Chiribella et al. [2008b] G. Chiribella, G. M. D’Ariano, and P. Perinotti, Transforming quantum operations: Quantum supermaps, Europhys. Lett. 83, 30004 (2008b).
  • Horn and Zhang [2005] R. A. Horn and F. Zhang, Basic properties of the Schur complement, in The Schur Complement and Its Applications, edited by F. Zhang (Springer US, Boston, MA, 2005) pp. 17–46.
  • Diamond and Boyd [2016] S. Diamond and S. Boyd, CVXPY: A Python-embedded modeling language for convex optimization, J. Mach. Learn. Res. 17, 1 (2016).
  • Agrawal et al. [2018] A. Agrawal, R. Verschueren, S. Diamond, and S. Boyd, A rewriting system for convex optimization problems, J. Control Decis. 5, 42 (2018).
  • MOSEK ApS [2021] MOSEK ApS, The MOSEK Optimizer API for Python manual. Version 9.3. (2021).
  • Bruzda et al. [2009] W. Bruzda, V. Cappellini, H.-J. Sommers, and K. Życzkowski, Random quantum operations, Phys. Lett. A 373, 320 (2009).
  • Johansson et al. [2012] J. Johansson, P. Nation, and F. Nori, Qutip: An open-source python framework for the dynamics of open quantum systems, Comput. Phys. Commun. 183, 1760 (2012).
  • Johansson et al. [2013] J. Johansson, P. Nation, and F. Nori, Qutip 2: A python framework for the dynamics of open quantum systems, Comput. Phys. Commun. 184, 1234 (2013).
  • Iten et al. [2016] R. Iten, R. Colbeck, I. Kukuljan, J. Home, and M. Christandl, Quantum circuits for isometries, Phys. Rev. A 93, 032318 (2016).
  • Žnidarič et al. [2008] M. Žnidarič, O. Giraud, and B. Georgeot, Optimal number of controlled-not gates to generate a three-qubit state, Phys. Rev. A 77, 032320 (2008).
  • Plesch and Brukner [2011] M. Plesch and Č. Brukner, Quantum-state preparation with universal gate decompositions, Phys. Rev. A 83, 032302 (2011).
  • Iten et al. [2019] R. Iten, O. Reardon-Smith, E. Malvetti, L. Mondada, G. Pauvert, E. Redmond, R. S. Kohli, and R. Colbeck, Introduction to UniversalQCompiler (2019), arXiv:1904.01072 [quant-ph] .
\do@columngrid

one´

Appendix A Proof of Eq. (6) of the main text

The formalism of this proof has been developed for strategies following a definite causal order in [38], and the generalization to indefinite-causal-order strategies considered here is straightforward.

In [28], the QFI of a quantum state is expressed as a minimization problem

JQ(ρϕ)=4min|ψϕ,ii=1qTr(|ψ˙ϕ,iψ˙ϕ,i|),J_{Q}(\rho_{\phi})=4\min_{\ket{\psi_{\phi,i}}}\sum_{i=1}^{q}\Tr\left(\outerproduct{\dot{\psi}_{\phi,i}}{\dot{\psi}_{\phi,i}}\right), (16)

for any integer qmaxϕ𝗋𝖺𝗇𝗄(ρϕ)q\geq\max_{\phi}\mathsf{rank}(\rho_{\phi}), where {|ψϕ,i}\left\{\ket{\psi_{\phi,i}}\right\} is a set of unnormalized vectors such that ρϕ=i|ψϕ,iψϕ,i|\rho_{\phi}=\sum_{i}\outerproduct{\psi_{\phi,i}}{\psi_{\phi,i}}111We assume {|ψϕ,i}\left\{\ket{\psi_{\phi,i}}\right\} is continuously differentiable with respect to ϕ\phi.. In the main text, the QFI of NN quantum channels ϕ\mathcal{E}_{\phi} is defined as the QFI of the output state obtained from the concatenation of the CJ operator NϕN_{\phi} of NN quantum channels and an optimal strategy PP in a given strategy set 𝖯\mathsf{P}:

J(𝖯)(Nϕ):=maxP𝖯JQ(PNϕ).J^{(\mathsf{P})}(N_{\phi}):=\max_{P\in\mathsf{P}}J_{Q}(P*N_{\phi}). (17)

Due to the monotonicity [36] of the QFI under CPTP maps (e.g., the partial trace operation over ancillary space in this case), by choosing a proper global future space F\mathcal{H}_{F} an optimal PP can be taken as a pure process (rank-1 operator) denoted by |PP|\outerproduct{P}{P}. For a fixed P=|PP|P=\outerproduct{P}{P}, minimization over decompositions of PNϕP*N_{\phi} is equivalent to minimization over decompositions of NϕN_{\phi}, as 𝗋𝖺𝗇𝗄(PNϕ)𝗋𝖺𝗇𝗄(Nϕ)\mathsf{rank}(P*N_{\phi})\leq\mathsf{rank}(N_{\phi}).

As a positive semidefinite operator, NϕN_{\phi} has a decomposition as:

Nϕ=i=1r|Nϕ,iNϕ,i|,N_{\phi}=\sum_{i=1}^{r}\outerproduct{N_{\phi,i}}{N_{\phi,i}}, (18)

where r:=maxϕ𝗋𝖺𝗇𝗄(Nϕ)r:=\max_{\phi}\mathsf{rank}(N_{\phi})222We also assume {|Nϕ,i}\left\{\ket{N_{\phi,i}}\right\} is continuously differentiable with respect to ϕ\phi.. Note that the decomposition is nonunique. Defining 𝐍ϕ:=(|Nϕ,1,,|Nϕ,r)\mathbf{N}_{\phi}:=\left(\ket{N_{\phi,1}},\dots,\ket{N_{\phi,r}}\right), an arbitrary alternative decomposition 𝐍~ϕ:=(|N~ϕ,1,,|N~ϕ,r)\tilde{\mathbf{N}}_{\phi}:=\left(\ket{\tilde{N}_{\phi,1}},\dots,\ket{\tilde{N}_{\phi,r}}\right) can be related to 𝐍ϕ\mathbf{N}_{\phi} by 𝐍~ϕ=𝐍ϕVϕ\tilde{\mathbf{N}}_{\phi}=\mathbf{N}_{\phi}V_{\phi}, where VϕV_{\phi} is an r×rr\times r unitary matrix. Then the QFI of NN channels can be expressed as

J(𝖯)(Nϕ)=maxP𝖯minVϕTr[P(𝟙FΩϕ)],J^{(\mathsf{P})}(N_{\phi})=\max_{P\in\mathsf{P}}\min_{V_{\phi}}\Tr\left[P\left(\mathds{1}_{F}\otimes\Omega_{\phi}\right)\right], (19)

having defined the performance operator Ωϕ:=4(𝐍~˙ϕ𝐍~˙ϕ)T=4i=1r(|N~˙ϕ,iN~˙ϕ,i|)T\Omega_{\phi}:=4\left(\dot{\tilde{\mathbf{N}}}_{\phi}\dot{\tilde{\mathbf{N}}}_{\phi}^{\dagger}\right)^{T}=4\sum_{i=1}^{r}\left(\outerproduct{\dot{\tilde{N}}_{\phi,i}}{\dot{\tilde{N}}_{\phi,i}}\right)^{T}, where the superscript TT denotes the transpose. We can further define an r×rr\times r Hermitian matrix h:=iV˙ϕVϕh:=\mathrm{i}\dot{V}_{\phi}V_{\phi}^{\dagger} to take care of the freedom of choice for the decomposition 𝐍~ϕ\tilde{\mathbf{N}}_{\phi}, by noting that 𝐍~˙ϕ𝐍~˙ϕ=(𝐍˙ϕi𝐍ϕh)(𝐍˙ϕi𝐍ϕh)\dot{\tilde{\mathbf{N}}}_{\phi}\dot{\tilde{\mathbf{N}}}_{\phi}^{\dagger}=\left(\dot{\mathbf{N}}_{\phi}-\mathrm{i}\mathbf{N}_{\phi}h\right)\left(\dot{\mathbf{N}}_{\phi}-\mathrm{i}\mathbf{N}_{\phi}h\right)^{\dagger}. Now by redefining 𝐍~˙ϕ:=(𝐍˙ϕi𝐍ϕh)\dot{\tilde{\mathbf{N}}}_{\phi}:=\left(\dot{\mathbf{N}}_{\phi}-\mathrm{i}\mathbf{N}_{\phi}h\right), we rewrite Ωϕ\Omega_{\phi} as Ωϕ(h)\Omega_{\phi}(h) to explicitly manifest its dependence on hh. Hence, we finally arrive at Eq. (6) of the main text:

J(𝖯)(Nϕ)\displaystyle J^{(\mathsf{P})}(N_{\phi}) =maxP𝖯minhrTr{P[𝟙FΩϕ(h)]}\displaystyle=\max_{P\in\mathsf{P}}\min_{h\in\mathbb{H}_{r}}\Tr\left\{P\left[\mathds{1}_{F}\otimes\Omega_{\phi}(h)\right]\right\} (20)
=maxP~𝖯~minhrTr[P~Ωϕ(h)],\displaystyle=\max_{\tilde{P}\in\tilde{\mathsf{P}}}\min_{h\in\mathbb{H}_{r}}\Tr\left[\tilde{P}\Omega_{\phi}(h)\right],

where 𝖯~:={P~=TrFP|P𝖯}\tilde{\mathsf{P}}:=\left\{\tilde{P}=\Tr_{F}P\middle|P\in\mathsf{P}\right\}. ∎

Appendix B Proof of Theorem 1

Starting from Eq. (6) of the main text, we exchange the order of minimization and maximization thanks to Fan’s minimax theorem [39], since Tr[P~Ωϕ(h)]\Tr\left[\tilde{P}\Omega_{\phi}(h)\right] is concave on P~\tilde{P} and convex on hh, and 𝖯~\tilde{\mathsf{P}} is assumed to be a compact set. Then the problem of QFI evaluation can be rewritten as

J(𝖯)(Nϕ)=minhmaxP~𝖯~Tr[P~Ωϕ(h)].J^{(\mathsf{P})}(N_{\phi})=\min_{h}\max_{\tilde{P}\in\tilde{\mathsf{P}}}\Tr\left[\tilde{P}\Omega_{\phi}(h)\right]. (21)

Reformulating the condition of Theorem 1, we require that each operator P~𝖯~\tilde{P}\in\tilde{\mathsf{P}} can be written as a convex combination of positive semidefinite operators SiS^{i}, i=1,,Ki=1,\dots,K:

P~=i=1KqiSi,fori=1Kqi=1,qi0,Si0,Si𝖲i,i=1,,K,\tilde{P}=\sum_{i=1}^{K}q^{i}S^{i},\ \mathrm{for}\ \sum_{i=1}^{K}q^{i}=1,\ q^{i}\geq 0,\ S^{i}\geq 0,\ S^{i}\in\mathsf{S}^{i},\ i=1,\dots,K, (22)

where each 𝖲i\mathsf{S}^{i} is an affine space of Hermitian operators. Thus Eq. (21) can be reformulated as

J(𝖯)(Nϕ)=\displaystyle J^{(\mathsf{P})}(N_{\phi})= minhmaxP~Tr[P~Ωϕ(h)],\displaystyle\min_{h}\max_{\tilde{P}}\Tr\left[\tilde{P}\Omega_{\phi}(h)\right], (23)
s.t.\displaystyle\mathrm{s.t.} P~=i=1KqiSi,\displaystyle\,\tilde{P}=\sum_{i=1}^{K}q^{i}S^{i},
i=1Kqi=1,\displaystyle\sum_{i=1}^{K}q^{i}=1,
qi0,Si0,Si𝖲i,i=1,,K.\displaystyle\,q^{i}\geq 0,\ S^{i}\geq 0,\ S^{i}\in\mathsf{S}^{i},\ i=1,\dots,K.

For now we fix hh and consider the dual problem of maximization over 𝖯~\tilde{\mathsf{P}}. For each affine space 𝖲i\mathsf{S}^{i} we have defined its dual affine space 𝖲¯i\overline{\mathsf{S}}^{i}, whose dual affine space in turn is exactly 𝖲i\mathsf{S}^{i} [43]. Choose an affine basis {Qi,j}j=1Li\{Q^{i,j}\}_{j=1}^{L_{i}} for 𝖲¯i\overline{\mathsf{S}}^{i}, and the maximization problem is further expressed as

maxP~\displaystyle\max_{\tilde{P}} Tr[P~Ωϕ(h)],\displaystyle\Tr\left[\tilde{P}\Omega_{\phi}(h)\right], (24)
s.t.\displaystyle\mathrm{s.t.} P~=i=1KqiSi,\displaystyle\,\tilde{P}=\sum_{i=1}^{K}q^{i}S^{i},
i=1Kqi=1,\displaystyle\sum_{i=1}^{K}q^{i}=1,
qi0,Si0,Tr(SiQi,j)=1,i=1,,K,j=1,,Li.\displaystyle\,q^{i}\geq 0,\ S^{i}\geq 0,\ \Tr\left(S^{i}Q^{i,j}\right)=1,\ i=1,\dots,K,\ j=1,\dots,L_{i}.

Defining Pi:=qiSiP^{i}:=q^{i}S^{i} to avoid the product of variables in optimization, we have

max\displaystyle\max Tr[P~Ωϕ(h)],\displaystyle\Tr\left[\tilde{P}\Omega_{\phi}(h)\right], (25)
s.t.\displaystyle\mathrm{s.t.} P~=i=1KPi,\displaystyle\,\tilde{P}=\sum_{i=1}^{K}P^{i},
i=1Kqi=1,\displaystyle\sum_{i=1}^{K}q^{i}=1,
Pi0,Tr(PiQi,j)=qi,i=1,,K,j=1,,Li,\displaystyle\,P^{i}\geq 0,\ \Tr\left(P^{i}Q^{i,j}\right)=q^{i},\ i=1,\dots,K,\ j=1,\dots,L_{i},

where the constraints qi0q^{i}\geq 0 can be safely removed, since TrSi=j=1Nd2j\Tr S^{i}=\prod_{j=1}^{N}d_{2j}, implying that 𝖲¯i\overline{\mathsf{S}}^{i} includes a positive operator proportional to identity for any i=1,,Ki=1,\dots,K, having denoted dj:=𝖽𝗂𝗆(j)d_{j}:=\mathsf{dim}(\mathcal{H}_{j}) for simplicity. The Lagrangian of the problem is given by

L\displaystyle L =iTr[PiΩϕ(h)]+(1iqi)λ+iTr(PiQ~i)+i,j[qiTr(PiQi,j)]λi,j\displaystyle=\sum_{i}\Tr\left[P^{i}\Omega_{\phi}(h)\right]+\left(1-\sum_{i}q^{i}\right)\lambda+\sum_{i}\Tr\left(P^{i}\tilde{Q}^{i}\right)+\sum_{i,j}\left[q^{i}-\Tr\left(P^{i}Q^{i,j}\right)\right]\lambda^{i,j} (26)
=λ+iTr{Pi[Ωϕ(h)+Q~ijλi,jQi,j]}+i[qi(jλi,jλ)],\displaystyle=\lambda+\sum_{i}\Tr\left\{P^{i}\left[\Omega_{\phi}(h)+\tilde{Q}^{i}-\sum_{j}\lambda^{i,j}Q^{i,j}\right]\right\}+\sum_{i}\left[q^{i}\left(\sum_{j}\lambda^{i,j}-\lambda\right)\right],

for Q~i0\tilde{Q}^{i}\geq 0. Hence, by removing Q~i\tilde{Q}^{i} the dual problem is written as

min\displaystyle\min λ,\displaystyle\,\lambda, (27)
s.t.\displaystyle\mathrm{s.t.} jλi,jQi,jΩϕ(h),λ=jλi,j,i=1,,K,j=1,,Li.\displaystyle\sum_{j}\lambda^{i,j}Q^{i,j}\geq\Omega_{\phi}(h),\ \lambda=\sum_{j}\lambda^{i,j},\ i=1,\dots,K,\ j=1,\dots,L_{i}.

We define Qi:=jλi,jQi,j/λQ^{i}:=\sum_{j}\lambda^{i,j}Q^{i,j}/\lambda if λ0\lambda\neq 0 (λ=0\lambda=0 corresponds to a trivial case where the QFI is zero), and clearly QiQ^{i} is an arbitrary operator in the set 𝖲¯i\overline{\mathsf{S}}^{i}. Therefore, we cast the dual problem into

min\displaystyle\min λ,\displaystyle\,\lambda, (28)
s.t.\displaystyle\mathrm{s.t.} λQiΩϕ(h),Qi𝖲¯i,i=1,,K.\displaystyle\,\lambda Q^{i}\geq\Omega_{\phi}(h),\ Q^{i}\in\overline{\mathsf{S}}^{i},\ i=1,\dots,K.

Slater’s theorem [59] implies that the strong duality holds, since the QFI is finite and the inequality constraints can be strictly satisfied for a positive semidefinite operator Ωϕ(h)\Omega_{\phi}(h), by choosing λQi=μΩϕ(h)𝟙1,2,,2N\lambda Q^{i}=\mu\lVert\Omega_{\phi}(h)\rVert\mathds{1}_{1,2,\dots,2N} for μ>1\mu>1 and any i=1,,Ki=1,\dots,K, having denoted the operator norm by \lVert\cdot\rVert. Finally, by optimizing the choice of hh we derive the result of Theorem 1. ∎

Appendix C Proof of the validity of Algorithm 1

We first recall the minimax theorem:

minxmaxyf(x,y)=maxyminxf(x,y)\min_{x}\max_{y}f(x,y)=\max_{y}\min_{x}f(x,y) (29)

for a function f(x,y)f(x,y) convex on xx and concave on yy. Assume (x0,y1)(x_{0},y_{1}) is a solution for the L.H.S. of Eq. (29) and (x1,y0)(x_{1},y_{0}) is a solution for the R.H.S. of Eq. (29). It is easy to see that

f(x0,y1)f(x0,y0)f(x1,y0).f(x_{0},y_{1})\geq f(x_{0},y_{0})\geq f(x_{1},y_{0}). (30)

In view of Eq. (29) both equalities hold. Therefore, (x0,y0)(x_{0},y_{0}) is a saddle point of f(x,y)f(x,y), i.e., x0=argminxf(x,y0)x_{0}=\operatorname*{arg\,min}_{x}{f(x,y_{0})} and y0=argmaxyf(x0,y)y_{0}=\operatorname*{arg\,max}_{y}{f(x_{0},y)}. Since the objective function Tr[P~Ωϕ(h)]\Tr\left[\tilde{P}\Omega_{\phi}(h)\right] in the primal problem of estimating QFI is convex on hh and concave on P~\tilde{P}, we can substitute x=hx=h and y=P~y=\tilde{P}. Obviously, h(opt)h^{(\mathrm{opt})} is an optimal solution for minhmaxP~Tr[P~Ωϕ(h)]\min_{h}\max_{\tilde{P}}\Tr\left[\tilde{P}\Omega_{\phi}(h)\right] and thus corresponds to a saddle point. Then x0=argminxf(x,y0)x_{0}=\operatorname*{arg\,min}_{x}{f(x,y_{0})} can be satisfied by requiring hf(h,P~(opt))|h=h(opt)=0\partial_{h}f\left(h,\tilde{P}^{(\mathrm{opt})}\right)|_{h=h^{(\mathrm{opt})}}=0, resulting in Eq. (11) in the main text:

Re{Tr{P~(opt)[i𝐍ϕ(𝐍˙ϕi𝐍ϕh(opt))]T}}=0forallr.\real\left\{\Tr\left\{\tilde{P}^{(\mathrm{opt})}\left[-\mathrm{i}\mathbf{N}_{\phi}\mathscr{H}\left(\dot{\mathbf{N}}_{\phi}-\mathrm{i}\mathbf{N}_{\phi}h^{(\mathrm{opt})}\right)^{\dagger}\right]^{T}\right\}\right\}=0\ \mathrm{for}\ \mathrm{all}\ \mathscr{H}\in\mathbb{H}_{r}. (31)

Meanwhile y0=argmaxyf(x0,y)y_{0}=\operatorname*{arg\,max}_{y}{f(x_{0},y)} corresponds to the requirement that P~(opt)\tilde{P}^{(\mathrm{opt})} is an optimal solution for fixed h=h(opt)h=h^{(\mathrm{opt})}. Therefore, (h(opt),P~(opt))\left(h^{(\mathrm{opt})},\tilde{P}^{(\mathrm{opt})}\right) is a saddle point and an optimal solution for maxP~minhTr[P~Ωϕ(h)]\max_{\tilde{P}}\min_{h}\Tr\left[\tilde{P}\Omega_{\phi}(h)\right]. By definition a purification of P~(opt)\tilde{P}^{(\mathrm{opt})} is an optimal physically implemented strategy, i.e., we can choose a strategy P(opt)P^{(\mathrm{opt})} such that TrFP(opt)=P~(opt)\Tr_{F}P^{(\mathrm{opt})}=\tilde{P}^{(\mathrm{opt})}. ∎

We remark that Eq. (31) can be reformulated as a set of linear constraints by choosing a basis {i}i=1r2\{\mathscr{H}^{i}\}_{i=1}^{r^{2}} for the space r\mathbb{H}_{r} of Hemitian matrices. For example, denoting by EijE_{ij} the r×rr\times r matrix of which only the (i,j)(i,j)-th element is 11 and all other elements are 0, we can choose =Ejj\mathscr{H}=E_{jj} for j=1,,rj=1,\dots,r, =Ejk+Ekj\mathscr{H}=E_{jk}+E_{kj} and =i(EjkEkj)\mathscr{H}=\mathrm{i}\left(E_{jk}-E_{kj}\right) for k=1,,rk=1,\dots,r and kjk\neq j, and obtain a series of constraints, which are equivalent to Eq.(31).

Appendix D Symmetry reduced optimization

This section demonstrates how to reduce the size of the optimization problems concerned in Theorem 1 and Algorithm 1, by exploiting the permutation symmetry as applicable.

Let us begin with some notations. We consider the action of SNS_{N}, the symmetric group of degree NN, on the finite-dimensional representation space i=1N𝒲i\otimes_{i=1}^{N}\mathcal{W}_{i}. GπG_{\pi} is a unitary (and orthogonal) operator on i=1N𝒲i\otimes_{i=1}^{N}\mathcal{W}_{i} corresponding to the permutation πSN\pi\in S_{N}: Gπ=𝐢=(i1,,iN)(j=1N|iπ(j))k=1Nik|G_{\pi}=\sum_{\mathbf{i}=(i_{1},\dots,i_{N})}\left(\otimes_{j=1}^{N}\ket{i_{\pi(j)}}\right)\otimes_{k=1}^{N}\bra{i_{k}}, where {|ij}i\{\ket{i_{j}}\}_{i} denotes an orthonormal basis of 𝒲j\mathcal{W}_{j}. Then an operator XX on i=1N𝒲i\otimes_{i=1}^{N}\mathcal{W}_{i} is said to be permutation invariant iff GπXGπ=XG_{\pi}XG_{\pi}^{\dagger}=X for all πSN\pi\in S_{N}. Analogously, a space 𝒳\mathcal{X} is permutation invariant iff GπXGπ𝒳G_{\pi}XG_{\pi}^{\dagger}\in\mathcal{X} for any X𝒳X\in\mathcal{X} and any πSN\pi\in S_{N}.

We now explicitly express the components of the performance operator Ωϕ(h)\Omega_{\phi}(h). Given NN identical quantum channels ϕ(ρ)=iKϕ,iρKϕ,i\mathcal{E}_{\phi}(\rho)=\sum_{i}K_{\phi,i}^{\dagger}\rho K_{\phi,i}, we decompose the CJ operator Eϕ=i|Eϕ,iEϕ,i|E_{\phi}=\sum_{i}\outerproduct{E_{\phi,i}}{E_{\phi,i}} corresponding to each channel. Note that |Eϕ,i\ket{E_{\phi,i}} is the vectorization of the Kraus operator Kϕ,iK_{\phi,i}, i.e., |Eϕ,i=m,nm|Kϕ,i|n|m|n\ket{E_{\phi,i}}=\sum_{m,n}\matrixelement{m}{K_{\phi,i}}{n}\ket{m}\ket{n}. The CJ operator of NN identical quantum channels is Nϕ=EϕN=𝐢|Nϕ,𝐢Nϕ,𝐢|N_{\phi}=E_{\phi}^{\otimes N}=\sum_{\mathbf{i}}\outerproduct{N_{\phi,\mathbf{i}}}{N_{\phi,\mathbf{i}}}, where we use the notation |Nϕ,𝐢=(i1,,iN)=n=1N|Eϕ,in\ket{N_{\phi,\mathbf{i}=(i_{1},\dots,i_{N})}}=\otimes_{n=1}^{N}\ket{E_{\phi,i_{n}}}. Taking the derivative results in (we omit the subscript ϕ\phi in |Eϕ,i\ket{E_{\phi,i}}) |N˙ϕ,𝐢=j=1N|Ei1|Eij1|E˙ij|Eij+1|EiN\ket{\dot{N}_{\phi,\mathbf{i}}}=\sum_{j=1}^{N}\ket{E_{i_{1}}}\cdots\ket{E_{i_{j-1}}}\ket{\dot{E}_{i_{j}}}\ket{E_{i_{j+1}}}\cdots\ket{E_{i_{N}}}, from which we can obtain the performance operator Ωϕ(h)=4(𝐍~˙ϕ𝐍~˙ϕ)T=4𝐢(|N~˙ϕ,𝐢N~˙ϕ,𝐢|)T\Omega_{\phi}(h)=4\left(\dot{\tilde{\mathbf{N}}}_{\phi}\dot{\tilde{\mathbf{N}}}_{\phi}^{\dagger}\right)^{T}=4\sum_{\mathbf{i}}\left(\outerproduct{\dot{\tilde{N}}_{\phi,\mathbf{i}}}{\dot{\tilde{N}}_{\phi,\mathbf{i}}}\right)^{T}, where |N~˙ϕ,𝐣=|N˙ϕ,𝐣i𝐤|Nϕ,𝐤h𝐤𝐣\ket{\dot{\tilde{N}}_{\phi,\mathbf{j}}}=\ket{\dot{N}_{\phi,\mathbf{j}}}-\mathrm{i}\sum_{\mathbf{k}}\ket{N_{\phi,\mathbf{k}}}h_{\mathbf{k}\mathbf{j}}.

D.1 Symmetry reduced QFI evaluation

With these notations, we first consider the optimization in QFI evaluation and have the following:

Lemma 1.

In the optimization problem of Theorem 1, if, for any πSN\pi\in S_{N} and any ii, there exists a jj such that the mapping SGπSGπS\mapsto G_{\pi}SG_{\pi}^{\dagger} on 𝖲i\mathsf{S}^{i} is a bijective function from 𝖲i\mathsf{S}^{i} to 𝖲j\mathsf{S}^{j}, then there must exist a permutation-invariant hh as a feasible solution.

Proof.

Without loss of generality we assume all 𝖲i\mathsf{S}^{i} are distinct spaces; otherwise, we just remove the duplicate ones. We first prove that, under any permutation operation QGπQGπQ\mapsto G_{\pi}^{\dagger}QG_{\pi}, for any i{1,,K}i\in\{1,\dots,K\} there exists a unique j{1,,K}j\in\{1,\dots,K\} such that the dual affine space 𝖲¯j\overline{\mathsf{S}}^{j} is bijectively mapped to 𝖲¯i\overline{\mathsf{S}}^{i}. The condition of the lemma implies that, for any GπG_{\pi} and any ii we can find jij_{i} such that GπSiGπ𝖲jiG_{\pi}S^{i}G_{\pi}^{\dagger}\in\mathsf{S}^{j_{i}} for all Si𝖲iS^{i}\in\mathsf{S}^{i}. Due to the bijectivity, 𝖲i\mathsf{S}^{i} and the corresponding 𝖲ji\mathsf{S}^{j_{i}} are isomorphic, with different jij_{i} for different ii. Apparently 𝖲¯i\overline{\mathsf{S}}^{i} and 𝖲¯ji\overline{\mathsf{S}}^{j_{i}} are also isomorphic. If we choose any Qji𝖲¯jiQ^{j_{i}}\in\overline{\mathsf{S}}^{j_{i}}, i.e., Tr(QjiSji)=1\Tr(Q^{j_{i}}S^{j_{i}})=1 for all Sji𝖲jiS^{j_{i}}\in\mathsf{S}^{j_{i}}, then we have Tr(GπQjiGπSi)=Tr(QjiGπSiGπ)=1\Tr(G_{\pi}^{\dagger}Q^{j_{i}}G_{\pi}S^{i})=\Tr(Q^{j_{i}}G_{\pi}S^{i}G_{\pi}^{\dagger})=1 for all Si𝖲iS^{i}\in\mathsf{S}^{i}. Therefore, GπQjiGπ𝖲¯iG_{\pi}^{\dagger}Q^{j_{i}}G_{\pi}\in\overline{\mathsf{S}}^{i} for all Qji𝖲¯jiQ^{j_{i}}\in\overline{\mathsf{S}}^{j_{i}}. Furthermore, the permutation operation QGπQGπQ\mapsto G_{\pi}^{\dagger}QG_{\pi} from 𝖲¯ji\overline{\mathsf{S}}^{j_{i}} to 𝖲¯i\overline{\mathsf{S}}^{i} is bijective as 𝖲¯ji\overline{\mathsf{S}}^{j_{i}} and 𝖲¯i\overline{\mathsf{S}}^{i} are isomorphic. In particular, any set of KK operators {Qi}i=1K\{Q^{i}\}_{i=1}^{K} for Qi𝖲¯iQ^{i}\in\overline{\mathsf{S}}^{i} is mapped to a set {Q~i}i=1K\{\tilde{Q}^{i}\}_{i=1}^{K} such that Q~i𝖲¯i\tilde{Q}^{i}\in\overline{\mathsf{S}}^{i}.

We then prove that there exists a permutation-invariant performance operator Ωϕ(h)\Omega_{\phi}(h) as a feasible solution. The group action is characterized by the permutation operator GπG_{\pi} on the representation space i𝒲i=i(2i12i)\otimes_{i}\mathcal{W}_{i}=\otimes_{i}\left(\mathcal{H}_{2i-1}\otimes\mathcal{H}_{2i}\right). Suppose h(opt)h^{(\mathrm{opt})} is an optimal solution, i.e., for any i{1,,K}i\in\{1,\dots,K\}, there exist optimal values of λ\lambda and QiQ^{i} such that λQiΩϕ(h(opt))\lambda Q^{i}\geq\Omega_{\phi}\left(h^{(\mathrm{opt})}\right) for Qi𝖲¯iQ^{i}\in\overline{\mathsf{S}}^{i}. Then for any permutation πSN\pi\in S_{N} we have

λGπQiGπGπΩϕ(h(opt))Gπ,Qi𝖲¯i,i=1,,K.\lambda G_{\pi}Q^{i}G_{\pi}^{\dagger}\geq G_{\pi}\Omega_{\phi}\left(h^{(\mathrm{opt})}\right)G_{\pi}^{\dagger},\ Q^{i}\in\overline{\mathsf{S}}^{i},\ i=1,\dots,K. (32)

Since GπG_{\pi} maps {Qi}i\{Q^{i}\}_{i} to a set {Q~i}i\{\tilde{Q}^{i}\}_{i} such that Q~i𝖲¯i\tilde{Q}^{i}\in\overline{\mathsf{S}}^{i}, the constraint Eq. (32) becomes

λQ~iGπΩϕ(h(opt))Gπ,Q~i𝖲¯i,i=1,,K,\lambda\tilde{Q}^{i}\geq G_{\pi}\Omega_{\phi}\left(h^{(\mathrm{opt})}\right)G_{\pi}^{\dagger},\ \tilde{Q}^{i}\in\overline{\mathsf{S}}^{i},\ i=1,\dots,K, (33)

which implies that GπΩϕ(h(opt))GπG_{\pi}\Omega_{\phi}(h^{(\mathrm{opt})})G_{\pi}^{\dagger} is also a feasible solution for any π\pi. Furthermore, the permutation-invariant solution of the performance operator 1N!πSNGπΩϕ(h(opt))Gπ\frac{1}{N!}\sum_{\pi\in S_{N}}G_{\pi}\Omega_{\phi}(h^{(\mathrm{opt})})G_{\pi}^{\dagger} is feasible.

Next, we show that we can choose a permutation-invariant hh such that the performance operator Ωϕ(h)=1N!πSNGπΩϕ(h(opt))Gπ\Omega_{\phi}(h)=\frac{1}{N!}\sum_{\pi\in S_{N}}G_{\pi}\Omega_{\phi}\left(h^{(\mathrm{opt})}\right)G_{\pi}^{\dagger}. hh is an operator on i=1Ns\otimes_{i=1}^{N}\mathbb{C}^{s}, where s:=maxϕ𝗋𝖺𝗇𝗄(Eϕ)ds:=\max_{\phi}\mathsf{rank}(E_{\phi})\leq d. For distinction we denote the group action on hh by GπhGπG_{\pi}^{\prime}hG_{\pi}^{\prime\dagger} and the group action on Ωϕ(h)\Omega_{\phi}(h) by GπΩϕ(h)GπG_{\pi}\Omega_{\phi}(h)G_{\pi}^{\dagger}. Writing hh in components, we have

𝐢,𝐣Gπh𝐢𝐣|𝐢𝐣|Gπ\displaystyle\sum_{\mathbf{i},\mathbf{j}}G_{\pi}^{\prime}h_{\mathbf{i}\mathbf{j}}\outerproduct{\mathbf{i}}{\mathbf{j}}G_{\pi}^{\prime\dagger} =𝐢,𝐣h𝐢𝐣|π(𝐢)π(𝐣)|\displaystyle=\sum_{\mathbf{i},\mathbf{j}}h_{\mathbf{i}\mathbf{j}}\outerproduct{\pi(\mathbf{i})}{\pi(\mathbf{j})} (34)
=π1(𝐢),π1(𝐣)hπ1(𝐢)π1(𝐣)|𝐢𝐣|\displaystyle=\sum_{\pi^{-1}(\mathbf{i}),\pi^{-1}(\mathbf{j})}h_{\pi^{-1}(\mathbf{i})\pi^{-1}(\mathbf{j})}\outerproduct{\mathbf{i}}{\mathbf{j}}
=𝐢,𝐣hπ1(𝐢)π1(𝐣)|𝐢𝐣|,\displaystyle=\sum_{\mathbf{i},\mathbf{j}}h_{\pi^{-1}(\mathbf{i})\pi^{-1}(\mathbf{j})}\outerproduct{\mathbf{i}}{\mathbf{j}},

where π(𝐢):=(iπ(1),,iπ(N))\pi(\mathbf{i}):=(i_{\pi(1)},\dots,i_{\pi(N)}). Note that Gπ|Nϕ,𝐢=|Nϕ,π(𝐢)G_{\pi}\ket{N_{\phi,\mathbf{i}}}=\ket{N_{\phi,\pi(\mathbf{i})}} and

Gπ|N˙ϕ,𝐢\displaystyle G_{\pi}\ket{\dot{N}_{\phi,\mathbf{i}}} =j=1N|Ei1|Eij1|E˙ij|Eij+1|EiN\displaystyle=\sum_{j=1}^{N}\ket{E_{i_{1}}}\cdots\ket{E_{i_{j-1}}}\ket{\dot{E}_{i_{j}}}\ket{E_{i_{j+1}}}\cdots\ket{E_{i_{N}}} (35)
=j=1N|Eiπ(1)|Eiπ[π1(j)1]|E˙ij|Eiπ[π1(j)+1]|Eiπ(N)\displaystyle=\sum_{j=1}^{N}\ket{E_{i_{\pi(1)}}}\cdots\ket{E_{i_{\pi[\pi^{-1}(j)-1]}}}\ket{\dot{E}_{i_{j}}}\ket{E_{i_{\pi[\pi^{-1}(j)+1]}}}\cdots\ket{E_{i_{\pi(N)}}}
=j=1N|Eiπ(1)|Eiπ(j1)|E˙iπ(j)|Eiπ(j+1)|Eiπ(N)\displaystyle=\sum_{j=1}^{N}\ket{E_{i_{\pi(1)}}}\cdots\ket{E_{i_{\pi(j-1)}}}\ket{\dot{E}_{i_{\pi(j)}}}\ket{E_{i_{\pi(j+1)}}}\cdots\ket{E_{i_{\pi(N)}}}
=|N˙ϕ,π(𝐢),\displaystyle=\ket{\dot{N}_{\phi,\pi(\mathbf{i})}},

which results in

GπΩϕ(h)Gπ\displaystyle G_{\pi}\Omega_{\phi}(h)G_{\pi}^{\dagger} =4𝐣Gπ(|N~˙ϕ,𝐣N~˙ϕ,𝐣|)TGπ\displaystyle=4\sum_{\mathbf{j}}G_{\pi}\left(\outerproduct{\dot{\tilde{N}}_{\phi,\mathbf{j}}}{\dot{\tilde{N}}_{\phi,\mathbf{j}}}\right)^{T}G_{\pi}^{\dagger} (36)
=4𝐣Gπ|N~˙ϕ,𝐣N~˙ϕ,𝐣|Gπ\displaystyle=4\sum_{\mathbf{j}}G_{\pi}\outerproduct{\dot{\tilde{N}}_{\phi,\mathbf{j}}^{*}}{\dot{\tilde{N}}_{\phi,\mathbf{j}}^{*}}G_{\pi}^{\dagger}
=4𝐣Gπ(|N˙ϕ,𝐣+i𝐤|Nϕ,𝐤h𝐤𝐣)(N˙ϕ,𝐣|i𝐥h𝐥𝐣Nϕ,𝐥|)Gπ\displaystyle=4\sum_{\mathbf{j}}G_{\pi}\left(\ket{\dot{N}_{\phi,\mathbf{j}}^{*}}+\mathrm{i}\sum_{\mathbf{k}}\ket{N_{\phi,\mathbf{k}}^{*}}h_{\mathbf{k}\mathbf{j}}^{*}\right)\left(\bra{\dot{N}_{\phi,\mathbf{j}}^{*}}-\mathrm{i}\sum_{\mathbf{l}}h_{\mathbf{l}\mathbf{j}}\bra{N_{\phi,\mathbf{l}}^{*}}\right)G_{\pi}^{\dagger}
=4𝐣(|N˙ϕ,π(𝐣)+i𝐤|Nϕ,π(𝐤)h𝐤𝐣)(N˙ϕ,π(𝐣)|i𝐥h𝐥𝐣Nϕ,π(𝐥)|)\displaystyle=4\sum_{\mathbf{j}}\left(\ket{\dot{N}_{\phi,\pi(\mathbf{j})}^{*}}+\mathrm{i}\sum_{\mathbf{k}}\ket{N_{\phi,\pi(\mathbf{k})}^{*}}h_{\mathbf{k}\mathbf{j}}^{*}\right)\left(\bra{\dot{N}_{\phi,\pi(\mathbf{j})}^{*}}-\mathrm{i}\sum_{\mathbf{l}}h_{\mathbf{l}\mathbf{j}}\bra{N_{\phi,\pi(\mathbf{l})}^{*}}\right)
=4𝐣(|N˙ϕ,𝐣+i𝐤|Nϕ,𝐤hπ1(𝐤)π1(𝐣))(N˙ϕ,𝐣|i𝐥hπ1(𝐥)π1(𝐣)Nϕ,𝐥|)\displaystyle=4\sum_{\mathbf{j}}\left(\ket{\dot{N}_{\phi,\mathbf{j}}^{*}}+\mathrm{i}\sum_{\mathbf{k}}\ket{N_{\phi,\mathbf{k}}^{*}}h_{\pi^{-1}(\mathbf{k})\pi^{-1}(\mathbf{j})}^{*}\right)\left(\bra{\dot{N}_{\phi,\mathbf{j}}^{*}}-\mathrm{i}\sum_{\mathbf{l}}h_{\pi^{-1}(\mathbf{l})\pi^{-1}(\mathbf{j})}\bra{N_{\phi,\mathbf{l}}^{*}}\right)
=Ωϕ(GπhGπ),\displaystyle=\Omega_{\phi}\left(G_{\pi}^{\prime}hG_{\pi}^{\prime\dagger}\right),

where in the last equation we have used Eq. (34). Therefore, by choosing the permutation-invariant solution h=1N!πSNGπh(opt)Gπh=\frac{1}{N!}\sum_{\pi\in S_{N}}G_{\pi}^{\prime}h^{(\mathrm{opt})}G_{\pi}^{\prime\dagger} we obtain Ωϕ(h)=1N!πSNGπΩϕ(h(opt))Gπ\Omega_{\phi}(h)=\frac{1}{N!}\sum_{\pi\in S_{N}}G_{\pi}\Omega_{\phi}(h^{(\mathrm{opt})})G_{\pi}^{\dagger}, and this permutation-invariant choice of hh is also a feasible solution. ∎

If a stronger condition holds, then not only can we choose a permutation-invariant hh, but we can also restrict Qi𝖲¯iQ^{i}\in\overline{\mathsf{S}}^{i} to be permutation invariant. In this case all matrix variables concerned in the optimization are permutation invariant.

Lemma 2.

In the optimization problem of Theorem 1, if each affine space 𝖲i\mathsf{S}^{i} is permutation invariant for any i=1,,Ki=1,\dots,K, then there must exist a permutation-invariant Qi𝖲¯iQ^{i}\in\overline{\mathsf{S}}^{i} as a feasible solution for each ii.

Proof.

By Lemma 1 we can restrict hh to be permutation invariant in optimization, and therefore the performance operator Ωϕ(h)\Omega_{\phi}(h) is permutation invariant. It is easy to see that each dual affine space 𝖲¯i\overline{\mathsf{S}}^{i} is also permutation invariant. Then for any i=1,,Ki=1,\dots,K, for Qi,(opt)𝖲¯iQ^{i,(\mathrm{opt})}\in\overline{\mathsf{S}}^{i} satisfying the constraint λQiΩϕ(h)\lambda Q^{i}\geq\Omega_{\phi}(h), we have the permutation-invariant solution Qi=1N!πSNGπQi,(opt)Gπ𝖲¯iQ^{i}=\frac{1}{N!}\sum_{\pi\in S_{N}}G_{\pi}Q^{i,(\mathrm{opt})}G_{\pi}^{\dagger}\in\overline{\mathsf{S}}^{i} which also satisfies the same constraint. Hence, this permutation-invariant choice of QiQ^{i} is also a feasible solution. ∎

Now since the optimization is restricted to the invariant subspace, we can reduce the matrix sizes by block diagonalization. Generally, for a group-invariant space of complex matrices n×n,(inv)\mathbb{C}^{n\times n,(\mathrm{inv})}, there exists an isomorphism φ\varphi preserving positive semidefiniteness between n×n,(inv)\mathbb{C}^{n\times n,(\mathrm{inv})} and a direct sum of complex matrix spaces [53, Theorem 9.1]:

φ:n×n,(inv)k=1Imk×mk,\varphi:\mathbb{C}^{n\times n,(\mathrm{inv})}\rightarrow\bigoplus_{k=1}^{I}\mathbb{C}^{m_{k}\times m_{k}}, (37)

where mkm_{k} is the multiplicity of the kk-th inequivalent irreducible representation of the group, and II is the number of inequivalent irreducible representations. To be more specific, for the symmetric group SNS_{N}, the representation space 𝒲=(d)N\mathcal{W}=(\mathbb{C}^{d})^{\otimes N} can be decomposed into |μ|=N𝒲μ\bigoplus_{|\mu|=N}\mathcal{W}^{\mu}, where each partition μ:=(μ1,,μd)\mu:=(\mu_{1},\dots,\mu_{d}) (with nonnegative integers μiμj\mu_{i}\geq\mu_{j} for any i<ji<j) corresponds to a Young diagram, and |μ|:=iμi|\mu|:=\sum_{i}\mu_{i}. Each isotypic component 𝒲μ\mathcal{W}^{\mu} can be further decomposed into a direct sum of equivalent irreducible subspaces: 𝒲μ=i=1mμ𝒲μ,i\mathcal{W}^{\mu}=\bigoplus_{i=1}^{m_{\mu}}\mathcal{W}^{\mu,i}. We define the dimension of the irreducible representation dμ:=𝖽𝗂𝗆(𝒲μ,i)d_{\mu}:=\mathsf{dim}\left(\mathcal{W}^{\mu,i}\right). It is worth mentioning that the first decomposition into isotypic components is unique while the second decomposition into equivalent irreducible representation spaces is not (see, e.g., [60]). In terms of the unitary group action GπG_{\pi} on 𝒲\mathcal{W}, there exists a unitary transformation of basis UU such that for any π\pi we have

UGπU=|μ|=NGπμ𝟙(mμ),U^{\dagger}G_{\pi}U=\bigoplus_{|\mu|=N}G_{\pi}^{\mu}\otimes\mathds{1}\left(m_{\mu}\right), (38)

where GπμG_{\pi}^{\mu} is a unitary operator on the irreducible representation space associated with the Young diagram label μ\mu, mμm_{\mu} is the corresponding multiplicity, and 𝟙(mμ)\mathds{1}\left(m_{\mu}\right) is an mμ×mμm_{\mu}\times m_{\mu} identity matrix acting on the multiplicity subspace. By Schur’s lemma, for any group-invariant operator XX on 𝒲\mathcal{W}, i.e., XX commuting with all GπG_{\pi}, we have

UXU=|μ|=N𝟙(dμ)Xμ,U^{\dagger}XU=\bigoplus_{|\mu|=N}\mathds{1}\left(d_{\mu}\right)\otimes X^{\mu}, (39)

for any π\pi, where XμX^{\mu} is an mμ×mμm_{\mu}\times m_{\mu} matrix. With such block diagonalization of the permutation-invariant operator, we reduce the dimension from d2Nd^{2N} to [61, Eq. (57)]

|μ|=Nmμ2=(N+d21d21)(N+1)d21,\sum_{|\mu|=N}m_{\mu}^{2}=\binom{N+d^{2}-1}{d^{2}-1}\leq(N+1)^{d^{2}-1}, (40)

where the upper bound is obtained straightforwardly from the definition of the binomial coefficient. Specifically, if XX is further restricted to be a Hermitian matrix variable, then the number of real scalar variables contained in all elements of XX is reduced from d2Nd^{2N} to (N+d21d21)\binom{N+d^{2}-1}{d^{2}-1}.

Now let us turn to the optimization problem of QFI evaluation. If the Hermitian matrix hh is taken to be permutation invariant by Lemma 1, the number of variables concerned in hh is reduced from s2Ns^{2N} to (N+s21s21)\binom{N+s^{2}-1}{s^{2}-1}. Similarly, if further by Lemma 2 each Hermitian matrix QiQ^{i} is permutation invariant, the number of variables in QiQ^{i} is also reduced from d2Nd^{2N} to (N+d21d21)\binom{N+d^{2}-1}{d^{2}-1}, where we redefine d:=𝖽𝗂𝗆(1)𝖽𝗂𝗆(2)d:=\mathsf{dim}(\mathcal{H}_{1})\mathsf{dim}(\mathcal{H}_{2}). By this reduction the complexity gets polynomial rather than exponential, with respect to NN.

We can then reformulate the optimization in Theorem 1 with the reduced form. We consider two cases relevant to the strategy sets mentioned in the main text. First, under certain circumstances we can reduce the number of constraints in optimization as follows:

Theorem 2 (Symmetry reduced Theorem 1, first case).

If, for any πSN\pi\in S_{N} and any ii, there exists a jj such that the mapping SGπSGπS\mapsto G_{\pi}SG_{\pi}^{\dagger} on 𝖲i\mathsf{S}^{i} is a bijective function from 𝖲i\mathsf{S}^{i} to 𝖲j\mathsf{S}^{j}, and meanwhile for any i,ji,j there exists some permutation operation such that 𝖲i\mathsf{S}^{i} is bijectively mapped to 𝖲j\mathsf{S}^{j}, then the QFI of NN quantum channels ϕ\mathcal{E}_{\phi} can be expressed as:

J(𝖯)(Nϕ)=\displaystyle J^{(\mathsf{P})}(N_{\phi})= minλ,Q1,hμλ,\displaystyle\min_{\lambda,Q^{1},h^{\mu}}\lambda, (41)
s.t.\displaystyle\mathrm{s.t.} λQ1Ωϕ(h),Q1𝖲¯1,\displaystyle\,\lambda Q^{1}\geq\Omega_{\phi}(h),\ Q^{1}\in\overline{\mathsf{S}}^{1},

where h=U[|μ|=N𝟙(dμ)hμ]Uh=U^{\prime}\left[\bigoplus_{|\mu|=N}\mathds{1}\left(d_{\mu}\right)\otimes h^{\mu}\right]U^{\prime\dagger} with each hμh^{\mu} as an mμ×mμm_{\mu}^{\prime}\times m_{\mu}^{\prime} matrix variable and UU^{\prime} as a unitary transformation of basis.

Proof.

We first prove that, for any i,ji,j there exists some permutation operation QGπQGπQ\mapsto G_{\pi}QG_{\pi}^{\dagger} such that 𝖲¯i\overline{\mathsf{S}}^{i} is bijectively mapped to 𝖲¯j\overline{\mathsf{S}}^{j}. Given any ii and jj, if we choose an arbitrary Qi𝖲¯iQ^{i}\in\overline{\mathsf{S}}^{i}, i.e., Tr(QiSi)=1\Tr(Q^{i}S^{i})=1 for all Si𝖲iS^{i}\in\mathsf{S}^{i}, then the condition of the theorem implies that there exists GπG_{\pi} such that GπSiGπ𝖲jG_{\pi}S^{i}G_{\pi}^{\dagger}\in\mathsf{S}^{j}. As the map is bijective, in fact GπSjGπ𝖲iG_{\pi}^{\dagger}S^{j}G_{\pi}\in\mathsf{S}^{i} for all Sj𝖲jS^{j}\in\mathsf{S}^{j}. Then we have Tr(GπQiGπSj)=Tr(QiGπSjGπ)=1\Tr(G_{\pi}Q^{i}G_{\pi}^{\dagger}S^{j})=\Tr(Q^{i}G_{\pi}^{\dagger}S^{j}G_{\pi})=1 for all Sj𝖲jS^{j}\in\mathsf{S}^{j}. Therefore, following the same argument in the proof of Lemma 1, under QGπQGπQ\mapsto G_{\pi}QG_{\pi}^{\dagger}, 𝖲¯i\overline{\mathsf{S}}^{i} is bijectively mapped to 𝖲¯j\overline{\mathsf{S}}^{j}.

By Lemma 1 we can take hh to be permutation invariant and apply the block diagonalization to hh given by Eq. (39). Then for any Q1𝖲¯1Q^{1}\in\overline{\mathsf{S}}^{1} satisfying the constraint λQ1Ωϕ(h)\lambda Q^{1}\geq\Omega_{\phi}(h), there exists GπG_{\pi} such that Qi=GπQ1Gπ𝖲¯iQ^{i}=G_{\pi}Q^{1}G_{\pi}^{\dagger}\in\overline{\mathsf{S}}^{i} satisfying λQiGπΩϕ(h)Gπ=Ωϕ(h)\lambda Q^{i}\geq G_{\pi}\Omega_{\phi}(h)G_{\pi}^{\dagger}=\Omega_{\phi}(h) for any i=2,,Ki=2,\dots,K. Therefore, all the constraints except for λQ1Ωϕ(h)\lambda Q^{1}\geq\Omega_{\phi}(h) are redundant and can be removed. ∎

Now we consider the second case. If by Lemmas 1 and 2 hh and each QiQ^{i} are permutation invariant, we can then reformulate the constraints in optimization with reduced matrix dimensions. To relate the group representation on i(2i12i)\otimes_{i}\left(\mathcal{H}_{2i-1}\otimes\mathcal{H}_{2i}\right) to the representation on (Cs)N\mathbb{(}C^{s})^{\otimes N}, we choose a unitary transformation of basis UU^{\prime} for hh, decompose h=U[|μ|=N𝟙(dμ)hμ]Uh=U^{\prime}\left[\bigoplus_{|\mu|=N}\mathds{1}\left(d_{\mu}\right)\otimes h^{\mu}\right]U^{\prime\dagger} with hμh^{\mu} as an mμ×mμm_{\mu}^{\prime}\times m_{\mu}^{\prime} matrix, and divide U=(U,μ1,,U,μI)U^{\prime}=\left(U^{\prime,\mu^{1}},\dots,U^{\prime,\mu^{I}}\right) into blocks, where μi\mu^{i} is the label of the ii-th Young diagram, and U,μiU^{\prime,\mu^{i}} has dμimμid_{\mu^{i}}m_{\mu^{i}}^{\prime} columns. Thus by defining the matrix

𝐍~˙ϕμ=𝐍˙ϕU,μi𝐍ϕU,μ[𝟙(dμ)hμ],\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu}=\dot{\mathbf{N}}_{\phi}U^{\prime,\mu}-\mathrm{i}\mathbf{N}_{\phi}U^{\prime,\mu}\left[\mathds{1}\left(d_{\mu}\right)\otimes h^{\mu}\right], (42)

we have the permutation-invariant performance operator

Ωϕ(h)=4|μ|=N(𝐍~˙ϕμ𝐍~˙ϕμ)T=4|μ|=N𝐍~˙ϕμ𝐍~˙ϕμ.\Omega_{\phi}(h)=4\sum_{|\mu|=N}\left(\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu}\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu\dagger}\right)^{T}=4\sum_{|\mu|=N}\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu*}\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu*\dagger}. (43)

Analogously, we apply the block diagonalization using a change of basis UU to

Qi=U[|μ|=N𝟙(dμ)Qi,μ]U𝖲¯i,Q^{i}=U\left[\bigoplus_{|\mu|=N}\mathds{1}\left(d_{\mu}\right)\otimes Q^{i,\mu}\right]U^{\dagger}\in\overline{\mathsf{S}}^{i}, (44)

where Qi,μQ^{i,\mu} is an mμ×mμm_{\mu}\times m_{\mu} matrix variable. The unitary transformation U=(Uμ1,,UμI)U=\left(U^{\mu^{1}},\dots,U^{\mu^{I}}\right) is first divided into blocks, and then each Uμi=(Uμi,1,,Uμi,dμi)U^{\mu^{i}}=\left(U^{\mu^{i},1},\dots,U^{\mu^{i},d_{\mu^{i}}}\right) is divided into blocks for i=1,,Ii=1,\dots,I, where μi\mu^{i} is the label of the ii-th Young diagram, and Uμi,jU^{\mu^{i},j} has mμim_{\mu^{i}} columns for j=1,,dμij=1,\dots,d_{\mu^{i}}. Note that UU also gives the block diagonalization of Ωϕ(h)\Omega_{\phi}(h).

Before proceeding further we prove that the range of 𝐍~˙ϕμ\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu*} is exactly contained in the irreducible representation space corresponding to μ\mu:

Lemma 3.

If we decompose the representation space 𝒲=i=1N(2i12i)=|μ|=N𝒲μ\mathcal{W}=\otimes_{i=1}^{N}\left(\mathcal{H}_{2i-1}\otimes\mathcal{H}_{2i}\right)=\bigoplus_{|\mu|=N}\mathcal{W}^{\mu}, then 𝖱𝖺𝗇𝗀𝖾(𝐍~˙ϕμ)𝒲μ\mathsf{Range}\left(\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu*}\right)\subseteq\mathcal{W}^{\mu} for any Young diagram label μ\mu.

Proof.

To characterize the representation space 𝒲μ\mathcal{W}^{\mu} we introduce the key notion of Young symmetrizer as well as other related concepts very briefly (see, e.g., [60] for mathematical details). A Young tableau is a filling into the boxes of the Young diagram with positive integers weakly increasing along each row and strictly increasing along each column. For a standard Young tableau labeled by ν\nu, i.e., a Young tableau filled with the entries 1,,N1,\dots,N, we define two permutation subgroups

Pν:={σSN|σpreserveseachrow}P_{\nu}:=\left\{\sigma\in S_{N}\middle|\sigma\ \mathrm{preserves}\ \mathrm{each}\ \mathrm{row}\right\} (45)

and

Qν:={σSN|σpreserveseachcolumn}.Q_{\nu}:=\left\{\sigma\in S_{N}\middle|\sigma\ \mathrm{preserves}\ \mathrm{each}\ \mathrm{column}\right\}. (46)

In the group algebra SN\mathbb{C}S_{N} we define two elements aν:=σPνeσa_{\nu}:=\sum_{\sigma\in P_{\nu}}e_{\sigma} and bν:=σQνsgn(σ)eσb_{\nu}:=\sum_{\sigma\in Q_{\nu}}\mathrm{sgn}(\sigma)e_{\sigma}, where eσe_{\sigma} is the unit vector corresponding to σ\sigma and sgn()\mathrm{sgn}(\cdot) denotes the parity of the permutation. Then the Young symmetrizer is defined by

cν:=aνbν=σPνπQνsgn(π)eσπSN.c_{\nu}:=a_{\nu}b_{\nu}=\sum_{\sigma\in P_{\nu}}\sum_{\pi\in Q_{\nu}}\mathrm{sgn}(\pi)e_{\sigma\pi}\in\mathbb{C}S_{N}. (47)

It is known that a Young diagram of μ\mu corresponds to dμd_{\mu} standard Young tableaux, with each Young tableau of ν\nu characterizing an irreducible representation space, given by the image of cνc_{\nu} on i𝒲i\otimes_{i}\mathcal{W}_{i} under the natural group algebra representation SNEnd(i𝒲i)\mathbb{C}S_{N}\rightarrow\mathrm{End}\left(\otimes_{i}\mathcal{W}_{i}\right), where End(V)\mathrm{End}(V) denotes the set of endomorphisms on VV.

With these notions we now have an explicit characterization of the representation space. Note that we can prove 𝖱𝖺𝗇𝗀𝖾(𝐍~˙ϕμ)𝒲μ\mathsf{Range}\left(\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu*}\right)\subseteq\mathcal{W}^{\mu} if 𝖱𝖺𝗇𝗀𝖾(𝐍˙ϕU,μ)𝒲μ\mathsf{Range}\left(\dot{\mathbf{N}}_{\phi}^{*}U^{\prime,\mu*}\right)\subseteq\mathcal{W}^{\mu} and 𝖱𝖺𝗇𝗀𝖾(𝐍ϕU,μ)𝒲μ\mathsf{Range}\left(\mathbf{N}_{\phi}^{*}U^{\prime,\mu*}\right)\subseteq\mathcal{W}^{\mu}. In the proof of Lemma 1 we have seen Gπ|N˙ϕ,𝐢=|N˙ϕ,π(𝐢)G_{\pi}\ket{\dot{N}_{\phi,\mathbf{i}}}=\ket{\dot{N}_{\phi,\pi(\mathbf{i})}} and Gπ|Nϕ,𝐢=|Nϕ,π(𝐢)G_{\pi}\ket{N_{\phi,\mathbf{i}}}=\ket{N_{\phi,\pi(\mathbf{i})}}, from which it follows that

Gπ𝐍˙ϕ=𝐍˙ϕGπG_{\pi}\dot{\mathbf{N}}_{\phi}=\dot{\mathbf{N}}_{\phi}G_{\pi}^{\prime} (48)

and Gπ𝐍ϕ=𝐍ϕGπG_{\pi}\mathbf{N}_{\phi}=\mathbf{N}_{\phi}G_{\pi}^{\prime}.

Now we prove that 𝖱𝖺𝗇𝗀𝖾(𝐍˙ϕU,μ)𝒲μ\mathsf{Range}\left(\dot{\mathbf{N}}_{\phi}^{*}U^{\prime,\mu*}\right)\subseteq\mathcal{W}^{\mu}. From the discussions above we know that

𝖱𝖺𝗇𝗀𝖾(U,μ)=νcν[is],\mathsf{Range}\left(U^{\prime,\mu}\right)=\bigoplus_{\nu}c_{\nu}\left[\otimes_{i}\mathbb{C}^{s}\right], (49)

for all Young tableau labels ν\nu corresponding to the Young diagram of μ\mu. Explicitly, we have

cν[is]=𝖱𝖺𝗇𝗀𝖾[σPνπQνsgn(π)Gσπ].c_{\nu}\left[\otimes_{i}\mathbb{C}^{s}\right]=\mathsf{Range}\left[\sum_{\sigma\in P_{\nu}}\sum_{\pi\in Q_{\nu}}\mathrm{sgn}(\pi)G_{\sigma\pi}^{\prime}\right]. (50)

Note that there always exists a unitary transformation of basis VV such that we can obtain a real matrix U(real),μ=U,μVU_{(\mathrm{real})}^{\prime,\mu}=U^{\prime,\mu}V^{\dagger}, which leads to

𝖱𝖺𝗇𝗀𝖾(U,μ)=𝖱𝖺𝗇𝗀𝖾(U(real),μV)=𝖱𝖺𝗇𝗀𝖾(U(real),μV)=𝖱𝖺𝗇𝗀𝖾(U,μ).\mathsf{Range}\left(U^{\prime,\mu*}\right)=\mathsf{Range}\left(U_{(\mathrm{real})}^{\prime,\mu*}V^{*}\right)=\mathsf{Range}\left(U_{(\mathrm{real})}^{\prime,\mu}V^{*}\right)=\mathsf{Range}\left(U^{\prime,\mu}\right). (51)

Thus 𝖱𝖺𝗇𝗀𝖾(𝐍˙ϕU,μ)=𝖱𝖺𝗇𝗀𝖾(𝐍˙ϕU,μ)\mathsf{Range}\left(\dot{\mathbf{N}}_{\phi}^{*}U^{\prime,\mu*}\right)=\mathsf{Range}\left(\dot{\mathbf{N}}_{\phi}^{*}U^{\prime,\mu}\right). Furthermore, from Eqs. (49) and (50) we have 𝖱𝖺𝗇𝗀𝖾(𝐍˙ϕU,μ)𝒲μ\mathsf{Range}\left(\dot{\mathbf{N}}_{\phi}^{*}U^{\prime,\mu*}\right)\subseteq\mathcal{W}^{\mu} if

𝖱𝖺𝗇𝗀𝖾[𝐍˙ϕσPνπQνsgn(π)Gσπ]𝒲μ\mathsf{Range}\left[\dot{\mathbf{N}}_{\phi}^{*}\sum_{\sigma\in P_{\nu}}\sum_{\pi\in Q_{\nu}}\mathrm{sgn}(\pi)G_{\sigma\pi}^{\prime}\right]\subseteq\mathcal{W}^{\mu} (52)

for all Young tableau labels ν\nu corresponding to the Young diagram of μ\mu. To show Eq. (52), note that by Eq. (48) we have

𝖱𝖺𝗇𝗀𝖾[𝐍˙ϕσPνπQνsgn(π)Gσπ]=𝖱𝖺𝗇𝗀𝖾[σPνπQνsgn(π)Gσπ𝐍˙ϕ].\mathsf{Range}\left[\dot{\mathbf{N}}_{\phi}^{*}\sum_{\sigma\in P_{\nu}}\sum_{\pi\in Q_{\nu}}\mathrm{sgn}(\pi)G_{\sigma\pi}^{\prime}\right]=\mathsf{Range}\left[\sum_{\sigma\in P_{\nu}}\sum_{\pi\in Q_{\nu}}\mathrm{sgn}(\pi)G_{\sigma\pi}\dot{\mathbf{N}}_{\phi}^{*}\right]. (53)

Since the R.H.S. of Eq. (53) is the image of the Young symmetrizer cνc_{\nu} on 𝖱𝖺𝗇𝗀𝖾(𝐍˙ϕ)\mathsf{Range}\left(\dot{\mathbf{N}}_{\phi}^{*}\right), it is a subset of 𝒲μ\mathcal{W}^{\mu}. Therefore, 𝖱𝖺𝗇𝗀𝖾(𝐍˙ϕU,μ)𝒲μ\mathsf{Range}\left(\dot{\mathbf{N}}_{\phi}^{*}U^{\prime,\mu*}\right)\subseteq\mathcal{W}^{\mu}. In the same way we can also show that 𝖱𝖺𝗇𝗀𝖾(𝐍ϕU,μ)𝒲μ\mathsf{Range}\left(\mathbf{N}_{\phi}^{*}U^{\prime,\mu*}\right)\subseteq\mathcal{W}^{\mu} and thus complete the proof. ∎

Therefore, UΩϕ(h)UU^{\dagger}\Omega_{\phi}(h)U is block diagonal with each block on the space 𝖱𝖺𝗇𝗀𝖾(𝐍~˙ϕμ)\mathsf{Range}\left(\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu}\right). This results in:

Theorem 3 (Symmetry reduced Theorem 1, second case).

If each affine space 𝖲i\mathsf{S}^{i} is permutation invariant for any i=1,,Ki=1,\dots,K, then the QFI of NN quantum channels ϕ\mathcal{E}_{\phi} can be expressed as:

J(𝖯)(Nϕ)=\displaystyle J^{(\mathsf{P})}(N_{\phi})= minλ,Qi,μ,hμλ,\displaystyle\min_{\lambda,Q^{i,\mu},h^{\mu}}\lambda, (54)
s.t.\displaystyle\mathrm{s.t.} λQi,μ4Uμ,1(𝐍~˙ϕμ𝐍~˙ϕμ)TUμ,1,i=1,,K,|μ|=N,\displaystyle\,\lambda Q^{i,\mu}\geq 4U^{\mu,1\dagger}\left(\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu}\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu\dagger}\right)^{T}U^{\mu,1},\ i=1,\dots,K,\ |\mu|=N,

where Qi,μQ^{i,\mu} is given by Eq. (44).

Proof.

By Lemmas 1 and 2, hh, Ωϕ(h)\Omega_{\phi}(h) and all QiQ^{i} are taken to be permutation invariant. The original constraint λQiΩϕ(h)\lambda Q^{i}\geq\Omega_{\phi}(h) on the permutation-invariant space is reformulated as

λ|μ|=N𝟙(dμ)Qi,μ|μ|=N4Uμ|ν|=N𝐍~˙ϕν𝐍~˙ϕνUμ.\lambda\bigoplus_{|\mu|=N}\mathds{1}\left(d_{\mu}\right)\otimes Q^{i,\mu}\geq\bigoplus_{|\mu|=N}4U^{\mu\dagger}\sum_{|\nu|=N}\dot{\tilde{\mathbf{N}}}_{\phi}^{\nu*}\dot{\tilde{\mathbf{N}}}_{\phi}^{\nu*\dagger}U^{\mu}. (55)

Since by Lemma 3 𝖱𝖺𝗇𝗀𝖾(𝐍~˙ϕμ)𝒲μ=𝖱𝖺𝗇𝗀𝖾(Uμ)\mathsf{Range}\left(\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu*}\right)\subseteq\mathcal{W}^{\mu}=\mathsf{Range}\left(U^{\mu}\right) for any μ\mu, we have Uμ|ν|=N𝐍~˙ϕν𝐍~˙ϕνUμ=Uμ𝐍~˙ϕμ𝐍~˙ϕμUμU^{\mu\dagger}\sum_{|\nu|=N}\dot{\tilde{\mathbf{N}}}_{\phi}^{\nu*}\dot{\tilde{\mathbf{N}}}_{\phi}^{\nu*\dagger}U^{\mu}=U^{\mu\dagger}\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu*}\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu*\dagger}U^{\mu}. Then Eq. (55) can be reformulated as

λ𝟙(dμ)Qi,μ4Uμ𝐍~˙ϕμ𝐍~˙ϕμUμ,|μ|=N.\lambda\mathds{1}\left(d_{\mu}\right)\otimes Q^{i,\mu}\geq 4U^{\mu\dagger}\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu*}\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu*\dagger}U^{\mu},\ |\mu|=N. (56)

Furthermore, both sides of the inequality in Eq. (56) are block diagonal with the same repeating blocks and we only need to compare one of the blocks. Without loss of generality, we choose the first block for comparison and obtain

λQi,μ4Uμ,1(𝐍~˙ϕμ𝐍~˙ϕμ)TUμ,1,\lambda Q^{i,\mu}\geq 4U^{\mu,1\dagger}\left(\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu}\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu\dagger}\right)^{T}U^{\mu,1}, (57)

which holds for all i=1,,Ki=1,\dots,K and |μ|=N|\mu|=N. ∎

We have not characterized each dual affine space 𝖲¯i\overline{\mathsf{S}}^{i} for i=1,,Ki=1,\dots,K yet. If we choose an affine basis {Si,j}j=1Mi\{S^{i,j}\}_{j=1}^{M_{i}} for each 𝖲i\mathsf{S}^{i}, then the constraint Qi𝖲¯iQ^{i}\in\overline{\mathsf{S}}^{i} can be reformulated as a set of linear constraints Tr(QiSi,j)=1\Tr(Q^{i}S^{i,j})=1 for all j=1,,Mij=1,\dots,M_{i}. If each affine space 𝖲i\mathsf{S}^{i} is permutation invariant, then by the proof of Lemma 2 we have, for a feasible solution of Qi𝖲¯iQ^{i}\in\overline{\mathsf{S}}^{i}, GπQiGπG_{\pi}Q^{i}G_{\pi}^{\dagger} is also a feasible solution for any πSN\pi\in S_{N}. Then by defining

S~i,j:=1N!πSNGπSi,jGπ,\tilde{S}^{i,j}:=\frac{1}{N!}\sum_{\pi\in S_{N}}G_{\pi}S^{i,j}G_{\pi}^{\dagger}, (58)

each linear constraint Tr(QiSi,j)=1\Tr(Q^{i}S^{i,j})=1 can be replaced by Tr(QiS~i,j)=1\Tr(Q^{i}\tilde{S}^{i,j})=1 without changing the problem. Since S~i,j\tilde{S}^{i,j} is permutation invariant, similar to Eq. (44) it can be decomposed as

S~i,j=U[|μ|=N𝟙(dμ)S~i,j,μ]U,\tilde{S}^{i,j}=U\left[\bigoplus_{|\mu|=N}\mathds{1}\left(d_{\mu}\right)\otimes\tilde{S}^{i,j,\mu}\right]U^{\dagger}, (59)

where S~i,j,μ\tilde{S}^{i,j,\mu} is an mμ×mμm_{\mu}\times m_{\mu} matrix. Combining Eqs. (44) and (59), we can reformulate the constraint Qi𝖲¯iQ^{i}\in\overline{\mathsf{S}}^{i} with reduced matrix sizes as

|μ|=NdμTr(Qi,μS~i,j,μ)=1,j=1,,Mi.\sum_{|\mu|=N}d_{\mu}\Tr(Q^{i,\mu}\tilde{S}^{i,j,\mu})=1,\ j=1,\dots,M_{i}. (60)

Finally, it remains to be seen how to find the unitary transformation UU and UU^{\prime} for block diagonalization. Known rigorous numerical algorithms for identifying the transformation are fairly expensive [62]. Fortunately, RepLAB [63], a numerical approach to decomposing representations based on randomized heuristics works very well in practice and is thus adopted here.

D.2 Symmetry reduced algorithm for optimal strategies

The idea is similar to the symmetry reduced evaluation of QFI. Since the first step of Algorithm 1 is simply solving the optimization problem in Theorem 1, now we only consider its second step, where we need to solve for the optimal value of P~\tilde{P} in

maxP~𝖯~Tr[P~Ωϕ(h(opt))],\max_{\tilde{P}\in\tilde{\mathsf{P}}}\Tr\left[\tilde{P}\Omega_{\phi}\left(h^{(\mathrm{opt})}\right)\right], (61)

where 𝖯~=𝖢𝗈𝗇𝗏{i=1K{Si0|Si𝖲i}}\tilde{\mathsf{P}}=\mathsf{Conv}\left\{\bigcup_{i=1}^{K}\left\{S^{i}\geq 0\middle|S^{i}\in\mathsf{S}^{i}\right\}\right\} and

Re{Tr{P~[i𝐍ϕ(𝐍˙ϕi𝐍ϕh(opt))]T}}=0forallr.\real\left\{\Tr\left\{\tilde{P}\left[-\mathrm{i}\mathbf{N}_{\phi}\mathscr{H}\left(\dot{\mathbf{N}}_{\phi}-\mathrm{i}\mathbf{N}_{\phi}h^{(\mathrm{opt})}\right)^{\dagger}\right]^{T}\right\}\right\}=0\ \mathrm{for}\ \mathrm{all}\ \mathscr{H}\in\mathbb{H}_{r}. (62)

We have the following:

Lemma 4.

If, for any πSN\pi\in S_{N} and any ii, there exists a jj such that the mapping SGπSGπS\mapsto G_{\pi}SG_{\pi}^{\dagger} on 𝖲i\mathsf{S}^{i} is a bijective function from 𝖲i\mathsf{S}^{i} to 𝖲j\mathsf{S}^{j}, then there must exist a permutation-invariant P~\tilde{P} as an optimal solution in Algorithm 1.

Proof.

By definition 𝖯~\tilde{\mathsf{P}} is permutation invariant. By Lemma 1 we can choose a permutation-invariant h(opt)h^{(\mathrm{opt})} and thus Ωϕ(h(opt))\Omega_{\phi}\left(h^{(\mathrm{opt})}\right) is also permutation invariant. If P~(opt)\tilde{P}^{(\mathrm{opt})} is an optimal solution of P~\tilde{P}, then GπP~(opt)GπG_{\pi}\tilde{P}^{(\mathrm{opt})}G_{\pi}^{\dagger} for any πSN\pi\in S_{N} is also an optimal solution, since

Tr[GπP~(opt)GπΩϕ(h(opt))]=Tr[P~(opt)GπΩϕ(h(opt))Gπ]=Tr[P~(opt)Ωϕ(h(opt))]\Tr\left[G_{\pi}\tilde{P}^{(\mathrm{opt})}G_{\pi}^{\dagger}\Omega_{\phi}\left(h^{(\mathrm{opt})}\right)\right]=\Tr\left[\tilde{P}^{(\mathrm{opt})}G_{\pi}^{\dagger}\Omega_{\phi}\left(h^{(\mathrm{opt})}\right)G_{\pi}\right]=\Tr\left[\tilde{P}^{(\mathrm{opt})}\Omega_{\phi}\left(h^{(\mathrm{opt})}\right)\right] (63)

and for all r\mathscr{H}\in\mathbb{H}_{r}

Re{Tr{GπP~(opt)Gπ[i𝐍ϕGπGπ(𝐍˙ϕi𝐍ϕh(opt))]T}}\displaystyle\real\left\{\Tr\left\{G_{\pi}\tilde{P}^{(\mathrm{opt})}G_{\pi}^{\dagger}\left[-\mathrm{i}\mathbf{N}_{\phi}G_{\pi}^{\prime}\mathscr{H}G_{\pi}^{\prime\dagger}\left(\dot{\mathbf{N}}_{\phi}-\mathrm{i}\mathbf{N}_{\phi}h^{(\mathrm{opt})}\right)^{\dagger}\right]^{T}\right\}\right\} (64)
=\displaystyle= Re{Tr{P~(opt)Gπ[i𝐍ϕGπGπ(𝐍˙ϕi𝐍ϕh(opt))]TGπ}}\displaystyle\real\left\{\Tr\left\{\tilde{P}^{(\mathrm{opt})}G_{\pi}^{\dagger}\left[-\mathrm{i}\mathbf{N}_{\phi}G_{\pi}^{\prime}\mathscr{H}G_{\pi}^{\prime\dagger}\left(\dot{\mathbf{N}}_{\phi}-\mathrm{i}\mathbf{N}_{\phi}h^{(\mathrm{opt})}\right)^{\dagger}\right]^{T}G_{\pi}\right\}\right\}
=\displaystyle= Re{Tr{P~(opt)[i𝐍ϕ(𝐍˙ϕi𝐍ϕh(opt))]T}}\displaystyle\real\left\{\Tr\left\{\tilde{P}^{(\mathrm{opt})}\left[-\mathrm{i}\mathbf{N}_{\phi}\mathscr{H}\left(\dot{\mathbf{N}}_{\phi}-\mathrm{i}\mathbf{N}_{\phi}h^{(\mathrm{opt})}\right)^{\dagger}\right]^{T}\right\}\right\}
=\displaystyle=  0,\displaystyle\,0,

having used Gπ𝐍ϕ=𝐍ϕGπG_{\pi}^{\dagger}\mathbf{N}_{\phi}=\mathbf{N}_{\phi}G_{\pi}^{\prime\dagger} and Gπ𝐍˙ϕ=𝐍˙ϕGπG_{\pi}^{\dagger}\dot{\mathbf{N}}_{\phi}=\dot{\mathbf{N}}_{\phi}G_{\pi}^{\prime\dagger} for any πSN\pi\in S_{N} in the second equality of Eq. (64). Therefore, there exists a permutation-invariant solution P~=1N!πSNGπP~(opt)Gπ\tilde{P}=\frac{1}{N!}\sum_{\pi\in S_{N}}G_{\pi}\tilde{P}^{(\mathrm{opt})}G_{\pi}^{\dagger}. ∎

Now by following the same line of arguments as used from Eqs. (58) to (60), we define

O:=\displaystyle O:= 1N!πSNGπ[i𝐍ϕ(𝐍˙ϕi𝐍ϕh(opt))]TGπ\displaystyle\,\frac{1}{N!}\sum_{\pi\in S_{N}}G_{\pi}\left[-\mathrm{i}\mathbf{N}_{\phi}\mathscr{H}\left(\dot{\mathbf{N}}_{\phi}-\mathrm{i}\mathbf{N}_{\phi}h^{(\mathrm{opt})}\right)^{\dagger}\right]^{T}G_{\pi}^{\dagger} (65)
=\displaystyle= [i𝐍ϕ~(𝐍˙ϕi𝐍ϕh(opt))]T,\displaystyle\,\left[-\mathrm{i}\mathbf{N}_{\phi}\tilde{\mathscr{H}}\left(\dot{\mathbf{N}}_{\phi}-\mathrm{i}\mathbf{N}_{\phi}h^{(\mathrm{opt})}\right)^{\dagger}\right]^{T},

having defined the permutation-invariant ~:=1N!πSNGπGπ\tilde{\mathscr{H}}:=\frac{1}{N!}\sum_{\pi\in S_{N}}G_{\pi}^{\prime}\mathscr{H}G_{\pi}^{\prime\dagger}. Similar to the arguments in Appendix C, by choosing a basis {~i}i=1J\left\{\tilde{\mathscr{H}}^{i}\right\}_{i=1}^{J} for the permutation-invariant subspace of r×r\mathbb{H}_{r\times r}, where J=(N+s21s21)J=\binom{N+s^{2}-1}{s^{2}-1}, Eq. (31) can be reformulated as a set of JJ linear constraints. We further define

Oi:=[i𝐍ϕ~i(𝐍˙ϕi𝐍ϕh(opt))]TO^{i}:=\left[-\mathrm{i}\mathbf{N}_{\phi}\tilde{\mathscr{H}}^{i}\left(\dot{\mathbf{N}}_{\phi}-\mathrm{i}\mathbf{N}_{\phi}h^{(\mathrm{opt})}\right)^{\dagger}\right]^{T} (66)

for i=1,,Ji=1,\dots,J. Now we can decompose Oi=U[|μ|=N𝟙(dμ)O~i,μ]UO^{i}=U\left[\bigoplus_{|\mu|=N}\mathds{1}\left(d_{\mu}\right)\otimes\tilde{O}^{i,\mu}\right]U^{\dagger}, P~=U[|μ|=N𝟙(dμ)P~μ]U\tilde{P}=U\left[\bigoplus_{|\mu|=N}\mathds{1}\left(d_{\mu}\right)\otimes\tilde{P}^{\mu}\right]U^{\dagger}, Ωϕ(h(opt))=U[|μ|=N𝟙(dμ)Ωϕμ(h(opt))]U\Omega_{\phi}\left(h^{(\mathrm{opt})}\right)=U\left[\bigoplus_{|\mu|=N}\mathds{1}\left(d_{\mu}\right)\otimes\Omega_{\phi}^{\mu}\left(h^{(\mathrm{opt})}\right)\right]U^{\dagger}, and then reformulate the optimization problem as the following reduced form:

maxP~μ\displaystyle\max_{\tilde{P}^{\mu}} |μ|=NdμTr[P~μΩϕμ(h(opt))],\displaystyle\sum_{|\mu|=N}d_{\mu}\Tr\left[\tilde{P}^{\mu}\Omega_{\phi}^{\mu}\left(h^{(\mathrm{opt})}\right)\right], (67)
s.t.\displaystyle\mathrm{s.t.} |μ|=NdμRe[Tr(P~μOi,μ)]=0foralli=1,,J.\displaystyle\sum_{|\mu|=N}d_{\mu}\real\left[\Tr\left(\tilde{P}^{\mu}O^{i,\mu}\right)\right]=0\ \mathrm{for}\ \mathrm{all}\ i=1,\dots,J.

Recall that P~=i=1KqiSi\tilde{P}=\sum_{i=1}^{K}q^{i}S^{i} for Si𝖲iS^{i}\in\mathsf{S}^{i}. By choosing an affine basis {Qi,j}j=1Li\{Q^{i,j}\}_{j=1}^{L_{i}} for 𝖲¯i\overline{\mathsf{S}}^{i} we can also characterize Si𝖲iS^{i}\in\mathsf{S}^{i} by a set of linear constraints Tr(SiQi,j)=1\Tr\left(S^{i}Q^{i,j}\right)=1 for i=1,,Ki=1,\dots,K and j=1,,Lij=1,\dots,L_{i}, and follow the same routine to tackle the constraints on the permutation-invariant subspace. Thus both the number of variables and the number of constraints are polynomial with respect to NN.

Appendix E Evaluation of QFI using different strategies

In this section we provide explicit formulas of the QFI for all strategy sets considered in the main text, in the forms which can be numerically solved by SDP. Without the positivity constraints, parallel, sequential and general indefinite-causal-order strategy sets are affine spaces themselves, while quantum SWITCH and causal superposition strategy sets are convex hulls of affine spaces. In some cases the result of Theorem 1 can be simplified a bit, as it is possible to trace over certain subspace while formulating the primal problem at the beginning.

E.1 Parallel strategies

When definite causal order is obeyed, a strategy can be described by a quantum comb [33, 64, 34]. The dual affine space is the set of dual combs without the positivity constraint [43]. For parallel strategies the primal problem can be written as

J(𝖯𝖺𝗋)(Nϕ)=\displaystyle J^{(\mathsf{Par})}(N_{\phi})= minhrmaxP~Tr[P~Ωϕ(h)],\displaystyle\min_{h\in\mathbb{H}_{r}}\max_{\tilde{P}}\Tr\left[\tilde{P}\Omega_{\phi}(h)\right], (68)
s.t.\displaystyle\mathrm{s.t.} P~0,\displaystyle\,\tilde{P}\geq 0,
P~=𝟙2,4,,2NP~(1),\displaystyle\,\tilde{P}=\mathds{1}_{2,4,\dots,2N}\otimes\tilde{P}^{(1)},
TrP~(1)=1.\displaystyle\Tr\tilde{P}^{(1)}=1.

Equivalently, the problem can be formulated as

minhrmaxP\displaystyle\min_{h\in\mathbb{H}_{r}}\max_{P} Tr[PTr2,4,,2NΩϕ(h)],\displaystyle\Tr\left[P\Tr_{2,4,\dots,2N}\Omega_{\phi}(h)\right], (69)
s.t.\displaystyle\mathrm{s.t.} P0,\displaystyle\,P\geq 0,
TrP=1.\displaystyle\Tr P=1.

The dual problem is given by

minλ,h\displaystyle\min_{\lambda,h} λ,\displaystyle\,\lambda, (70)
s.t.\displaystyle\mathrm{s.t.} λ𝟙1,3,,2N1Tr2,4,,2NΩϕ(h),\displaystyle\,\lambda\mathds{1}_{1,3,\dots,2N-1}\geq\Tr_{2,4,\dots,2N}\Omega_{\phi}(h),

which simplifies the result directly obtained from Theorem 1 a bit. To solve the problem via SDP, we define a block matrix

A:=(λ4𝟙(ri=1Nd2i)n1,1|nr,i=1Nd2i||n1,1|nr,i=1Nd2i𝟙1,3,,2N1),A:=\left(\begin{array}[]{c | c}\frac{\lambda}{4}\mathds{1}\left(r\prod_{i=1}^{N}d_{2i}\right)&\begin{array}[]{c}\bra{n_{1,1}}\\ \vdots\\ \bra{n_{r,\prod_{i=1}^{N}d_{2i}}}\end{array}\\ \hline\cr\begin{array}[]{ccc}\ket{n_{1,1}}&\ldots&\ket{n_{r,\prod_{i=1}^{N}d_{2i}}}\end{array}&\mathds{1}_{1,3,\dots,2N-1}\end{array}\right), (71)

wherein 𝟙(d)\mathds{1}(d) denotes a dd-dimensional identity matrix, and

|ni,j:=j|N~˙ϕ,i,\ket{n_{i,j}}:=\innerproduct{j}{\dot{\tilde{N}}_{\phi,i}^{*}}, (72)

where |N~˙ϕ,j=|N˙ϕ,jik|Nϕ,khkj\ket{\dot{\tilde{N}}_{\phi,j}}=\ket{\dot{N}_{\phi,j}}-\mathrm{i}\sum_{k}\ket{N_{\phi,k}}h_{kj} and {|j,j=1,,k=1Nd2k}\left\{\ket{j},\ j=1,\dots,\prod_{k=1}^{N}d_{2k}\right\} forms an orthonormal basis of k=1N2k\otimes_{k=1}^{N}\mathcal{H}_{2k}, having assumed that the identity map trivially acts on the subspace where the dual vector j|\bra{j} does not affect. Note that

i,j|ni,jni,j|=14Tr2,4,,2NΩϕ(h).\sum_{i,j}\outerproduct{n_{i,j}}{n_{i,j}}=\frac{1}{4}\Tr_{2,4,\dots,2N}\Omega_{\phi}(h). (73)

By Schur complement lemma [65, Theorem 1.12], the constraint in Eq. (70) is equivalent to the requirement that A0A\geq 0. Hence, the QFI for parallel strategies is solved by

minλ,h\displaystyle\min_{\lambda,h} λ,\displaystyle\,\lambda, (74)
s.t.\displaystyle\mathrm{s.t.} A0,\displaystyle\,A\geq 0,
hr.\displaystyle\,h\in\mathbb{H}_{r}.

The problem can be solved by SDP since hh is incorporated linearly in the blocks of AA.

Symmetry reduction.—We can reduce the problem using permutation symmetry. For 𝖯𝖺𝗋\mathsf{Par}, the set 𝖯~\tilde{\mathsf{P}} is given by 𝖯~={S0|S𝖲}\tilde{\mathsf{P}}=\left\{S\geq 0\middle|S\in\mathsf{S}\right\}, and the affine space 𝖲\mathsf{S} is permutation invariant. As explained in Appendix D we can decompose the permutation-invariant h=U[|μ|=N𝟙(dμ)hμ]Uh=U^{\prime}\left[\bigoplus_{|\mu|=N}\mathds{1}\left(d_{\mu}\right)\otimes h^{\mu}\right]U^{\prime\dagger} with hμh^{\mu} as an mμ×mμm_{\mu}^{\prime}\times m_{\mu}^{\prime} matrix. Q𝖲¯Q\in\overline{\mathsf{S}} is characterized by the constraint Tr2,4,,2NQ=𝟙1,3,,2N1\Tr_{2,4,\dots,2N}Q=\mathds{1}_{1,3,\dots,2N-1}, and we can decompose Q=U[|μ|=N𝟙(dμ)Qμ]UQ=U\left[\bigoplus_{|\mu|=N}\mathds{1}\left(d_{\mu}\right)\otimes Q^{\mu}\right]U^{\dagger} with QμQ^{\mu} as an mμ×mμm_{\mu}\times m_{\mu} matrix. If we define

Aμ:=(λ4𝟙(dμmμ)(Uμ,1𝐍~˙ϕμ)Uμ,1𝐍~˙ϕμQμ),A^{\mu}:=\left(\begin{array}[]{c | c}\frac{\lambda}{4}\mathds{1}\left(d_{\mu}m_{\mu}^{\prime}\right)&\left(U^{\mu,1\dagger}\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu*}\right)^{\dagger}\\ \hline\cr U^{\mu,1\dagger}\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu*}&Q^{\mu}\end{array}\right), (75)

where 𝐍~˙ϕμ\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu} is given by Eq. (42), then by Theorem 3 we can reformulate the optimization problem as

minλ,Qμ,h\displaystyle\min_{\lambda,Q^{\mu},h} λ,\displaystyle\,\lambda, (76)
s.t.\displaystyle\mathrm{s.t.} Aμ0,hμmμ,|μ|=N,\displaystyle\,A^{\mu}\geq 0,\ h^{\mu}\in\mathbb{H}_{m_{\mu}^{\prime}},\ |\mu|=N,
Tr2,4,,2NQ=𝟙1,3,,2N1.\displaystyle\Tr_{2,4,\dots,2N}Q=\mathds{1}_{1,3,\dots,2N-1}.

The constraint Tr2,4,,2NQ=𝟙1,3,,2N1\Tr_{2,4,\dots,2N}Q=\mathds{1}_{1,3,\dots,2N-1} only requires to be explicitly characterized on the permutation-invariant subspace, since Tr2,4,,2NQ\Tr_{2,4,\dots,2N}Q is permutation invariant. Therefore, not only the number of scalar variables but also the number of constraints in terms of scalar variables are polynomial with respect to NN.

E.2 Sequential strategies

For sequential strategies the problem can be written as (having traced over 2N\mathcal{H}_{2N})

J(𝖲𝖾𝗊)(Nϕ)=\displaystyle J^{(\mathsf{Seq})}(N_{\phi})= minhrmaxP(k)Tr[P(N)Tr2NΩϕ(h)],\displaystyle\min_{h\in\mathbb{H}_{r}}\max_{P^{(k)}}\Tr\left[P^{(N)}\Tr_{2N}\Omega_{\phi}(h)\right], (77)
s.t.\displaystyle\mathrm{s.t.} P(N)0,\displaystyle\,P^{(N)}\geq 0,
Tr2k1P(k)=𝟙2k2P(k1),k=2,,N,\displaystyle\Tr_{2k-1}P^{(k)}=\mathds{1}_{2k-2}\otimes P^{(k-1)},\ k=2,\dots,N,
TrP(1)=1,\displaystyle\Tr P^{(1)}=1,

from which it follows that the dual problem is

minλ,Q(k),h\displaystyle\min_{\lambda,Q^{(k)},h} λ,\displaystyle\,\lambda, (78)
s.t.\displaystyle\mathrm{s.t.} λ𝟙2N1Q(N1)Tr2NΩϕ(h),\displaystyle\,\lambda\mathds{1}_{2N-1}\otimes Q^{(N-1)}\geq\Tr_{2N}\Omega_{\phi}(h),
Tr2kQ(k)=𝟙2k1Q(k1),k=2,,N1,\displaystyle\Tr_{2k}Q^{(k)}=\mathds{1}_{2k-1}\otimes Q^{(k-1)},\ k=2,\dots,N-1,
Tr2Q(1)=𝟙1,\displaystyle\Tr_{2}Q^{(1)}=\mathds{1}_{1},

where Q(N1)Q^{(N-1)} is Hermitian.

Similarly, in order to solve the problem via SDP we rewrite it as

minλ,Q(k),h\displaystyle\min_{\lambda,Q^{(k)},h} λ,\displaystyle\,\lambda, (79)
s.t.\displaystyle\mathrm{s.t.} A0,\displaystyle\,A\geq 0,
Tr2kQ(k)=𝟙2k1Q(k1),k=2,,N1,\displaystyle\Tr_{2k}Q^{(k)}=\mathds{1}_{2k-1}\otimes Q^{(k-1)},\ k=2,\dots,N-1,
Tr2Q(1)=𝟙1,\displaystyle\Tr_{2}Q^{(1)}=\mathds{1}_{1},
hr,\displaystyle\,h\in\mathbb{H}_{r},

for

A:=(λ4𝟙(rd2N)n1,1|nr,d2N||n1,1|nr,d2N𝟙2N1Q(N1)),A:=\left(\begin{array}[]{c | c}\frac{\lambda}{4}\mathds{1}\left(rd_{2N}\right)&\begin{array}[]{c}\bra{n_{1,1}}\\ \vdots\\ \bra{n_{r,d_{2N}}}\end{array}\\ \hline\cr\begin{array}[]{ccc}\ket{n_{1,1}}&\ldots&\ket{n_{r,d_{2N}}}\end{array}&\mathds{1}_{2N-1}\otimes Q^{(N-1)}\end{array}\right), (80)

having defined

|ni,j:=j|N~˙ϕ,i,\ket{n_{i,j}}:=\innerproduct{j}{\dot{\tilde{N}}_{\phi,i}^{*}}, (81)

where |N~˙ϕ,j=|N˙ϕ,jik|Nϕ,khkj\ket{\dot{\tilde{N}}_{\phi,j}}=\ket{\dot{N}_{\phi,j}}-\mathrm{i}\sum_{k}\ket{N_{\phi,k}}h_{kj} and {|j,j=1,,d2N}\left\{\ket{j},\ j=1,\dots,d_{2N}\right\} forms an orthonormal basis of 2N\mathcal{H}_{2N}.

E.3 Quantum SWITCH strategies

We first formally define a quantum SWITCH strategy set 𝖲𝖶𝖨\mathsf{SWI} as the collection of P𝖲𝗍𝗋𝖺𝗍P\in\mathsf{Strat} such that

P=(ρT,A,C)|P(SW)P(SW)|,ρT,A,C0,TrρT,A,C=1,P=\left(\rho_{T,A,C}\right)*\outerproduct{P^{(\mathrm{SW})}}{P^{(\mathrm{SW})}},\ \rho_{T,A,C}\geq 0,\ \Tr\rho_{T,A,C}=1, (82)

where |P(SW):=|I\rrangleA,FAπSN[|πC|I\rrangleT,2π(1)1(i=1N1|I\rrangle2π(i),2π(i+1)1)|I\rrangle2π(N),FT|πFC]\ket{P^{(\mathrm{SW})}}:=\lvert I\rrangle_{A,F_{A}}\sum_{\pi\in S_{N}}\left[\ket{\pi}_{C}\lvert I\rrangle_{T,2\pi(1)-1}\left(\otimes_{i=1}^{N-1}\lvert I\rrangle_{2\pi(i),2\pi(i+1)-1}\right)\lvert I\rrangle_{2\pi(N),F_{T}}\ket{\pi}_{F_{C}}\right] corresponds to a (generalized) quantum SWITCH for NN operations, each permutation π\pi is an element of the symmetric group SNS_{N} whose order is N!N!, and {|πC}\{\ket{\pi}_{C}\} forms an orthonormal basis. We suppose each i\mathcal{H}_{i} for i=1,,2Ni=1,\dots,2N has the same dimension d1d_{1}. Ti\mathcal{H}_{T}\simeq\mathcal{H}_{i} denotes the input space of the target system, A\mathcal{H}_{A} the ancillary space, and C\mathcal{H}_{C} the space of the control system. Correspondingly, FT\mathcal{H}_{F_{T}}, FA\mathcal{H}_{F_{A}} and FC\mathcal{H}_{F_{C}} denote the future output spaces of each part. The global future space F=FTFAFC\mathcal{H}_{F}=\mathcal{H}_{F_{T}}\otimes\mathcal{H}_{F_{A}}\otimes\mathcal{H}_{F_{C}}.

Using the quantum SWITCH strategy set, after tracing over the global future space F\mathcal{H}_{F} the QFI evaluation problem is written as

J(𝖲𝖶𝖨)(Nϕ)=\displaystyle J^{(\mathsf{SWI})}(N_{\phi})= minhrmaxP~Tr[P~Ωϕ(h)],\displaystyle\min_{h\in\mathbb{H}_{r}}\max_{\tilde{P}}\Tr\left[\tilde{P}\Omega_{\phi}(h)\right], (83)
s.t.\displaystyle\mathrm{s.t.} P~=πSNqπρ2π(1)1π(i=1N1|I\rrangle2π(i),2π(i+1)1\llangleI|2π(i),2π(i+1)1)𝟙2π(N),\displaystyle\,\tilde{P}=\sum_{\pi\in S_{N}}q^{\pi}\rho_{2\pi(1)-1}^{\pi}\left(\otimes_{i=1}^{N-1}\lvert I\rrangle_{2\pi(i),2\pi(i+1)-1}\llangle I\rvert_{2\pi(i),2\pi(i+1)-1}\right)\otimes\mathds{1}_{2\pi(N)},
πSNqπ=1,\displaystyle\sum_{\pi\in S_{N}}q^{\pi}=1,
ρ2π(1)1π0,Trρ2π(1)1π=1,qπ0,πSN,\displaystyle\,\rho_{2\pi(1)-1}^{\pi}\geq 0,\ \Tr\rho_{2\pi(1)-1}^{\pi}=1,\ q^{\pi}\geq 0,\ \pi\in S_{N},

where the superscript π\pi of an operator denotes a permutation label, and the subscript denotes the subspace it lies in. Note that the primal set of P~\tilde{P} is a convex hull of affine spaces. Equivalently the problem can be rewritten as

minhrmaxρ2π(1)1π,qπ\displaystyle\min_{h\in\mathbb{H}_{r}}\max_{\rho_{2\pi(1)-1}^{\pi},q^{\pi}} πSNTr[qπρ2π(1)1π(i=1N1\llangleI|2π(i),2π(i+1)1)Tr2π(N)Ωϕ(h)(j=1N1|I\rrangle2π(j),2π(j+1)1)],\displaystyle\sum_{\pi\in S_{N}}\Tr\left[q^{\pi}\rho_{2\pi(1)-1}^{\pi}\left(\otimes_{i=1}^{N-1}\llangle I\rvert_{2\pi(i),2\pi(i+1)-1}\right)\Tr_{2\pi(N)}\Omega_{\phi}(h)\left(\otimes_{j=1}^{N-1}\lvert I\rrangle_{2\pi(j),2\pi(j+1)-1}\right)\right], (84)
s.t.\displaystyle\mathrm{s.t.} πSNqπ=1,\displaystyle\sum_{\pi\in S_{N}}q^{\pi}=1,
ρ2π(1)1π0,Trρ2π(1)1π=1,qπ0,πSN.\displaystyle\,\rho_{2\pi(1)-1}^{\pi}\geq 0,\ \Tr\rho_{2\pi(1)-1}^{\pi}=1,\ q^{\pi}\geq 0,\ \pi\in S_{N}.

Following the method in the proof of Theorem 1, the dual problem is given by

min\displaystyle\min λ,\displaystyle\,\lambda, (85)
s.t.\displaystyle\mathrm{s.t.} λ𝟙2π(1)1Ωϕπ(h),Ωϕπ(h):=(i=1N1\llangleI|2π(i),2π(i+1)1)Tr2π(N)Ωϕ(h)(j=1N1|I\rrangle2π(j),2π(j+1)1),πSN.\displaystyle\,\lambda\mathds{1}_{2\pi(1)-1}\geq\Omega_{\phi}^{\pi}(h),\ \Omega_{\phi}^{\pi}(h):=\left(\otimes_{i=1}^{N-1}\llangle I\rvert_{2\pi(i),2\pi(i+1)-1}\right)\Tr_{2\pi(N)}\Omega_{\phi}(h)\left(\otimes_{j=1}^{N-1}\lvert I\rrangle_{2\pi(j),2\pi(j+1)-1}\right),\ \pi\in S_{N}.

Equivalently in an SDP form the problem is written as

minλ,h\displaystyle\min_{\lambda,h} λ,\displaystyle\,\lambda, (86)
s.t.\displaystyle\mathrm{s.t.} Aπ0,πSN,\displaystyle\,A^{\pi}\geq 0,\ \pi\in S_{N},
hr,\displaystyle\,h\in\mathbb{H}_{r},

having defined

Aπ:=(λ4𝟙(rd1)n1,1π|nr,d1π||n1,1π|nr,d1π𝟙2π(1)1),A^{\pi}:=\left(\begin{array}[]{c | c}\frac{\lambda}{4}\mathds{1}(rd_{1})&\begin{array}[]{c}\bra{n_{1,1}^{\pi}}\\ \vdots\\ \bra{n_{r,d_{1}}^{\pi}}\end{array}\\ \hline\cr\begin{array}[]{ccc}\ket{n_{1,1}^{\pi}}&\ldots&\ket{n_{r,d_{1}}^{\pi}}\end{array}&\mathds{1}_{2\pi(1)-1}\end{array}\right), (87)

for

|ni,jπ:=jπ|(k=1N1\llangleI|2π(k),2π(k+1)1)|N~˙ϕ,i,\ket{n_{i,j}^{\pi}}:=\bra{j^{\pi}}\left(\otimes_{k=1}^{N-1}\llangle I\rvert_{2\pi(k),2\pi(k+1)-1}\right)\ket{\dot{\tilde{N}}_{\phi,i}^{*}}, (88)

where |N~˙ϕ,j=|N˙ϕ,jik|Nϕ,khkj\ket{\dot{\tilde{N}}_{\phi,j}}=\ket{\dot{N}_{\phi,j}}-\mathrm{i}\sum_{k}\ket{N_{\phi,k}}h_{kj} and {|jπ}\{\ket{j^{\pi}}\} forms an orthonormal basis of 2π(N)\mathcal{H}_{2\pi(N)}.

Symmetry reduction.—For 𝖲𝖶𝖨\mathsf{SWI}, by Theorem 2, hh can be taken to be permutation invariant and the constraint corresponding to one permutation π\pi (e.g., the identity element of SNS_{N}) is sufficient. As explained in Appendix D we can decompose the permutation-invariant h=U[|μ|=N𝟙(dμ)hμ]Uh=U^{\prime}\left[\bigoplus_{|\mu|=N}\mathds{1}\left(d_{\mu}\right)\otimes h^{\mu}\right]U^{\prime\dagger} with hμh^{\mu} as an mμ×mμm_{\mu}^{\prime}\times m_{\mu}^{\prime} matrix. We define

𝐧j2Nμ:=j2N|(i=1N1\llangleI|2i,2i+1)𝐍~˙ϕμ,\mathbf{n}_{j_{2N}}^{\mu}:=\bra{j_{2N}}\left(\otimes_{i=1}^{N-1}\llangle I\rvert_{2i,2i+1}\right)\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu*}, (89)

where {|j2N}\{\ket{j_{2N}}\} forms an orthonormal basis of 2N\mathcal{H}_{2N} and 𝐍~˙ϕμ\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu} is given by Eq. (42). If we define

A(inv):=(λ4𝟙(rd1)𝐧1μ1𝐧d1μI𝐧1μ1𝐧d1μI𝟙1),A^{(\mathrm{inv})}:=\left(\begin{array}[]{c | c}\frac{\lambda}{4}\mathds{1}(rd_{1})&\begin{array}[]{c}\mathbf{n}_{1}^{\mu^{1}\dagger}\\ \vdots\\ \mathbf{n}_{d_{1}}^{\mu^{I}\dagger}\end{array}\\ \hline\cr\begin{array}[]{ccc}\mathbf{n}_{1}^{\mu^{1}}&\ldots&\mathbf{n}_{d_{1}}^{\mu^{I}}\end{array}&\mathds{1}_{1}\end{array}\right), (90)

then by Theorem 2 we can reformulate the optimization problem as

minλ,hμ\displaystyle\min_{\lambda,h^{\mu}} λ,\displaystyle\,\lambda, (91)
s.t.\displaystyle\mathrm{s.t.} A(inv)0,\displaystyle\,A^{(\mathrm{inv})}\geq 0,
hμmμ,|μ|=N.\displaystyle\,h^{\mu}\in\mathbb{H}_{m_{\mu}^{\prime}},\ |\mu|=N.

E.4 Causal superposition strategies

Following a similar route for the causal superposition strategy set the problem can be written as

J(𝖲𝗎𝗉)(Nϕ)=\displaystyle J^{(\mathsf{Sup})}(N_{\phi})= minhrmaxPπ,(N),qππSNTr(qπPπ,(N)Tr2π(N)Ωϕ(h)),\displaystyle\min_{h\in\mathbb{H}_{r}}\max_{P^{\pi,(N)},q^{\pi}}\sum_{\pi\in S_{N}}\Tr\left(q^{\pi}P^{\pi,(N)}\Tr_{2\pi(N)}\Omega_{\phi}(h)\right), (92)
s.t.\displaystyle\mathrm{s.t.} πSNqπ=1,\displaystyle\sum_{\pi\in S_{N}}q^{\pi}=1,
qπ0,Pπ,(N)0,TrPπ,(1)=1,Tr2k1Pπ,(k)=𝟙2k2Pπ,(k1)fork=2,,N,πSN.\displaystyle\,q^{\pi}\geq 0,\ P^{\pi,(N)}\geq 0,\ \Tr P^{\pi,(1)}=1,\ \Tr_{2k-1}P^{\pi,(k)}=\mathds{1}_{2k-2}\otimes P^{\pi,(k-1)}\ \mathrm{for}\ k=2,\dots,N,\ \pi\in S_{N}.

For each causal order in the superposition the dual affine space is the set of dual combs without the positivity constraint. Thus the dual problem is given by

min\displaystyle\min λ,\displaystyle\,\lambda, (93)
s.t.\displaystyle\mathrm{s.t.} λ𝟙2π(N)1Qπ,(N1)Tr2π(N)Ωϕ(h),Tr2π(1)Qπ,(1)=𝟙2π(1)1,πSN,\displaystyle\,\lambda\mathds{1}_{2\pi(N)-1}\otimes Q^{\pi,(N-1)}\geq\Tr_{2\pi(N)}\Omega_{\phi}(h),\ \Tr_{2\pi(1)}Q^{\pi,(1)}=\mathds{1}_{2\pi(1)-1},\ \pi\in S_{N},
Tr2π(k)Qπ,(k)=𝟙2π(k)1Qπ,(k1)fork=2,,N1,πSN,\displaystyle\Tr_{2\pi(k)}Q^{\pi,(k)}=\mathds{1}_{2\pi(k)-1}\otimes Q^{\pi,(k-1)}\ \mathrm{for}\ k=2,\dots,N-1,\ \pi\in S_{N},

where the constraints hold for any πSN\pi\in S_{N}. To solve the problem via SDP we can formulate it as

minλ,h\displaystyle\min_{\lambda,h} λ,\displaystyle\,\lambda, (94)
s.t.\displaystyle\mathrm{s.t.} Aπ0,Tr2π(1)Qπ,(1)=𝟙2π(1)1,Tr2π(k)Qπ,(k)=𝟙2π(k)1Qπ,(k1)fork=2,,N1,πSN,\displaystyle\,A^{\pi}\geq 0,\ \Tr_{2\pi(1)}Q^{\pi,(1)}=\mathds{1}_{2\pi(1)-1},\ \Tr_{2\pi(k)}Q^{\pi,(k)}=\mathds{1}_{2\pi(k)-1}\otimes Q^{\pi,(k-1)}\ \mathrm{for}\ k=2,\dots,N-1,\ \pi\in S_{N},
hr,\displaystyle\,h\in\mathbb{H}_{r},

having defined

Aπ:=(λ4𝟙(rd2π(N))n1,1π|nr,d2π(N)π||n1,1π|nr,d2π(N)π𝟙2π(N)1Qπ,(N1)),A^{\pi}:=\left(\begin{array}[]{c | c}\frac{\lambda}{4}\mathds{1}\left(rd_{2\pi(N)}\right)&\begin{array}[]{c}\bra{n_{1,1}^{\pi}}\\ \vdots\\ \bra{n_{r,d_{2\pi(N)}}^{\pi}}\end{array}\\ \hline\cr\begin{array}[]{ccc}\ket{n_{1,1}^{\pi}}&\ldots&\ket{n_{r,d_{2\pi(N)}}^{\pi}}\end{array}&\mathds{1}_{2\pi(N)-1}\otimes Q^{\pi,(N-1)}\end{array}\right), (95)

for

|ni,jπ:=jπ|N~˙ϕ,i,\ket{n_{i,j}^{\pi}}:=\bra{j^{\pi}}\ket{\dot{\tilde{N}}_{\phi,i}^{*}}, (96)

where |N~˙ϕ,j=|N˙ϕ,jik|Nϕ,khkj\ket{\dot{\tilde{N}}_{\phi,j}}=\ket{\dot{N}_{\phi,j}}-\mathrm{i}\sum_{k}\ket{N_{\phi,k}}h_{kj} and {|jπ}\{\ket{j^{\pi}}\} forms an orthonormal basis of 2π(N)\mathcal{H}_{2\pi(N)}.

Symmetry reduction.—Similar to 𝖲𝖶𝖨\mathsf{SWI}, by Theorem 2, for 𝖲𝗎𝗉\mathsf{Sup} we take a permutation-invariant hh and only need the constraint corresponding to the identity element of SNS_{N}. As explained in Appendix D we can decompose the permutation-invariant h=U[|μ|=N𝟙(dμ)hμ]Uh=U^{\prime}\left[\bigoplus_{|\mu|=N}\mathds{1}\left(d_{\mu}\right)\otimes h^{\mu}\right]U^{\prime\dagger} with hμh^{\mu} as an mμ×mμm_{\mu}^{\prime}\times m_{\mu}^{\prime} matrix. We define

𝐧j2Nμ:=j2N|𝐍~˙ϕμ,\mathbf{n}_{j_{2N}}^{\mu}:=\bra{j_{2N}}\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu*}, (97)

where {|j2N}\{\ket{j_{2N}}\} forms an orthonormal basis of 2N\mathcal{H}_{2N} and 𝐍~˙ϕμ\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu} is given by Eq. (42). If we define

A(inv):=(λ4𝟙(rd2N)𝐧1μ1𝐧d1μI𝐧1μ1𝐧d1μI𝟙2N1Q(N1)),A^{(\mathrm{inv})}:=\left(\begin{array}[]{c | c}\frac{\lambda}{4}\mathds{1}\left(rd_{2N}\right)&\begin{array}[]{c}\mathbf{n}_{1}^{\mu^{1}\dagger}\\ \vdots\\ \mathbf{n}_{d_{1}}^{\mu^{I}\dagger}\end{array}\\ \hline\cr\begin{array}[]{ccc}\mathbf{n}_{1}^{\mu^{1}}&\ldots&\mathbf{n}_{d_{1}}^{\mu^{I}}\end{array}&\mathds{1}_{2N-1}\otimes Q^{(N-1)}\end{array}\right), (98)

then by Theorem 2 we can reformulate the optimization problem as

minλ,Q(k),h\displaystyle\min_{\lambda,Q^{(k)},h} λ,\displaystyle\,\lambda, (99)
s.t.\displaystyle\mathrm{s.t.} A(inv)0,\displaystyle\,A^{(\mathrm{inv})}\geq 0,
Tr2kQ(k)=𝟙2k1Q(k1),k=2,,N1,\displaystyle\Tr_{2k}Q^{(k)}=\mathds{1}_{2k-1}\otimes Q^{(k-1)},\ k=2,\dots,N-1,
Tr2Q(1)=𝟙1,\displaystyle\Tr_{2}Q^{(1)}=\mathds{1}_{1},
hμmμ,|μ|=N.\displaystyle\,h^{\mu}\in\mathbb{H}_{m_{\mu}^{\prime}},\ |\mu|=N.

E.5 General indefinite-causal-order strategies

In this case the explicit linear constraints on strategies have been derived in [21], and the dual affine space turns out to be the set of CJ operators of NN-partite no-signaling quantum channels without the positivity constraint [16, 43], mathematically defined by

Tr2kQ\displaystyle\Tr_{2k}Q =𝟙2k1d2k1Tr2k1,2kQ,k=1,,N,\displaystyle=\frac{\mathds{1}_{2k-1}}{d_{2k-1}}\otimes\Tr_{2k-1,2k}Q,\ k=1,\dots,N, (100)
TrQ\displaystyle\Tr Q =i=1Nd2i1.\displaystyle=\prod_{i=1}^{N}d_{2i-1}.

The intuitive interpretation for no-signaling channels is that locally the input of each channel only affects the output of this single channel, but cannot transmit any information to N1N-1 other channels. To solve the QFI evaluation problem via SDP we can write it in the form

minλ,Q,h\displaystyle\min_{\lambda,Q,h} λ,\displaystyle\,\lambda, (101)
s.t.\displaystyle\mathrm{s.t.} A0,\displaystyle\,A\geq 0,
Tr2kQ=𝟙2k1d2k1Tr2k1,2kQ,k=1,,N,\displaystyle\Tr_{2k}Q=\frac{\mathds{1}_{2k-1}}{d_{2k-1}}\otimes\Tr_{2k-1,2k}Q,\ k=1,\dots,N,
TrQ=i=1Nd2i1,\displaystyle\Tr Q=\prod_{i=1}^{N}d_{2i-1},
hr,\displaystyle\,h\in\mathbb{H}_{r},

having defined

A=(λ4𝟙(r)N~˙ϕ,1|N~˙ϕ,r||N~˙ϕ,1|N~˙ϕ,rQ).A=\left(\begin{array}[]{c | c}\frac{\lambda}{4}\mathds{1}\left(r\right)&\begin{array}[]{c}\bra{\dot{\tilde{N}}_{\phi,1}^{*}}\\ \vdots\\ \bra{\dot{\tilde{N}}_{\phi,r}^{*}}\end{array}\\ \hline\cr\begin{array}[]{ccc}\ket{\dot{\tilde{N}}_{\phi,1}^{*}}&\ldots&\ket{\dot{\tilde{N}}_{\phi,r}^{*}}\end{array}&Q\end{array}\right). (102)

Symmetry reduction.—For 𝖨𝖢𝖮\mathsf{ICO}, by Lemmas 1 and 2, both hh and QQ can be taken to be permutation invariant. As explained in Appendix D we can decompose the permutation-invariant h=U[|μ|=N𝟙(dμ)hμ]Uh=U^{\prime}\left[\bigoplus_{|\mu|=N}\mathds{1}\left(d_{\mu}\right)\otimes h^{\mu}\right]U^{\prime\dagger} with hμh^{\mu} as an mμ×mμm_{\mu}^{\prime}\times m_{\mu}^{\prime} matrix, and decompose the permutation-invariant Q=U[|μ|=N𝟙(dμ)Qμ]UQ=U\left[\bigoplus_{|\mu|=N}\mathds{1}\left(d_{\mu}\right)\otimes Q^{\mu}\right]U^{\dagger} with QμQ^{\mu} as an mμ×mμm_{\mu}\times m_{\mu} matrix. If we define

Aμ:=(λ4𝟙(dμmμ)(Uμ,1𝐍~˙ϕμ)Uμ,1𝐍~˙ϕμQμ),A^{\mu}:=\left(\begin{array}[]{c | c}\frac{\lambda}{4}\mathds{1}\left(d_{\mu}m_{\mu}^{\prime}\right)&\left(U^{\mu,1\dagger}\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu*}\right)^{\dagger}\\ \hline\cr U^{\mu,1\dagger}\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu*}&Q^{\mu}\end{array}\right), (103)

where 𝐍~˙ϕμ\dot{\tilde{\mathbf{N}}}_{\phi}^{\mu} is given by Eq. (42), then by Theorem 3 we can reformulate the optimization problem as

minλ,Qμ,h\displaystyle\min_{\lambda,Q^{\mu},h} λ,\displaystyle\,\lambda, (104)
s.t.\displaystyle\mathrm{s.t.} Aμ0,hμmμ,|μ|=N,\displaystyle\,A^{\mu}\geq 0,\ h^{\mu}\in\mathbb{H}_{m_{\mu}^{\prime}},\ |\mu|=N,
Tr2NQ=𝟙2N1d2N1Tr2N1,2NQ,\displaystyle\Tr_{2N}Q=\frac{\mathds{1}_{2N-1}}{d_{2N-1}}\otimes\Tr_{2N-1,2N}Q,
|μ|=NdμTrQμ=i=1Nd2i1,\displaystyle\sum_{|\mu|=N}d_{\mu}\Tr Q^{\mu}=\prod_{i=1}^{N}d_{2i-1},

having removed the redundant constraints on QQ by permutation symmetry. Similar to the case of 𝖯𝖺𝗋\mathsf{Par}, we only need to consider the constraint on QQ on the permutation-invariant subspace, since both QQ and Tr2N1,2NQ\Tr_{2N-1,2N}Q are permutation invariant. Therefore, both the number of scalar variables and the number of constraints in terms of scalar variables are polynomial with respect to NN.

Appendix F Complexity analysis

Here we refer to the number of real scalar variables concerned in optimization as the complexity. In Appendix E we have presented both the original and the symmetry reduced QFI evaluation for all families of strategies considered in this work as applicable, from which we can obtain Table 1 in the main text.

Now we consider the algorithm for identifying an optimal strategy. Since the algorithm is based on Theorem 1, its complexity is no less than that of the QFI evaluation. For the second step of the algorithm, by Lemma 4 there exists a permutation-invariant optimal strategy for all families of strategies except for the sequential one, and thus we can apply the group-invariant SDP to achieve the permutation-invariant solution. Recall that P~=i=1KqiSi\tilde{P}=\sum_{i=1}^{K}q^{i}S^{i} for Si𝖲iS^{i}\in\mathsf{S}^{i}. In fact, for 𝖲𝖶𝖨\mathsf{SWI} and 𝖲𝗎𝗉\mathsf{Sup}, we only need to characterize S1𝖲1S^{1}\in\mathsf{S}^{1} and obtain P~=i=1K1KS1\tilde{P}=\sum_{i=1}^{K}\frac{1}{K}S^{1} due to the permutation invariance. In summary, taking both steps of the algorithm into account, we obtain Table 2:

Table 2: Complexity of Algorithm 1 for each family of strategies (with respect to NN). The asymptotic numbers of variables in optimization are compared between the original (Ori.) and group-invariant (Inv.) SDP. We denote d:=d1d2d:=d_{1}d_{2} for di:=𝖽𝗂𝗆(i)d_{i}:=\mathsf{dim}(\mathcal{H}_{i}) and s:=maxϕ𝗋𝖺𝗇𝗄(Eϕ)ds:=\max_{\phi}\mathsf{rank}(E_{\phi})\leq d.
SDP 𝖯𝖺𝗋\mathsf{Par} 𝖲𝖾𝗊\mathsf{Seq} 𝖲𝖶𝖨\mathsf{SWI} 𝖲𝗎𝗉\mathsf{Sup} 𝖨𝖢𝖮\mathsf{ICO}
Ori. O(max(s,d1)N)O\left(\max(s,d_{1})^{N}\right) O(dN)O\left(d^{N}\right) O(N!)O\left(N!\right) O(N!dN)O\left(N!\,d^{N}\right) O(dN)O\left(d^{N}\right)
Inv. O(Nd21)O\left(N^{d^{2}-1}\right) O(dN)O\left(d^{N}\right) O(Ns21)O(N^{s^{2}-1}) O(dN)O\left(d^{N}\right) O(Nd21)O\left(N^{d^{2}-1}\right)

As a concrete example, we apply the group-invariant SDP to the evaluation of the QFI J(𝖨𝖢𝖮)J^{(\mathsf{ICO})} for the general indefinite-causal-order strategies up to N=5N=5. We consider the amplitude damping noise and take ϕ=1.0\phi=1.0, t=1.0t=1.0 and p=0.5p=0.5. As illustrated in Fig. 3, the growth of J(𝖨𝖢𝖮)J^{(\mathsf{ICO})} is faster than linear growth but slower than quadratic growth. Table 3 compares the complexity between the original and the group-invariant SDP and indicates that the symmetry reduced approach can save the computational resources dramatically.

Refer to caption
Figure 3: Growth of QFI J(𝖨𝖢𝖮)J^{(\mathsf{ICO})} as NN increases. For the amplitude damping noise, we take ϕ=1.0\phi=1.0, t=1.0t=1.0 and p=0.5p=0.5. The dashed and dotted lines illustrate the linear and quadratic growth with respect to NN respectively, while matching the QFI at N=1N=1.
Table 3: Number of real scalar variables concerned in the evaluation of J(𝖨𝖢𝖮)J^{(\mathsf{ICO})}. For the original SDP the total number of real scalar variables for QQ, hh and λ\lambda is 16N+4N+116^{N}+4^{N}+1, while for the group-invariant SDP the total number of real scalar variables for QμQ^{\mu}, hμh^{\mu} and λ\lambda for all Young diagram labels |μ|=N|\mu|=N is (N+1515)+(N+33)+1\binom{N+15}{15}+\binom{N+3}{3}+1.
NN 1 2 3 4 5
No. of variables in the original SDP 21 273 4161 65793 1049601
No. of variables in the group-invariant SDP 21 147 837 3912 15561

Appendix G Supplementary numerical results

G.1 Hierarchy for the N=3N=3 case

Numerical results in this work are obtained by implementing SDP using the open source Python package CVXPY [66, 67] with the solver MOSEK [68]. We plot the QFI for N=3N=3, amplitude damping noise in Fig. 4 and observe a similar hierarchy of the estimation performance using parallel, sequential and indefinite-causal-order strategies. Different from the N=2N=2 case presented in the main text, general indefinite-causal-order strategies indeed provide a small advantage over causal superposition strategies, which is presented in Table. 4.

Refer to caption
Figure 4: Hierarchy of QFI using parallel, sequential, and indefinite-causal-order strategies for N=3N=3, the amplitude damping noise. We take ϕ=1.0\phi=1.0, fix the evolution time t=1.0t=1.0 and vary the decay parameter pp. To see the gaps between different strategies more clearly we insert an inset plot where the decay parameter pp ranges from 0.1 to 0.2.
Table 4: Hierarchy of QFI using 𝖨𝖢𝖮\mathsf{ICO} and 𝖲𝗎𝗉\mathsf{Sup}.
pp J(𝖲𝗎𝗉)(Nϕ)J^{(\mathsf{Sup})}(N_{\phi}) J(𝖨𝖢𝖮)(Nϕ)J^{(\mathsf{ICO})}(N_{\phi})
0.1 8.185 8.200
0.2 7.364 7.375
0.3 6.523 6.524
0.4 5.642 5.647
0.5 4.725 4.743
0.6 3.786 3.815
0.7 2.832 2.870
0.8 1.871 1.909
0.9 0.918 0.930

On the other hand, analogous to the N=2N=2 case, simple quantum SWITCH strategies without any additional intermediate control operations could have advantage over any definite-causal-order strategies when the decay parameter pp is small, but this advantage becomes more insignificant. This should not be surprising since the control can make a bigger difference as NN grows.

G.2 Estimation of randomly sampled channels

To demonstrate the universality of the hierarchy of different families of strategies considered in the main text, we randomly sample noise channels drawn from an ensemble of CPTP maps defined by Bruzda et al. in [69]. In this work we only sample rank-2 qubit channels for N=2N=2, which is enough to show the hierarchy. The sampling process is implemented via an open source Python package QuTiP [70, 71]. We set an error tolerance of 10810^{-8}, i.e., we claim J1>J2J_{1}>J_{2} only if the gap is no smaller than 10810^{-8}. We find that for 984 of 1000 random channels, a strict hierarchy J(𝖯𝖺𝗋)<J(𝖲𝖾𝗊)<J(𝖲𝗎𝗉)<J(𝖨𝖢𝖮)J^{(\mathsf{Par})}<J^{(\mathsf{Seq})}<J^{(\mathsf{Sup})}<J^{(\mathsf{ICO})} holds, implying that general indefinite-causal-order strategies can provide advantage over causal superposition strategies. In addition, we find that of the same 1000 channels J(𝖯𝖺𝗋)<J(𝖲𝖶𝖨)J^{(\mathsf{Par})}<J^{(\mathsf{SWI})} for 34 channels and J(𝖲𝖾𝗊)<J(𝖲𝖶𝖨)J^{(\mathsf{Seq})}<J^{(\mathsf{SWI})} only for 1 channel, so with a high probability quantum SWITCH strategies cannot outperform strategies following definite causal order for a random noise channel, which highlights the estimation enhancement from intermediate control in the general case.

Appendix H Comparison with asymptotic results

In this section we focus on strategies following definite causal order, i.e., parallel and sequential ones, and compare our results and those of the extensively studied asymptotic theory.

H.1 Preliminaries

We first introduce some basic notions. If we write the operation-sum representation of the channel ϕ(ρ)=i=1rKϕ,iρKϕ,i\mathcal{E}_{\phi}(\rho)=\sum_{i=1}^{r}K_{\phi,i}^{\dagger}\rho K_{\phi,i}, where {Kϕ,i}\{K_{\phi,i}\} are a set of Kraus operators and rr is the rank of the channel, the channel QFI can be evaluated by optimization:

JQ(chan)(ϕ)=4minhrα,J_{Q}^{(\mathrm{chan})}(\mathcal{E}_{\phi})=4\min_{h\in\mathbb{H}_{r}}\lVert\alpha\rVert, (105)

where \lVert\cdot\rVert denotes the operator norm and α=iK~˙ϕ,iK~˙ϕ,i\alpha=\sum_{i}\dot{\tilde{K}}^{\dagger}_{\phi,i}\dot{\tilde{K}}_{\phi,i}. Here K~˙ϕ,i=K˙ϕ,iij=1rhijKϕ,j\dot{\tilde{K}}_{\phi,i}=\dot{K}_{\phi,i}-\mathrm{i}\sum_{j=1}^{r}h_{ij}K_{\phi,j} is nothing but the derivative of an equivalent Kraus representation, given an r×rr\times r Hermitian matrix hh.

The upper bounds on QFI of NN quantum channels have been derived for both sequential and parallel strategies. For parallel strategies an asymptotically tight upper bound is [28, 30]

J(𝖯𝖺𝗋)(Nϕ)4minhr[Nα+N(N1)β2],J^{(\mathsf{Par})}(N_{\phi})\leq 4\min_{h\in\mathbb{H}_{r}}[N\lVert\alpha\rVert+N(N-1)\lVert\beta\rVert^{2}], (106)

where β=iiKϕ,iK~˙ϕ,i\beta=\mathrm{i}\sum_{i}K_{\phi,i}^{\dagger}\dot{\tilde{K}}_{\phi,i}. An upper bound was also derived for sequential strategies [7, 45]:

J(𝖲𝖾𝗊)(Nϕ)4minhr[Nα+N(N1)β(β+2α)].J^{(\mathsf{Seq})}(N_{\phi})\leq 4\min_{h\in\mathbb{H}_{r}}[N\lVert\alpha\rVert+N(N-1)\lVert\beta\rVert(\lVert\beta\rVert+2\sqrt{\lVert\alpha\rVert})]. (107)

It has been shown that the QFI follows the standard quantum limit if and only if there exists an hh such that β=0\beta=0 [45]. In this case sequential strategies provide no advantage asymptotically, and we have

limN1NJ(𝖯𝖺𝗋)(Nϕ)=limN1NJ(𝖲𝖾𝗊)(Nϕ)=4minhrs.t.β=0α.\lim_{N\rightarrow\infty}\frac{1}{N}J^{(\mathsf{Par})}(N_{\phi})=\lim_{N\rightarrow\infty}\frac{1}{N}J^{(\mathsf{Seq})}(N_{\phi})=4\min_{h\in\mathbb{H}_{r}\ \mathrm{s.t.}\ \beta=0}\lVert\alpha\rVert. (108)

We remark that the minimization in Eq. (106) can be efficiently evaluated via SDP [30].

H.2 Tightness of QFI bounds in nonasymptotic channel estimation

We compare our nonasymptotic results and existing asymptotically tight bounds for two types of quantum channels. Apart from the amplitude damping noise described by K1(AD)=|00|+1p|11|K_{1}^{(\mathrm{AD})}=\outerproduct{0}{0}+\sqrt{1-p}\outerproduct{1}{1} and K2(AD)=p|01|K_{2}^{(\mathrm{AD})}=\sqrt{p}\outerproduct{0}{1} considered in the main text, here we also present a second example, where the noise is described by a SWAP-type interaction Vint=eigτHSWAPV_{\mathrm{int}}=e^{-\mathrm{i}g\tau H_{\mathrm{SWAP}}} between a qubit system SS and a qubit environment EE, where gg is the interaction strength, τ\tau is the interaction time and the Hamiltonian is given by HSWAP(|iS|jE)=|jS|iEH_{\mathrm{SWAP}}(\ket{i}_{S}\ket{j}_{E})=\ket{j}_{S}\ket{i}_{E}. The initial environment state is |0\ket{0}, and the Kraus operators can be written as K1(SWAP)=0|EVint|0E=eigτ|00|+cos(gτ)|11|K_{1}^{(\mathrm{SWAP})}=\bra{0}_{E}V_{\mathrm{int}}\ket{0}_{E}=e^{-\mathrm{i}g\tau}\outerproduct{0}{0}+\cos(g\tau)\outerproduct{1}{1}, and K2(SWAP)=1|EVint|0E=isin(gτ)|01|K_{2}^{(\mathrm{SWAP})}=\bra{1}_{E}V_{\mathrm{int}}\ket{0}_{E}=-\mathrm{i}\sin(g\tau)\outerproduct{0}{1}.

We plot the QFI for the two examples in Fig. 5. Both of them show the advantage of sequential strategies over parallel ones, and the gaps between exact results of QFI for sequential strategies and the parallel upper bounds given by Eq. (106).

\sidesubfloat

[]Refer to caption \sidesubfloat[]Refer to caption
\sidesubfloat[]Refer to caption \sidesubfloat[]Refer to caption

Figure 5: Comparison of our results with the existing asymptotically tight QFI bound. We compare our results to the asymptotically tight bound on the maximal QFI of parallel strategies. We plot the QFI versus the evolution time tt for N=2N=2, ϕ=1.0\phi=1.0, g=1.0g=1.0, τ=t\tau=t and p=0.5p=0.5. LABEL:sub@subfig:N_2_ad N=2N=2 and LABEL:sub@subfig:N_3_ad N=3N=3 for the amplitude damping noise. LABEL:sub@subfig:N_2_swap N=2N=2 and LABEL:sub@subfig:N_3_swap N=3N=3 for the SWAP-type noise.

H.3 Elusive advantage of sequential strategies in the asymptotic limit

We observe a gap between parallel and sequential strategies for amplitude damping channels and SWAP-type interactions for N=2N=2 and 33. Now we show that for both examples there is no advantage of sequential strategies asymptotically since there exists an hh such that β=0\beta=0.

For the amplitude damping channel, Kϕ,i(AD)=Ki(AD)Uz(ϕ),i=1,2K_{\phi,i}^{(\mathrm{AD})}=K_{i}^{(\mathrm{AD})}U_{z}(\phi),\ i=1,2. Direct calculation leads to

β(AD)=(t2+h11(AD))|00|+[h11(AD)t2+(h22(AD)h11(AD))p]|11|.\beta^{(\mathrm{AD})}=\left(\frac{t}{2}+h^{(\mathrm{AD})}_{11}\right)\outerproduct{0}{0}+\left[h^{(\mathrm{AD})}_{11}-\frac{t}{2}+\left(h^{(\mathrm{AD})}_{22}-h^{(\mathrm{AD})}_{11}\right)p\right]\outerproduct{1}{1}. (109)

To obtain β(AD)=0\beta^{(\mathrm{AD})}=0 we just need to take h11(AD)=t/2h^{(\mathrm{AD})}_{11}=-t/2 and h22(AD)=(2p)t/2ph^{(\mathrm{AD})}_{22}=(2-p)t/2p.

Similarly, for the SWAP-type interaction we have

β(SWAP)=(t2+h11(SWAP))|00|+[h11(SWAP)t2+(h22(SWAP)h11(SWAP))sin2(gτ)]|11|.\beta^{(\mathrm{SWAP})}=\left(\frac{t}{2}+h^{(\mathrm{SWAP})}_{11}\right)\outerproduct{0}{0}+\left[h^{(\mathrm{SWAP})}_{11}-\frac{t}{2}+\left(h^{(\mathrm{SWAP})}_{22}-h^{(\mathrm{SWAP})}_{11}\right)\sin^{2}(g\tau)\right]\outerproduct{1}{1}. (110)

Thus there exists h11(SWAP)=t/2h^{(\mathrm{SWAP})}_{11}=-t/2 and h22(SWAP)=[2sin2(gτ)]t/2sin2(gτ)h^{(\mathrm{SWAP})}_{22}=\left[2-\sin^{2}(g\tau)\right]t/2\sin^{2}(g\tau) such that β(SWAP)=0\beta^{(\mathrm{SWAP})}=0.

Appendix I Implementation of optimal strategies with universal quantum gates

In this section we apply Algorithm 1 in the main text to numerically solve for an optimal strategy in the set of sequential and causal superposition ones respectively. The CJ operator of an optimal sequential strategy corresponds to a sequence of isometries with a minimal ancilla-space implementation provided by [46], and can then be decomposed into single-qubit gates and CNOT gates [72]. By taking advantage of the freedom of choosing a parameter-independent unitary on the final output state, we can adjust the strategy to reduce the CNOT count without affecting the QFI. In terms of an optimal causal superposition strategy we follow the same routine for each sequential strategy branch respectively in the superposition.

I.1 Optimal sequential strategy

Let 0=\mathcal{H}_{0}=\mathbb{C} and 2N+1=F\mathcal{H}_{2N+1}=\mathcal{H}_{F}, and we have the CJ operator of a sequential strategy P(Fi=02Ni)P\in\mathcal{L}\left(\mathcal{H}_{F}\otimes_{i=0}^{2N}\mathcal{H}_{i}\right) (an (N+1)(N+1)-step quantum comb), as illustrated in Fig. 6. In this way of relabeling Hilbert spaces we have

P=P(N+1)0,Tr2k1P(k)=𝟙2k2P(k1)fork=2,,N+1,TrP(1)=P(0)=1.P=P^{(N+1)}\geq 0,\ \Tr_{2k-1}P^{(k)}=\mathds{1}_{2k-2}\otimes P^{(k-1)}\ \mathrm{for}\ k=2,\dots,N+1,\ \Tr P^{(1)}=P^{(0)}=1. (111)
Refer to caption
Figure 6: Concatenation of a sequential strategy and NN quantum channels. PP is the CJ operator of a sequential strategy and EϕE_{\phi} is the CJ operator of a parametrized quantum channel. missingH0=\mathcal{\mathcal{missing}}H_{0}=\mathbb{C} is a trivial one-dimensional Hilbert space, and therefore the first step of the strategy is the process of state preparation.

According to Theorems 1 and 2 in [46], PP corresponds to a sequence of isometries {V(k)}\{V^{(k)}\} for k=1,,N+1k=1,\dots,N+1 by Stinespring dilation:

𝒫(ρ)=TrAN+1[(V(N+1)𝟙1,3,,2N1)(V(1)𝟙2,4,,2N)ρ(V(1)𝟙2,4,,2N)(V(N+1)𝟙1,3,,2N1)]\mathcal{P}(\rho)=\Tr_{A_{N+1}}\left[\left(V^{(N+1)}\otimes\mathds{1}_{1,3,\dots,2N-1}\right)\cdots\left(V^{(1)}\otimes\mathds{1}_{2,4,\dots,2N}\right)\rho\left(V^{(1)}\otimes\mathds{1}_{2,4,\dots,2N}\right)^{\dagger}\cdots\left(V^{(N+1)}\otimes\mathds{1}_{1,3,\dots,2N-1}\right)^{\dagger}\right] (112)

for any input state ρ(i=0N2i)\rho\in\mathcal{L}\left(\otimes_{i=0}^{N}\mathcal{H}_{2i}\right), where the whole process corresponding to PP is described by an isometry 𝒫(i=0N2i,i=0N2i+1)\mathcal{P}\in\mathcal{L}\left(\otimes_{i=0}^{N}\mathcal{H}_{2i},\otimes_{i=0}^{N}\mathcal{H}_{2i+1}\right), and in each step a choice of isometry V(k)(2k2Ak1,2k1Ak)V^{(k)}\in\mathcal{L}\left(\mathcal{H}_{2k-2}\otimes\mathcal{H}_{A_{k-1}},\mathcal{H}_{2k-1}\otimes\mathcal{H}_{A_{k}}\right) with minimal ancilla space is given by

V(k)=𝟙2k1(P(k))12[|I\rrangle2k1,(2k1)𝟙2k2(2k2)(P(k1))12],V^{(k)}=\mathds{1}_{2k-1}\otimes\left(P^{(k)*}\right)^{\frac{1}{2}}\left[\lvert I\rrangle_{2k-1,(2k-1)^{\prime}}\mathds{1}_{2k-2\rightarrow(2k-2)^{\prime}}\otimes\left(P^{(k-1)*}\right)^{-\frac{1}{2}}\right], (113)

where Ak=𝖲𝗎𝗉𝗉(P(k))\mathcal{H}_{A_{k}}=\mathsf{Supp}(P^{(k)*}) is an ancillary space given by the support of P(k)P^{(k)*} with A0=\mathcal{H}_{A_{0}}=\mathbb{C}, i\mathcal{H}_{i^{\prime}} is a copy of the Hilbert space i\mathcal{H}_{i}, 𝟙2k2(2k2):=i|i(2k2)i|2k2\mathds{1}_{2k-2\rightarrow(2k-2)^{\prime}}:=\sum_{i}\ket{i}_{(2k-2)^{\prime}}\bra{i}_{2k-2} is an identity map from 2k2\mathcal{H}_{2k-2} to (2k2)\mathcal{H}_{(2k-2)^{\prime}}, and (P(k1))12\left(P^{(k-1)*}\right)^{-\frac{1}{2}} denotes the Moore–Penrose pseudoinverse of (P(k1))12\left(P^{(k-1)*}\right)^{\frac{1}{2}} with its support on Ak1\mathcal{H}_{A_{k-1}}.

As the last isometry V(N+1)V^{(N+1)} preserves the QFI, it is therefore only necessary to consider the implementation of P(N)P^{(N)} instead of the full strategy P(N+1)P^{(N+1)}. From this explicit construction it follows that the minimal dimension of the ancilla space for implementing the sequential strategy PP is 𝖽𝗂𝗆(AN)=𝗋𝖺𝗇𝗄(P(N))\mathsf{dim}(\mathcal{H}_{A_{N}})=\mathsf{rank}(P^{(N)}). In the case of N=2N=2 for the amplitude damping noise considered in the main text, it is easy to see that 𝖽𝗂𝗆(A1)2\mathsf{dim}(\mathcal{H}_{A_{1}})\leq 2 and 𝖽𝗂𝗆(A2)8\mathsf{dim}(\mathcal{H}_{A_{2}})\leq 8, so V(1)V^{(1)} is an isometry from 0 to (at most) 22 qubits and V(2)V^{(2)} is an isometry from 22 to (at most) 44 qubits, as illustrated in Fig. 7.

\Qcircuit@C=1em @R=.7em \lstick|0⟩ & \multigate1V^(1) \gateE_ϕ \multigate3V^(2) \gateE_ϕ \qw
\lstick|0⟩ \ghostV^(1) \qw \ghostV^(2) \qw \qw
\lstick|0⟩ \qw \qw \ghostV^(2) \qw \qw
\lstick|0⟩ \qw \qw \ghostV^(2) \qw \qw

Figure 7: A sequence of isometries corresponding to a sequential strategy for N=2N=2. The first qubit is the system qubit going through the channel ϕ\mathcal{E}_{\phi} twice, while the three other qubits are ancillary.

Next, we apply a circuit decomposition of each isometry into single-qubit gates and CNOT gates. In practice it is often desirable to achieve a CNOT count as low as possible. Note that V(1)V^{(1)} actually corresponds to the preparation of a 22-qubit state, which in general requires only one CNOT gate [73]. In terms of V(2)V^{(2)}, an isometry from 22 to 44 qubits, the state-of-the-art decomposition scheme is the column-by-column approach which requires at most 54 CNOT gates [72]. However, as an arbitrary parameter-independent unitary on 33 ancillae can always be deferred and regrouped into V(3)V^{(3)} and therefore does not affect the QFI, in fact we can choose a proper V(2)V^{(2)} to further reduce the worst CNOT count to 4747 without changing the QFI. To see this we need to briefly introduce the main ideas behind the column-by-column decomposition scheme.

An isometry VV from mm to nn qubits (mnm\leq n) can be represented in the matrix form by V=U𝟙(2n×2m)V=U^{\dagger}\mathds{1}(2^{n}\times 2^{m}), where UU^{\dagger} is a 2n×2n2^{n}\times 2^{n} unitary matrix and 𝟙(2n×2m)\mathds{1}(2^{n}\times 2^{m}) is the first 2m2^{m} columns of the 2n×2n2^{n}\times 2^{n} identity matrix. If we obtain a decomposition of UU^{\dagger}, then we can simply initialize the state of the first nmn-m qubits to |0\ket{0} to implement VV. Equivalently we can find a decomposition of UU such that UV=𝟙(2n×2m)UV=\mathds{1}(2^{n}\times 2^{m}) and then inverse the circuit representing UU. The idea is to find a sequence of unitary operations such that U=U2m1U0U=U_{2^{m}-1}\cdots U_{0} transforms VV into 𝟙(2n×2m)\mathds{1}(2^{n}\times 2^{m}) column by column. More specifically, we first choose a proper U0U_{0} to map the first column of VV to the first column of 𝟙(2n×2m)\mathds{1}(2^{n}\times 2^{m}), i.e., U0V|0m=|0nU_{0}V\ket{0}_{m}=\ket{0}_{n}, then choose U1U_{1} satisfying U1U0V|1m=|1nU_{1}U_{0}V\ket{1}_{m}=\ket{1}_{n} as well as U1U0V|0m=|0nU_{1}U_{0}V\ket{0}_{m}=\ket{0}_{n} …until we determine U2m1U_{2^{m}-1}.

Here we only focus on U0U_{0}, the inverse of which can be seen as a process preparing a state V|0mV\ket{0}_{m} from |0n\ket{0}_{n}. In terms of decomposing V(2)V^{(2)} from m=2m=2 to n=4n=4 qubits, preparing a 44-qubit state in general requires 88 CNOT gates [74]. Fortunately, without changing the QFI, we have the freedom to choose a unitary UancU_{\mathrm{anc}} on the ancillae after applying V(2)V^{(2)} such that the state V(2)|02=UancV(2)|02V^{(2)\prime}\ket{0}_{2}=U_{\mathrm{anc}}V^{(2)}\ket{0}_{2} can be prepared using only one CNOT gate. This can be seen by dividing the 44 qubits into two parties, including the single system qubit (in the space S\mathcal{H}_{S}) and the three ancillary qubits (in the space A\mathcal{H}_{A}), and taking the Schimidt decomposition of the 44-qubit state V(2)|02V^{(2)}\ket{0}_{2}

|ψSA:=V(2)|02=i=01λi|eiS|fiA,\ket{\psi}_{SA}:=V^{(2)}\ket{0}_{2}=\sum_{i=0}^{1}\lambda_{i}\ket{e_{i}}_{S}\ket{f_{i}}_{A}, (114)

where {|ei/fiS/A}\{\ket{e_{i}/f_{i}}_{S/A}\} forms an orthonormal basis of S/A\mathcal{H}_{S/A}, and {λi}\{\lambda_{i}\} is a set of nonnegative real numbers satisfying iλi2=1\sum_{i}\lambda_{i}^{2}=1. Therefore, to prepare V(2)|02V^{(2)}\ket{0}_{2}, we only need a local unitary on S\mathcal{H}_{S} to generate i=01λi|iS|0A\sum_{i=0}^{1}\lambda_{i}\ket{i}_{S}\ket{0}_{A}, then apply one CNOT gate taking the system qubit as the control to obtain i=01λi|iS|iA\sum_{i=0}^{1}\lambda_{i}\ket{i}_{S}\ket{i}_{A}, and finally apply local unitary operations US=i|eiSi|SU_{S}=\sum_{i}\ket{e_{i}}_{S}\bra{i}_{S} on the system and UA=i|fiAi|AU_{A}=\sum_{i}\ket{f_{i}}_{A}\bra{i}_{A} on the ancillae respectively. If we take Uanc=UAU_{\mathrm{anc}}=U_{A}^{\dagger}, then it is easy to see that V(2)|02=UancV(2)|02=i=01λi|eiS|iAV^{(2)\prime}\ket{0}_{2}=U_{\mathrm{anc}}V^{(2)}\ket{0}_{2}=\sum_{i=0}^{1}\lambda_{i}\ket{e_{i}}_{S}\ket{i}_{A} can thus be prepared using one CNOT gate. This choice of V(2)V^{(2)\prime} saves 77 CNOT gates compared to the general state preparation scheme, and leads to a worst CNOT count of 4747 in total.

Now we present numerical results of the circuit implementation of an optimal sequential strategy. The decomposition of ismometries is implemented using the Mathematica package UniversalQCompiler [75] based on the method described above. As in the main text, we consider the amplitude damping noise and take N=2N=2, ϕ=1.0\phi=1.0, p=0.5p=0.5 and t=1.0t=1.0. The circuits implementing V(1)V^{(1)} and V(2)V^{(2)\prime} are illustrated in Fig. 8. The state preparation V(1)V^{(1)} requires 11 CNOT gate and the intermediate control operation V(2)V^{(2)\prime} requires 3333 CNOT gates.

\Qcircuit@C=1em @R=.7em \push  & \lstick|0⟩ \gateR_y \gateR_z \ctrl1 \gateR_y \qw \push 
\push  \lstick|0⟩ \gateR_y \qw \targ \qw \qw \push 

(a) Decomposition of V(1)V^{(1)}. For simplicity the angles of single-qubit rotation gates are not depicted.

\Qcircuit@C=.6em @R=.7em & \ctrl1 \ctrl1 \ctrl2 \qw \ctrl2 \ctrl3 \qw \ctrl3 \qw \qw \qw \qw \qw \qw \targ \targ \targ \targ \targ \targ \targ \ctrl2 \qw \ctrl3 \qw \qw \qw \targ \targ \targ \targ \targ \ctrl1 \qw
\targ \targ \qw \ctrl1 \qw \qw \ctrl2 \qw \ctrl1 \qw \ctrl2 \targ \targ \targ \ctrl-1 \qw \ctrl-1 \qw \ctrl-1 \qw \ctrl-1 \qw \qw \qw \targ \targ \targ \qw \qw \ctrl-1 \qw \ctrl-1 \targ \qw
\lstick|0⟩ \qw \qw \targ \targ \targ \qw \qw \qw \targ \ctrl1 \qw \qw \ctrl-1 \qw \qw \qw \qw \ctrl-2 \qw \qw \qw \targ \ctrl1 \qw \qw \ctrl-1 \qw \qw \ctrl-2 \qw \qw \qw \qw \qw
\lstick|0⟩ \qw \qw \qw \qw \qw \targ \targ \targ \qw \targ \targ \ctrl-2 \qw \ctrl-2 \qw \ctrl-3 \qw \qw \qw \ctrl-3 \qw \qw \targ \targ \ctrl-2 \qw \ctrl-2 \ctrl-3 \qw \qw \ctrl-3 \qw \qw \qw

(b) Decomposition of V(2)=UancV(2)V^{(2)\prime}=U_{\mathrm{anc}}V^{(2)}. For simplicity single-qubit gates, which might be required in addition to CNOT gates, are not depicted.
Figure 8: Decomposition of isometries corresponding to an optimal sequential strategy for N=2N=2. We apply V(2)V^{(2)\prime} instead of V(2)V^{(2)} to achieve the maximal QFI with fewer CNOT gates.

I.2 Optimal causal superposition strategy

A causal superposition strategy for estimating NN channels can be implemented by an N!N!-dim quantum control system entangled with N!N! sequential strategies of applying the channels:

P=|PP|for|P=πSN|Pπ|πC,P=\outerproduct{P}{P}\ \mathrm{for}\ \ket{P}=\sum_{\pi\in S_{N}}\ket{P^{\pi}}\ket{\pi}_{C}, (115)

where {|πC}\{\ket{\pi}_{C}\} forms an orthonormal basis of the Hilbert space C\mathcal{H}_{C} of the control system, and each Pπ=|PπPπ|P^{\pi}=\outerproduct{P^{\pi}}{P^{\pi}} is a sequential strategy. Once we obtain an optimal causal superposition strategy by applying Algorithm 1, we can apply the circuit decomposition for each sequential strategy in the superposition, following the method described in Appendix I.1. Taking account of the permutation symmetry, we can choose an optimal causal superposition strategy such that apart from the execution order of two channels, sequential strategies in the decomposition of the strategy are the same, containing the same state preparation and intermediate control.

As a concrete example, we again take N=2N=2, ϕ=1.0\phi=1.0, p=0.5p=0.5 and t=1.0t=1.0 for the amplitude damping noise and present numerical results of the circuit implementation of an optimal causal superposition strategy. As illustrated in Fig. 9, we use the qubit |ψC\ket{\psi}_{C} to coherently control which sequential order is executed. Due to the permutation invariance of the optimal strategy, we can simply control the query order of the identical channels while fixing V(1)V^{(1)} and V(2)V^{(2)} for all sequential orders. In view of this, generally we can use a (2N1)(2N-1)-quantum SWITCH to control the order of NN channels ϕ\mathcal{E}_{\phi} and N1N-1 intermediate control operations V(i)V^{(i)}.

\Qcircuit@C=1em @R=.7em & If |ψC=|0C\ket{\psi}_{C}=\ket{0}_{C},
\lstick|0⟩ \multigate1V^(1) \gateE_ϕ^(1) \multigate3V^(2) \gateE_ϕ^(2) \qw
\lstick|0⟩ \ghostV^(1) \qw \ghostV^(2) \qw \qw
\lstick|0⟩ \qw \qw \ghostV^(2) \qw \qw
\lstick|0⟩ \qw \qw \ghostV^(2) \qw \qw
                   \Qcircuit@C=1em @R=.7em & If |ψC=|1C\ket{\psi}_{C}=\ket{1}_{C},
\lstick|0⟩ \multigate1V^(1) \gateE_ϕ^(2) \multigate3V^(2) \gateE_ϕ^(1) \qw
\lstick|0⟩ \ghostV^(1) \qw \ghostV^(2) \qw \qw
\lstick|0⟩ \qw \qw \ghostV^(2) \qw \qw
\lstick|0⟩ \qw \qw \ghostV^(2) \qw \qw

Figure 9: Sequences of isometries corresponding to each sequential order in the causal superposition for N=2N=2. The first qubit of the circuit is the system qubit, and the query order of two identical channels ϕ(1)\mathcal{E}_{\phi}^{(1)} and ϕ(2)\mathcal{E}_{\phi}^{(2)} is entangled with the state of the control qubit |ψC\ket{\psi}_{C}. When |ψC\ket{\psi}_{C} is a superposition of the two states shown in the figure, the causal order is also in a superposition given by Eq. (115).

Further decomposition shows that each sequential branch requires one CNOT gate for state preparation V(1)V^{(1)} and 36 CNOT gates for intermediate control V(2)V^{(2)}, as illustrated in Fig. 10.

\Qcircuit@C=1em @R=.7em \push  & \lstick|0⟩ \gateR_y \gateR_z \ctrl1 \gateR_y \qw \push 
\push  \lstick|0⟩ \gateR_y \qw \targ \qw \qw \push 

(a) Decomposition of V(1)V^{(1)}. For simplicity the angles of single-qubit rotation gates are not depicted.

\Qcircuit@C=.6em @R=.7em & \ctrl1 \ctrl1 \ctrl2 \qw \ctrl2 \qw \ctrl3 \qw \ctrl3 \qw \qw \qw \qw \qw \qw \qw \qw \qw \targ \targ \targ \targ \targ \targ \targ \ctrl2 \qw \ctrl3 \qw \qw \qw \ctrl1 \targ \targ \targ \ctrl1 \qw
\targ \targ \qw \ctrl1 \qw \qw \qw \ctrl2 \qw \targ \targ \targ \ctrl1 \qw \ctrl2 \targ \targ \targ \ctrl-1 \qw \ctrl-1 \qw \ctrl-1 \qw \ctrl-1 \qw \qw \qw \targ \targ \targ \targ \ctrl-1 \qw \ctrl-1 \targ \qw
\lstick|0⟩ \qw \qw \targ \targ \targ \ctrl1 \qw \qw \qw \qw \ctrl-1 \qw \targ \ctrl1 \qw \qw \ctrl-1 \qw \qw \qw \qw \ctrl-2 \qw \qw \qw \targ \ctrl1 \qw \qw \ctrl-1 \qw \qw \qw \ctrl-2 \qw \qw \qw
\lstick|0⟩ \qw \qw \qw \qw \qw \targ \targ \targ \targ \ctrl-2 \qw \ctrl-2 \qw \targ \targ \ctrl-2 \qw \ctrl-2 \qw \ctrl-3 \qw \qw \qw \ctrl-3 \qw \qw \targ \targ \ctrl-2 \qw \ctrl-2 \qw \qw \qw \qw \qw \qw

(b) Decomposition of V(2)V^{(2)}. For simplicity single-qubit gates, which might be required in addition to CNOT gates, are not depicted.
Figure 10: Decomposition of isometries corresponding to one causal order in the decomposition of an optimal causal superposition strategy for N=2N=2. We have already taken advantage of the freedom to choose a V(2)V^{(2)} implemented by fewer CNOT gates, as explained in Appendix I.1.