Quantum Fisher information

Quantum

The quantum Fisher information is a central quantity in quantum metrology and is the quantum analogue of the classical Fisher information.[1][2][3][4][5] It is one of the central quantities used to qualify the utility of an input state, especially in Mach–Zehnder (or, equivalently, Ramsey) interferometer-based phase or parameter estimation.[1][3][6] It is shown that the quantum Fisher information can also be a sensitive probe of a quantum phase transition (e.g. recognizing the superradiant quantum phase transition in the Dicke model[6]). The quantum Fisher information F Q [ ϱ , A ] {\displaystyle F_{\rm {Q}}[\varrho ,A]} of a state ϱ {\displaystyle \varrho } with respect to the observable A {\displaystyle A} is defined as

F Q [ ϱ , A ] = 2 k , l ( λ k λ l ) 2 ( λ k + λ l ) | k | A | l | 2 , {\displaystyle F_{\rm {Q}}[\varrho ,A]=2\sum _{k,l}{\frac {(\lambda _{k}-\lambda _{l})^{2}}{(\lambda _{k}+\lambda _{l})}}\vert \langle k\vert A\vert l\rangle \vert ^{2},}

where λ k {\displaystyle \lambda _{k}} and | k {\displaystyle \vert k\rangle } are the eigenvalues and eigenvectors of the density matrix ϱ , {\displaystyle \varrho ,} respectively, and the summation goes over all k {\displaystyle k} and l {\displaystyle l} such that λ k + λ l > 0 {\displaystyle \lambda _{k}+\lambda _{l}>0} .

When the observable generates a unitary transformation of the system with a parameter θ {\displaystyle \theta } from initial state ϱ 0 {\displaystyle \varrho _{0}} ,

ϱ ( θ ) = exp ( i A θ ) ϱ 0 exp ( + i A θ ) , {\displaystyle \varrho (\theta )=\exp(-iA\theta )\varrho _{0}\exp(+iA\theta ),}

the quantum Fisher information constrains the achievable precision in statistical estimation of the parameter θ {\displaystyle \theta } via the quantum Cramér–Rao bound as

( Δ θ ) 2 1 m F Q [ ϱ , A ] , {\displaystyle (\Delta \theta )^{2}\geq {\frac {1}{mF_{\rm {Q}}[\varrho ,A]}},}

where m {\displaystyle m} is the number of independent repetitions.

It is often desirable to estimate the magnitude of an unknown parameter α {\displaystyle \alpha } that controls the strength of a system's Hamiltonian H = α A {\displaystyle H=\alpha A} with respect to a known observable A {\displaystyle A} during a known dynamical time t {\displaystyle t} . In this case, defining θ = α t {\displaystyle \theta =\alpha t} , so that θ A = t H {\displaystyle \theta A=tH} , means estimates of θ {\displaystyle \theta } can be directly translated into estimates of α {\displaystyle \alpha } .

Connection with Fisher information

Classical Fisher information of measuring observable B {\displaystyle B} on density matrix ϱ ( θ ) {\displaystyle \varrho (\theta )} is defined as F [ B , θ ] = b 1 p ( b | θ ) ( p ( b | θ ) θ ) 2 {\displaystyle F[B,\theta ]=\sum _{b}{\frac {1}{p(b|\theta )}}\left({\frac {\partial p(b|\theta )}{\partial \theta }}\right)^{2}} , where p ( b | θ ) = b | ϱ ( θ ) | b {\displaystyle p(b|\theta )=\langle b\vert \varrho (\theta )\vert b\rangle } is the probability of obtaining outcome b {\displaystyle b} when measuring observable B {\displaystyle B} on the transformed density matrix ϱ ( θ ) {\displaystyle \varrho (\theta )} . b {\displaystyle b} is the eigenvalue corresponding to eigenvector | b {\displaystyle \vert b\rangle } of observable B {\displaystyle B} .

Quantum Fisher information is the supremum of the classical Fisher information over all such observables,[7]

F Q [ ϱ , A ] = sup B F [ B , θ ] . {\displaystyle F_{\rm {Q}}[\varrho ,A]=\sup _{B}F[B,\theta ].}

Relation to the symmetric logarithmic derivative

The quantum Fisher information equals the expectation value of L ϱ 2 {\displaystyle L_{\varrho }^{2}} , where L ϱ {\displaystyle L_{\varrho }} is the symmetric logarithmic derivative

Equivalent expressions

For a unitary encoding operation ϱ ( θ ) = exp ( i A θ ) ϱ 0 exp ( + i A θ ) , {\displaystyle \varrho (\theta )=\exp(-iA\theta )\varrho _{0}\exp(+iA\theta ),} , the quantum Fisher information can be computed as an integral,[8]

F Q [ ϱ , A ] = 2 0 tr ( exp ( ρ 0 t ) [ ϱ 0 , A ] exp ( ρ 0 t ) [ ϱ 0 , A ] )   d t , {\displaystyle F_{\rm {Q}}[\varrho ,A]=-2\int _{0}^{\infty }{\text{tr}}\left(\exp(-\rho _{0}t)[\varrho _{0},A]\exp(-\rho _{0}t)[\varrho _{0},A]\right)\ dt,}

where [   ,   ] {\displaystyle [\ ,\ ]} on the right hand side denotes commutator. It can be also expressed in terms of Kronecker product and vectorization,[9]

F Q [ ϱ , A ] = 2 vec ( [ ϱ 0 , A ] ) ( ρ 0 I + I ρ 0 ) 1 vec ( [ ϱ 0 , A ] ) , {\displaystyle F_{\rm {Q}}[\varrho ,A]=2\,{\text{vec}}([\varrho _{0},A])^{\dagger }{\big (}\rho _{0}^{*}\otimes {\rm {I}}+{\rm {I}}\otimes \rho _{0}{\big )}^{-1}{\text{vec}}([\varrho _{0},A]),}

where {\displaystyle ^{*}} denotes complex conjugate, and {\displaystyle ^{\dagger }} denotes conjugate transpose. This formula holds for invertible density matrices. For non-invertible density matrices, the inverse above is substituted by the Moore-Penrose pseudoinverse. Alternatively, one can compute the quantum Fisher information for invertible state ρ ν = ( 1 ν ) ρ 0 + ν π {\displaystyle \rho _{\nu }=(1-\nu )\rho _{0}+\nu \pi } , where π {\displaystyle \pi } is any full-rank density matrix, and then perform the limit ν 0 + {\displaystyle \nu \rightarrow 0^{+}} to obtain the quantum Fisher information for ρ 0 {\displaystyle \rho _{0}} . Density matrix π {\displaystyle \pi } can be, for example, I d e n t i t y / dim H {\displaystyle {\rm {Identity}}/\dim {\mathcal {H}}} in a finite-dimensional system, or a thermal state in infinite dimensional systems.

Generalization and relations to Bures metric and quantum fidelity

For any differentiable parametrization of the density matrix ϱ ( θ ) {\displaystyle \varrho ({\boldsymbol {\theta }})} by a vector of parameters θ = ( θ 1 , , θ n ) {\displaystyle {\boldsymbol {\theta }}=(\theta _{1},\dots ,\theta _{n})} , the quantum Fisher information matrix is defined as

F Q i j [ ϱ ( θ ) ] = 2 k , l Re ( k | i ϱ | l l | j ϱ | k ) λ k + λ l , {\displaystyle F_{\rm {Q}}^{ij}[\varrho ({\boldsymbol {\theta }})]=2\sum _{k,l}{\frac {\operatorname {Re} (\langle k\vert \partial _{i}\varrho \vert l\rangle \langle l\vert \partial _{j}\varrho \vert k\rangle )}{\lambda _{k}+\lambda _{l}}},}

where i {\displaystyle \partial _{i}} denotes partial derivative with respect to parameter θ i {\displaystyle \theta _{i}} . The formula also holds without taking the real part Re {\displaystyle \operatorname {Re} } , because the imaginary part leads to an antisymmetric contribution that disappears under the sum. Note that all eigenvalues λ k {\displaystyle \lambda _{k}} and eigenvectors | k {\displaystyle \vert k\rangle } of the density matrix potentially depend on the vector of parameters θ {\displaystyle {\boldsymbol {\theta }}} .

This definition is identical to four times the Bures metric, up to singular points where the rank of the density matrix changes (those are the points at which λ k + λ l {\displaystyle \lambda _{k}+\lambda _{l}} suddenly becomes zero.) Through this relation, it also connects with quantum fidelity F ( ϱ , σ ) = ( t r [ ϱ σ ϱ ] ) 2 {\displaystyle F(\varrho ,\sigma )=\left(\mathrm {tr} \left[{\sqrt {{\sqrt {\varrho }}\sigma {\sqrt {\varrho }}}}\right]\right)^{2}} of two infinitesimally close states,[10]

F ( ϱ θ , ϱ θ + d θ ) = 1 1 4 i , j ( F Q i j [ ϱ ( θ ) ] + 2 λ k ( θ ) = 0 i j λ k ) d θ i d θ j + O ( d θ 3 ) , {\displaystyle F(\varrho _{\boldsymbol {\theta }},\varrho _{{\boldsymbol {\theta }}+d{\boldsymbol {\theta }}})=1-{\frac {1}{4}}\sum _{i,j}{\Big (}F_{\rm {Q}}^{ij}[\varrho ({\boldsymbol {\theta }})]+2\!\!\sum _{\lambda _{k}({\boldsymbol {\theta }})=0}\!\!\partial _{i}\partial _{j}\lambda _{k}{\Big )}d\theta _{i}d\theta _{j}+{\mathcal {O}}(d\theta ^{3}),}

where the inner sum goes over all k {\displaystyle k} at which eigenvalues λ k ( θ ) = 0 {\displaystyle \lambda _{k}({\boldsymbol {\theta }})=0} . The extra term (which is however zero in most applications) can be avoided by taking a symmetric expansion of fidelity,[11]

F ( ϱ θ d θ / 2 , ϱ θ + d θ / 2 ) = 1 1 4 i , j F Q i j [ ϱ ( θ ) ] d θ i d θ j + O ( d θ 3 ) . {\displaystyle F\left(\varrho _{{\boldsymbol {\theta }}-d{\boldsymbol {\theta }}/2},\varrho _{{\boldsymbol {\theta }}+d{\boldsymbol {\theta }}/2}\right)=1-{\frac {1}{4}}\sum _{i,j}F_{\rm {Q}}^{ij}[\varrho ({\boldsymbol {\theta }})]d\theta _{i}d\theta _{j}+{\mathcal {O}}(d\theta ^{3}).}

For n = 1 {\displaystyle n=1} and unitary encoding, the quantum Fisher information matrix reduces to the original definition.

Quantum Fisher information matrix is a part of a wider family of quantum statistical distances.[12]

Relation to fidelity susceptibility

Assuming that | ψ 0 ( θ ) {\displaystyle \vert \psi _{0}(\theta )\rangle } is a ground state of a parameter-dependent non-degenerate Hamiltonian H ( θ ) {\displaystyle H(\theta )} , four times the quantum Fisher information of this state is called fidelity susceptibility, and denoted[13]

χ F = 4 F Q ( | ψ 0 ( θ ) ) . {\displaystyle \chi _{F}=4F_{Q}(\vert \psi _{0}(\theta )\rangle ).}

Fidelity susceptibility measures the sensitivity of the ground state to the parameter, and its divergence indicates a quantum phase transition. This is because of the aforementioned connection with fidelity: a diverging quantum Fisher information means that | ψ 0 ( θ ) {\displaystyle \vert \psi _{0}(\theta )\rangle } and | ψ 0 ( θ + d θ ) {\displaystyle \vert \psi _{0}(\theta +d\theta )\rangle } are orthogonal to each other, for any infinitesimal change in parameter d θ {\displaystyle d\theta } , and thus are said to undergo a phase-transition at point θ {\displaystyle \theta } .

Convexity properties

The quantum Fisher information equals four times the variance for pure states

F Q [ | Ψ , H ] = 4 ( Δ H ) Ψ 2 {\displaystyle F_{\rm {Q}}[\vert \Psi \rangle ,H]=4(\Delta H)_{\Psi }^{2}} .

For mixed states, when the probabilities are parameter independent, i.e., when p ( θ ) = p {\displaystyle p(\theta )=p} , the quantum Fisher information is convex:

F Q [ p ϱ 1 ( θ ) + ( 1 p ) ϱ 2 ( θ ) , H ] p F Q [ ϱ 1 ( θ ) , H ] + ( 1 p ) F Q [ ϱ 2 ( θ ) , H ] . {\displaystyle F_{\rm {Q}}[p\varrho _{1}(\theta )+(1-p)\varrho _{2}(\theta ),H]\leq pF_{\rm {Q}}[\varrho _{1}(\theta ),H]+(1-p)F_{\rm {Q}}[\varrho _{2}(\theta ),H].}

The quantum Fisher information is the largest function that is convex and that equals four times the variance for pure states. That is, it equals four times the convex roof of the variance[14][15]

F Q [ ϱ , H ] = 4 inf { p k , | Ψ k } k p k ( Δ H ) Ψ k 2 , {\displaystyle F_{\rm {Q}}[\varrho ,H]=4\inf _{\{p_{k},\vert \Psi _{k}\rangle \}}\sum _{k}p_{k}(\Delta H)_{\Psi _{k}}^{2},}

where the infimum is over all decompositions of the density matrix

ϱ = k p k | Ψ k Ψ k | . {\displaystyle \varrho =\sum _{k}p_{k}\vert \Psi _{k}\rangle \langle \Psi _{k}\vert .}

Note that | Ψ k {\displaystyle \vert \Psi _{k}\rangle } are not necessarily orthogonal to each other. The above optimization can be rewritten as an optimization over the two-copy space as [16]

F Q [ ϱ , H ] = min ϱ 12 2 T r [ ( H I d e n t i t y I d e n t i t y H ) 2 ϱ 12 ] , {\displaystyle F_{Q}[\varrho ,H]=\min _{\varrho _{12}}2{\rm {Tr}}[(H\otimes {\rm {Identity}}-{\rm {Identity}}\otimes H)^{2}\varrho _{12}],}

such that ϱ 12 {\displaystyle \varrho _{12}} is a symmetric separable state and

T r 1 ( ϱ 12 ) = T r 2 ( ϱ 12 ) = ϱ . {\displaystyle {\rm {Tr}}_{1}(\varrho _{12})={\rm {Tr}}_{2}(\varrho _{12})=\varrho .}

Later the above statement has been proved even for the case of a minimization over general (not necessarily symmetric) separable states.[17]

When the probabilities are θ {\displaystyle \theta } -dependent, an extended-convexity relation has been proved:[18]

F Q [ i p i ( θ ) ϱ i ( θ ) ] i p i ( θ ) F Q [ ϱ i ( θ ) ] + F C [ { p i ( θ ) } ] , {\displaystyle F_{\rm {Q}}{\Big [}\sum _{i}p_{i}(\theta )\varrho _{i}(\theta ){\Big ]}\leq \sum _{i}p_{i}(\theta )F_{\rm {Q}}[\varrho _{i}(\theta )]+F_{\rm {C}}[\{p_{i}(\theta )\}],}

where F C [ { p i ( θ ) } ] = i θ p i ( θ ) 2 p i ( θ ) {\displaystyle F_{\rm {C}}[\{p_{i}(\theta )\}]=\sum _{i}{\frac {\partial _{\theta }p_{i}(\theta )^{2}}{p_{i}(\theta )}}} is the classical Fisher information associated to the probabilities contributing to the convex decomposition. The first term, in the right hand side of the above inequality, can be considered as the average quantum Fisher information of the density matrices in the convex decomposition.

Inequalities for composite systems

We need to understand the behavior of quantum Fisher information in composite system in order to study quantum metrology of many-particle systems.[19] For product states,

F Q [ ϱ 1 ϱ 2 , H 1 I d e n t i t y + I d e n t i t y H 2 ] = F Q [ ϱ 1 , H 1 ] + F Q [ ϱ 2 , H 2 ] {\displaystyle F_{\rm {Q}}[\varrho _{1}\otimes \varrho _{2},H_{1}\otimes {\rm {Identity}}+{\rm {Identity}}\otimes H_{2}]=F_{\rm {Q}}[\varrho _{1},H_{1}]+F_{\rm {Q}}[\varrho _{2},H_{2}]}

holds.

For the reduced state, we have

F Q [ ϱ 12 , H 1 I d e n t i t y 2 ] F Q [ ϱ 1 , H 1 ] , {\displaystyle F_{\rm {Q}}[\varrho _{12},H_{1}\otimes {\rm {Identity}}_{2}]\geq F_{\rm {Q}}[\varrho _{1},H_{1}],}

where ϱ 1 = T r 2 ( ϱ 12 ) {\displaystyle \varrho _{1}={\rm {Tr}}_{2}(\varrho _{12})} .

Relation to entanglement

There are strong links between quantum metrology and quantum information science. For a multiparticle system of N {\displaystyle N} spin-1/2 particles [20]

F Q [ ϱ , J z ] N {\displaystyle F_{\rm {Q}}[\varrho ,J_{z}]\leq N}

holds for separable states, where

J z = n = 1 N j z ( n ) , {\displaystyle J_{z}=\sum _{n=1}^{N}j_{z}^{(n)},}

and j z ( n ) {\displaystyle j_{z}^{(n)}} is a single particle angular momentum component. The maximum for general quantum states is given by

F Q [ ϱ , J z ] N 2 . {\displaystyle F_{\rm {Q}}[\varrho ,J_{z}]\leq N^{2}.}

Hence, quantum entanglement is needed to reach the maximum precision in quantum metrology. Moreover, for quantum states with an entanglement depth k {\displaystyle k} ,

F Q [ ϱ , J z ] s k 2 + r 2 {\displaystyle F_{\rm {Q}}[\varrho ,J_{z}]\leq sk^{2}+r^{2}}

holds, where s = N / k {\displaystyle s=\lfloor N/k\rfloor } is the largest integer smaller than or equal to N / k , {\displaystyle N/k,} and r = N s k {\displaystyle r=N-sk} is the remainder from dividing N {\displaystyle N} by k {\displaystyle k} . Hence, a higher and higher levels of multipartite entanglement is needed to achieve a better and better accuracy in parameter estimation.[21][22] It is possible to obtain a weaker but simpler bound [23]

F Q [ ϱ , J z ] N k . {\displaystyle F_{\rm {Q}}[\varrho ,J_{z}]\leq Nk.}

Hence, a lower bound on the entanglement depth is obtained as

F Q [ ϱ , J z ] N k . {\displaystyle {\frac {F_{\rm {Q}}[\varrho ,J_{z}]}{N}}\leq k.}

Relation to the Wigner–Yanase skew information

The Wigner–Yanase skew information is defined as [24]

I ( ϱ , H ) = T r ( H 2 ϱ ) T r ( H ϱ H ϱ ) . {\displaystyle I(\varrho ,H)={\rm {Tr}}(H^{2}\varrho )-{\rm {Tr}}(H{\sqrt {\varrho }}H{\sqrt {\varrho }}).}

It follows that I ( ϱ , H ) {\displaystyle I(\varrho ,H)} is convex in ϱ . {\displaystyle \varrho .}

For the quantum Fisher information and the Wigner–Yanase skew information, the inequality

F Q [ ϱ , H ] 4 I ( ϱ , H ) {\displaystyle F_{\rm {Q}}[\varrho ,H]\geq 4I(\varrho ,H)}

holds, where there is an equality for pure states.

Relation to the variance

For any decomposition of the density matrix given by p k {\displaystyle p_{k}} and | Ψ k {\displaystyle \vert \Psi _{k}\rangle } the relation [14]

( Δ H ) 2 k p k ( Δ H ) Ψ k 2 1 4 F Q [ ϱ , H ] {\displaystyle (\Delta H)^{2}\geq \sum _{k}p_{k}(\Delta H)_{\Psi _{k}}^{2}\geq {\frac {1}{4}}F_{\rm {Q}}[\varrho ,H]}

holds, where both inequalities are tight. That is, there is a decomposition for which the second inequality is saturated, which is the same as stating that the quantum Fisher information is the convex roof of the variance over four, discussed above. There is also a decomposition for which the first inequality is saturated, which means that the variance is its own concave roof [14]

( Δ H ) 2 = sup { p k , | Ψ k } k p k ( Δ H ) Ψ k 2 . {\displaystyle (\Delta H)^{2}=\sup _{\{p_{k},\vert \Psi _{k}\rangle \}}\sum _{k}p_{k}(\Delta H)_{\Psi _{k}}^{2}.}

Uncertainty relations with the quantum Fisher information and the variance

Knowing that the quantum Fisher information is the convex roof of the variance times four, we obtain the relation [25] ( Δ A ) 2 F Q [ ϱ , B ] | i [ A , B ] | 2 , {\displaystyle (\Delta A)^{2}F_{Q}[\varrho ,B]\geq \vert \langle i[A,B]\rangle \vert ^{2},} which is stronger than the Heisenberg uncertainty relation. For a particle of spin- j , {\displaystyle j,} the following uncertainty relation holds ( Δ J x ) 2 + ( Δ J y ) 2 + ( Δ J z ) 2 j , {\displaystyle (\Delta J_{x})^{2}+(\Delta J_{y})^{2}+(\Delta J_{z})^{2}\geq j,} where J l {\displaystyle J_{l}} are angular momentum components. The relation can be strengthened as [26][27] ( Δ J x ) 2 + ( Δ J y ) 2 + F Q [ ϱ , J z ] / 4 j . {\displaystyle (\Delta J_{x})^{2}+(\Delta J_{y})^{2}+F_{Q}[\varrho ,J_{z}]/4\geq j.}

References

  1. ^ a b Helstrom, C (1976). Quantum detection and estimation theory. Academic Press. ISBN 0123400503.
  2. ^ Holevo, Alexander S (1982). Probabilistic and statistical aspects of quantum theory (2nd English ed.). Scuola Normale Superiore. ISBN 978-88-7642-378-9.
  3. ^ a b Braunstein, Samuel L.; Caves, Carlton M. (1994-05-30). "Statistical distance and the geometry of quantum states". Physical Review Letters. 72 (22). American Physical Society (APS): 3439–3443. Bibcode:1994PhRvL..72.3439B. doi:10.1103/physrevlett.72.3439. ISSN 0031-9007. PMID 10056200.
  4. ^ Braunstein, Samuel L.; Caves, Carlton M.; Milburn, G.J. (April 1996). "Generalized Uncertainty Relations: Theory, Examples, and Lorentz Invariance". Annals of Physics. 247 (1): 135–173. arXiv:quant-ph/9507004. Bibcode:1996AnPhy.247..135B. doi:10.1006/aphy.1996.0040. S2CID 358923.
  5. ^ Paris, Matteo G. A. (21 November 2011). "Quantum Estimation for Quantum Technology". International Journal of Quantum Information. 07 (supp01): 125–137. arXiv:0804.2981. doi:10.1142/S0219749909004839. S2CID 2365312.
  6. ^ a b Wang, Teng-Long; Wu, Ling-Na; Yang, Wen; Jin, Guang-Ri; Lambert, Neill; Nori, Franco (2014-06-17). "Quantum Fisher information as a signature of the superradiant quantum phase transition". New Journal of Physics. 16 (6): 063039. arXiv:1312.1426. Bibcode:2014NJPh...16f3039W. doi:10.1088/1367-2630/16/6/063039. ISSN 1367-2630.
  7. ^ Paris, Matteo G. A. (2009). "Quantum estimation for quantum technology". International Journal of Quantum Information. 07 (supp01): 125–137. arXiv:0804.2981. doi:10.1142/s0219749909004839. ISSN 0219-7499. S2CID 2365312.
  8. ^ PARIS, MATTEO G. A. (2009). "Quantum estimation for quantum technology". International Journal of Quantum Information. 07 (supp01): 125–137. arXiv:0804.2981. doi:10.1142/s0219749909004839. ISSN 0219-7499. S2CID 2365312.
  9. ^ Šafránek, Dominik (2018-04-12). "Simple expression for the quantum Fisher information matrix". Physical Review A. 97 (4): 042322. arXiv:1801.00945. Bibcode:2018PhRvA..97d2322S. doi:10.1103/physreva.97.042322. ISSN 2469-9926.
  10. ^ Šafránek, Dominik (2017-05-11). "Discontinuities of the quantum Fisher information and the Bures metric". Physical Review A. 95 (5): 052320. arXiv:1612.04581. Bibcode:2017PhRvA..95e2320S. doi:10.1103/physreva.95.052320. ISSN 2469-9926. S2CID 118962619.
  11. ^ Zhou, Sisi; Jiang, Liang (18 Oct 2019). "An exact correspondence between the quantum Fisher information and the Bures metric". arXiv:1910.08473 [quant-ph].
  12. ^ Jarzyna, M.; Kołodyński, J. (18 August 2020). "Geometric Approach to Quantum Statistical Inference". IEEE Journal on Selected Areas in Information Theory. 1 (2): 367–386. arXiv:2008.09129. doi:10.1109/JSAIT.2020.3017469. ISSN 2641-8770. S2CID 221245983.
  13. ^ Gu, S.-J. (2010). "Fidelity approach to quantum phase transitions". International Journal of Modern Physics B. 24 (23): 4371–4458. arXiv:0811.3127. Bibcode:2010IJMPB..24.4371G. doi:10.1142/S0217979210056335. S2CID 118375103.
  14. ^ a b c Tóth, Géza; Petz, Dénes (20 March 2013). "Extremal properties of the variance and the quantum Fisher information". Physical Review A. 87 (3): 032324. arXiv:1109.2831. Bibcode:2013PhRvA..87c2324T. doi:10.1103/PhysRevA.87.032324. S2CID 55088553.
  15. ^ Yu, Sixia (2013). "Quantum Fisher Information as the Convex Roof of Variance". arXiv:1302.5311 [quant-ph].
  16. ^ Tóth, Géza; Moroder, Tobias; Gühne, Otfried (21 April 2015). "Evaluating Convex Roof Entanglement Measures". Physical Review Letters. 114 (16): 160501. arXiv:1409.3806. Bibcode:2015PhRvL.114p0501T. doi:10.1103/PhysRevLett.114.160501. PMID 25955038. S2CID 39578286.
  17. ^ Tóth, Géza; Pitrik, József (16 October 2023). "Quantum Wasserstein distance based on an optimization over separable states". Quantum. 7: 1143. arXiv:2209.09925. Bibcode:2023Quant...7.1143T. doi:10.22331/q-2023-10-16-1143. S2CID 252408568.
  18. ^ Alipour, S.; Rezakhani, A. T. (2015-04-07). "Extended convexity of quantum Fisher information in quantum metrology". Physical Review A. 91 (4): 042104. arXiv:1403.8033. Bibcode:2015PhRvA..91d2104A. doi:10.1103/PhysRevA.91.042104. ISSN 1050-2947. S2CID 124094775.
  19. ^ Tóth, Géza; Apellaniz, Iagoba (24 October 2014). "Quantum metrology from a quantum information science perspective". Journal of Physics A: Mathematical and Theoretical. 47 (42): 424006. arXiv:1405.4878. Bibcode:2014JPhA...47P4006T. doi:10.1088/1751-8113/47/42/424006. S2CID 119261375.
  20. ^ Pezzé, Luca; Smerzi, Augusto (10 March 2009). "Entanglement, Nonlinear Dynamics, and the Heisenberg Limit". Physical Review Letters. 102 (10): 100401. arXiv:0711.4840. Bibcode:2009PhRvL.102j0401P. doi:10.1103/PhysRevLett.102.100401. PMID 19392092. S2CID 13095638.
  21. ^ Hyllus, Philipp (2012). "Fisher information and multiparticle entanglement". Physical Review A. 85 (2): 022321. arXiv:1006.4366. Bibcode:2012PhRvA..85b2321H. doi:10.1103/physreva.85.022321. S2CID 118652590.
  22. ^ Tóth, Géza (2012). "Multipartite entanglement and high-precision metrology". Physical Review A. 85 (2): 022322. arXiv:1006.4368. Bibcode:2012PhRvA..85b2322T. doi:10.1103/physreva.85.022322. S2CID 119110009.
  23. ^ Tóth, Géza (2021). Entanglement detection and quantum metrology in quantum optical systems (PDF). Budapest: Doctoral Dissertation submitted to the Hungarian Academy of Sciences. p. 68.
  24. ^ Wigner, E. P.; Yanase, M. M. (1 June 1963). "Information Contents of Distributions". Proceedings of the National Academy of Sciences. 49 (6): 910–918. Bibcode:1963PNAS...49..910W. doi:10.1073/pnas.49.6.910. PMC 300031. PMID 16591109.
  25. ^ Fröwis, Florian; Schmied, Roman; Gisin, Nicolas (2 July 2015). "Tighter quantum uncertainty relations following from a general probabilistic bound". Physical Review A. 92 (1): 012102. arXiv:1409.4440. Bibcode:2015PhRvA..92a2102F. doi:10.1103/PhysRevA.92.012102. S2CID 58912643.
  26. ^ Tóth, Géza; Fröwis, Florian (31 January 2022). "Uncertainty relations with the variance and the quantum Fisher information based on convex decompositions of density matrices". Physical Review Research. 4 (1): 013075. arXiv:2109.06893. Bibcode:2022PhRvR...4a3075T. doi:10.1103/PhysRevResearch.4.013075. S2CID 237513549.
  27. ^ Chiew, Shao-Hen; Gessner, Manuel (31 January 2022). "Improving sum uncertainty relations with the quantum Fisher information". Physical Review Research. 4 (1): 013076. arXiv:2109.06900. Bibcode:2022PhRvR...4a3076C. doi:10.1103/PhysRevResearch.4.013076. S2CID 237513883.