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

Unfolding of relative gg-entropies and monotone metrics

F. Di Nocera1,2[Uncaptioned image],
1 Max Planck Institute for Mathematics in the Sciences, Leipzig, Germany
2
e-mail: fabiodncr[at]gmail.com and dinocer[at]mis.mpg.de
Abstract

We discuss the geometric aspects of a recently described unfolding procedure and show the form of objects relevant in the field of Quantum Information Geometry in the unfolding space. In particular, we show the form of the quantum monotone metric tensors characterized by Petz and retrace in this unfolded perspective a recently introduced procedure of extracting a covariant tensor from a relative gg-entropy.

1 Introduction

Fisher-Rao metric tensor [21] on classical statistical models can be seen as a “second-order expansion” of Kullback-Leibler relative entropy between points in the model. A more general fact actually holds, namely that it is possible to extract covariant tensors from second-order derivatives of relative entropies [2] satisfying the monotonicity property with respect to Markov morphisms [11]. This is in perfect agreement with a crucial result obtained by Cencov [6], i.e. the characterization of the Fisher-Rao metric tensor as the only metric tensor which is invariant under congruent embeddings.

In the quantum case, however, the situation is more complex. In [20], (developing on the work of Cencov and Morozova in [19]) Petz showed that there is a family of quantum generalizations of Fisher-Rao metric tensor labelled by operator monotone functions [17]. This characterization is largely used in the context of Quantum Information Geometry and was obtained considering monotonicity under the class of completely positive, trace preserving (CPTP) maps, that are the appropriate generalization to the quantum case of Markov’s maps [7].

In [16], the authors show that the characterization obtained by Petz can be recovered by considering some “second-order expansion” of a family of relative entropies labelled by a convex operator function [15] gg defined on density matrices, called relative g-entropies.

Our main goal is to retrace these basic facts about Quantum Information Geometry in an unfolded framework described in [9, 18, 8]. The rough idea behind the unfolding procedure is to specify a density operator ρ\rho by means of its spectrum and a unitary operator UU that diagonalizes ρ\rho. Then since density matrices are positive semi-definite, trace-one matrices, their spectrum is specified (up to permutations) by a probability distribution over a finite sample space. In the language of Information Geometry, by points on a simplex. This idea will be made clear in section 2, while in section 3 we will show the aforementioned results from [20] and [16] in this framework.

2 Unfolding of the space of quantum states

Let \mathcal{H} be a finite-dimensional Hilbert space of complex dimension nn, let also ()\mathcal{B}(\mathcal{H}) be the space of (bounded) operators acting on the Hilbert space \mathcal{H}, the (real) subspace of ()\mathcal{B}(\mathcal{H}) of self-adjoint operators will be denoted as sa()\mathcal{B}_{sa}(\mathcal{H}). A mixed quantum state is defined as a positive semi-definite (hence self-adjoint) operator with trace one. If a mixed state is also invertible, i.e. is positive definite, it is said to be faithful. The space of all such operators is called space of faithful quantum states,

𝒮()={ρsa()|ρ>0,Trρ=1}.\mathscr{S}(\mathcal{H})=\{\,\rho\in\mathcal{B}_{sa}(\mathcal{H})\,|\quad\rho>0,\quad\textit{Tr}{\,\rho}=1\}. (1)

For space constraint, here we will not discuss in detail the geometry of the space of quantum states, and refer the reader to [13, 3, 4]. The space of faithful quantum states is a smooth manifold and our discussion will be restricted to this setting because the monotone quantum metric tensors classified by Petz are well-defined only on this manifold. However, it is worth mentioning that the whole space of quantum states 𝒮()¯\overline{\mathscr{S}(\mathcal{H})} (i.e., the topological closure of 𝒮()\mathscr{S}(\mathcal{H}) in sa()\mathcal{B}_{sa}(\mathcal{H})) is a stratified manifold, and that its extremal points (i.e., the pure states) form a smooth manifold diffeomorphic to ()\mathbb{CP}(\mathcal{H}) on which the Riemannian structure is fiexd to be the Fubini-Study metric tensor because of the requirement of unitary invariance [12].

For self-adjoint operators holds a crucial result, known as spectral theorem, that states that for any self-adjoint operator Asa()A\in\mathcal{B}_{sa}(\mathcal{H}) there exist a orthonormal basis of \mathcal{H} consisting of eigenvectors of AA. This means that we can always find a basis in which the self-adjoint operator is diagonal. Moreover, the eigenvalues are real numbers.

Change of basis on an Hilbert space \mathcal{H} are performed by means of an action of unitary operators on \mathcal{H}. An operator U()U\in\mathcal{B}(\mathcal{H}) is said to be unitary if UU=UU=𝕀UU^{\dagger}=U^{\dagger}U=\mathbb{I}. For these reasons, we can write any mixed state ρ\rho with eigenvalues {λ1,λ2,,λn}\{\lambda_{1},\lambda_{2},\dots,\lambda_{n}\} as

ρ=Uρ0U,\rho=U\rho_{0}U^{\dagger}, (2)

with ρ0\rho_{0} given, with respect to some reference basis in \mathcal{H}, by

ρ0=[λ1000λ2000λn].\rho_{0}=\begin{bmatrix}\lambda_{1}&0&\dots&0\\ 0&\lambda_{2}&\dots&0\\ \vdots&\vdots&\ddots&\vdots\\ 0&0&\dots&\lambda_{n}\end{bmatrix}. (3)

Notice that the positive definiteness of ρ0\rho_{0} implies that all λj\lambda_{j}’s are strictly positive, while the condition on the trace implies that their sum is 1. This means that, roughly speaking, the eigenvalues of ρ\rho are the components of a probability distribution on a sample space of cardinality nn. We now want to start from this observation to introduce our unfolding procedure, in order to do so, let us make this analogy between probability vectors and diagonal states more formal.

The set of probability distributions on a finite sample space of cardinality nn can be geometrically represented as the n1n-1-dimensional standard simplex Δn1\Delta_{n-1}. This can be realized as a subset of n\mathbb{R}^{n} as

Δn1={𝐩=(p1,p2,,pn)|jpj=1,pj0j=1,2,,n}.\Delta_{n-1}=\{\mathbf{p}=(p_{1},p_{2},\dots,p_{n})\in\mathbb{R}\,|\quad\sum_{j}p_{j}=1,\quad p_{j}\geq 0\quad\forall j=1,2,\dots,n\}. (4)

We now want to look at probability distributions as diagonal mixed states, basically identifying the components of the vector 𝐩\mathbf{p} with the eigenvalues of some mixed state. Since we want to restrict our discussion to faithful states, we need to consider only the cases in which the components of 𝐩\mathbf{p} are strictly positive, meaning that we have to consider the space

Δn1o={𝐩=(p1,p2,,pn)|jpj=1,pj>0j=1,2,,n}.\Delta^{\mathrm{o}}_{n-1}=\{\mathbf{p}=(p_{1},p_{2},\dots,p_{n})\in\mathbb{R}\,|\quad\sum_{j}p_{j}=1,\quad p_{j}>0\quad\forall j=1,2,\dots,n\}. (5)

In this way, we can immerse Δn1o\Delta^{\mathrm{o}}_{n-1} in the space of faithful quantum states as follows,

i:Δn1o𝐩=(p1,p2,,pn)i(𝐩)=[p1000p2000pn]𝒮().i:\Delta^{\mathrm{o}}_{n-1}\ni\mathbf{p}=(p_{1},p_{2},\dots,p_{n})\mapsto i(\mathbf{p})=\begin{bmatrix}p_{1}&0&\dots&0\\ 0&p_{2}&\dots&0\\ \vdots&\vdots&\ddots&\vdots\\ 0&0&\dots&p_{n}\end{bmatrix}\in\mathscr{S}(\mathcal{H}). (6)
Remark 1

Let us stress here that in the definition of ii there is an implicit choice of a basis in ()\mathcal{B}(\mathcal{H}). In fact, choosing to immerse the probability distribution diagonally amounts to choosing a basis {|e1,|e2,,|en}\{\ket{e_{1}},\ket{e_{2}},\dots,\ket{e_{n}}\} of \mathcal{H} and write

i(𝐩)=j=1npj|ejej|.i(\mathbf{p})=\sum_{j=1}^{n}p^{j}\ket{e_{j}}\bra{e_{j}}. (7)

Clearly this implicit choice does not lead to any loss of generality of the discussion.

Any diagonal faithful state can be realized as the image via ii of some probability distribution, implying that for any faithful state ρ\rho there exists a 𝐩Δn1o\mathbf{p}\in\Delta^{\mathrm{o}}_{n-1} and a unitary operator UU such that

ρ=Ui(𝐩)U,\rho=U\,i(\mathbf{p})U^{\dagger}, (8)

but it is easy to see that this doesn’t specify uniquely neither 𝐩\mathbf{p} nor UU. In fact, the action of the unitary group considered has a non.trivial isotropy group, so that UU is not uniquely determined. On the other hand 𝐩\mathbf{p} is determined only up to permutations, the redundancy on 𝐩\mathbf{p} can be removed by considering equivalence classes of probability distributions with respect to permutations, meaning going to the so called Weyl chamber.

Remark 2

Let us notice here that while the degeneracy on the choice of the probability distribution can be solved by taking the quotient with respect to the permutation group, the same can not be done for 𝒰()\mathcal{U}(\mathcal{H}). In fact, it would be possible to quotient with respect to the isotropy group of the considered action if it did not depend on the point i(𝐩)i(\mathbf{p}) on which it acts, which is not the case for our action.

Equation (8) gives a many-to-one correspondence between elements in the cartesian product Δn1o×𝒰()\Delta^{\mathrm{o}}_{n-1}\times\mathcal{U}(\mathcal{H}), i.e. couples (U,𝐩)(U,\mathbf{p}), and faithful states. This is exactly what we want to achieve with our unfolding procedure, ():=Δn1o×𝒰()\mathcal{M}(\mathcal{H}):=\Delta^{\mathrm{o}}_{n-1}\times\mathcal{U}(\mathcal{H}) will be our unfolding space, and we use equation (8) to define the map

π:𝒰()×Δn1o(U,𝐩)π(U,𝐩)=Ui(𝐩)U𝒮().\pi:\mathcal{U}(\mathcal{H})\times\Delta^{\mathrm{o}}_{n-1}\ni(U,\mathbf{p})\mapsto\pi(U,\mathbf{p})=U\,i(\mathbf{p})U^{\dagger}\in\mathscr{S}(\mathcal{H}). (9)

The map π\pi can be seen to be a smooth projection of ()\mathcal{M(H)} on the space of faithful states [8].

Like for all product spaces, it is possible to project \mathcal{M} on its two factors, in particular, let us consider the projection on the second factor,

πD:𝒰()×Δn1o(U,𝐩)𝐩Δn1o.\pi_{D}:\mathcal{U}(\mathcal{H})\times\Delta^{\mathrm{o}}_{n-1}\ni(U,\mathbf{p})\mapsto\mathbf{p}\in\Delta^{\mathrm{o}}_{n-1}. (10)

The projection πD\pi_{D} will be useful for recognizing classical structures in the unfolded space, in fact, it can be regarded as some sort of “dequantization map”, taking a mixed quantum state and giving as its image a classical mixture with weights given by the eigenvalues of the mixed state; basically forgetting about the non-classical features of the system.

Since we want to discuss Information Geometry on the unfolding space, a brief discussion regarding the tangent spaces to the unfolding space ()\mathcal{M}(\mathcal{H}) is unavoidable. Since ()\mathcal{M}(\mathcal{H}) is the product of two manifolds, we can consider vectors in T(U,𝐩)()T_{(U,\mathbf{p})}\mathcal{M}(\mathcal{H}) as couples of vectors tangent to 𝒰()\mathcal{U}(\mathcal{H}) and Δn1o\Delta^{\mathrm{o}}_{n-1} respectively [1]. Vectors tangent to 𝒰()\mathcal{U}(\mathcal{H}) are given by skew-adjoint matrices [14], while it is easy to see that the tangent space of Δn1o\Delta^{\mathrm{o}}_{n-1} is isomorphic to n1\mathbb{R}^{n-1} and given by vectors in n\mathbb{R}^{n} whose components sum to zero.

This means that every V(U,𝐩)T(U,𝐩)()V_{(U,\mathbf{p})}\in T_{(U,\mathbf{p})}\mathcal{M}(\mathcal{H}) can be written as

V(U,𝐩)=(iH,v)V_{(U,\mathbf{p})}=(iH,v) (11)

with HH a self-adjoint matrix and vnv\in\mathbb{R}^{n} such that jvj=0\sum_{j}v^{j}=0.

3 Monotone metrics and relative gg-entropies on the unfolding space

Petz’s characterization of monotone quantum metrics is used in the context of Quantum Information Theory when considering the possible quantum analogues of Fisher-Rao metric tensor. In [20], Petz characterized all metric tensors on the space of quantum states obeying monotonicity under completely positive, trace-preserving maps [20] as those metric whose action on the vectors A,BTρ𝒮()A,B\in T_{\rho}\mathscr{S}(\mathcal{H}) can be written as

Gf|ρ(A,B)=Tr(A(𝐊ρf)1(B)).G^{f}|_{\rho}(A,B)=\textit{Tr}\left(A\left(\mathbf{K}^{f}_{\rho}\right)^{-1}(B)\right). (12)

Here 𝐊ρf\mathbf{K}^{f}_{\rho} is a superoperator depending on the point ρ\rho and on the operator monotone function ff. The superoperator 𝐊f\mathbf{K}_{f} is defined as

𝐊ρf=f(𝐋ρ𝐑ρ1)𝐑ρ,\mathbf{K}^{f}_{\rho}=f(\mathbf{L}_{\rho}\mathbf{R}_{\rho^{-1}})\mathbf{R}_{\rho}, (13)

where 𝐋ρ\mathbf{L}_{\rho} and 𝐑ρ\mathbf{R}_{\rho} are respectively the left and right multiplications by means of ρ\rho, namely

𝐋ρ(B)=ρB,𝐑ρ(B)=Bρ.\mathbf{L}_{\rho}(B)=\rho B,\quad\mathbf{R}_{\rho}(B)=B\rho. (14)

Moreover, if the function ff is such that f(x)=xf(x1)f(x)=xf(x^{-1}) and f(1)=1f(1)=1, the correspondence between such functions and monotone metrics is one-to-one.

Making use of the projection map introduced in (9) we can pull-back the monotone metrics in (12) to the unfolding space. In fact, we can define a twice covariant tensor GfG^{f}_{\mathcal{M}} on ()\mathcal{M(H)} via

Gf:=πGf,G^{f}_{\mathcal{M}}:=\pi^{*}G^{f}, (15)

this means that the action of GfG^{f}_{\mathcal{M}} on two tangent vectors can be written as

(Gf)(U,𝐩)((iH,v),(iK,u)):=(πGf)(U,𝐩)((iH,v),(iK,u))=(Gf)π(U,𝐩)(T(U,𝐩)π(iH,v),T(U,𝐩)π(iK,u)),\begin{split}\left(G^{f}_{\mathcal{M}}\right)_{(U,\mathbf{p})}\left((iH,v),(iK,u)\right):&=\left(\pi^{*}G^{f}\right)_{(U,\mathbf{p})}\left((iH,v),(iK,u)\right)\\ &=(G^{f})_{\pi(U,\mathbf{p})}\left(T_{(U,\mathbf{p})}\pi(iH,v),T_{(U,\mathbf{p})}\pi(iK,u)\right),\end{split} (16)

where T(U,𝐩)πT_{(U,\mathbf{p})}\pi denotes the tangent map of π\pi at the point (U,𝐩)(U,\mathbf{p}), HH and KK are self-adjoint operators on \mathcal{H} and u,vT𝐩Δn1ou,v\in T_{\mathbf{p}}\Delta^{\mathrm{o}}_{n-1}.

Remark 3

A crucial thing to notice is that the tensor GfG^{f}_{\mathcal{M}} is not a metric tensor, since it has a non-trivial kernel. To see this, consider a curve m(t)=(U(t),𝐩)m(t)=(U(t),\mathbf{p}) on ()\mathcal{M}(\mathcal{H}) such that 𝐩\mathbf{p} does not depend on tt, while U(t)U(t) is in the isotropy group of π(m(0))\pi(m(0)) for all tt, then holds

π(m(t))=π(m(0)):=ρ~t,\pi(m(t))=\pi(m(0)):=\tilde{\rho}\quad\forall t, (17)

implying that for any vector VπV_{\pi} tangent to such curve we have

Tm(0)π(Vπ)=𝟎,T_{m(0)}\pi(V_{\pi})=\mathbf{0}, (18)

where 𝟎\mathbf{0} is the null vector in Tρ~𝒮()T_{\tilde{\rho}}\mathscr{S}(\mathcal{H}). This implies

(Gf)m(0)(Vπ,Vπ)=(Gf)ρ~(𝟎,𝟎)=0.\left(G^{f}_{\mathcal{M}}\right)_{m(0)}\left(V_{\pi},V_{\pi}\right)=(G^{f})_{\tilde{\rho}}\left(\mathbf{0},\mathbf{0}\right)=0. (19)

In [9], the authors find that the tensor GfG^{f}_{\mathcal{M}} can be written as

(Gf)(U,𝐩)((iH,v),(iK,u))=G𝒰f(iH,iK)+πD(GFRn1(v,u)).\left(G^{f}_{\mathcal{M}}\right)_{(U,\mathbf{p})}\left((iH,v),(iK,u)\right)=G^{f}_{\mathcal{U}}(iH,iK)+\pi_{D}^{*}\left(G^{n-1}_{FR}(v,u)\right). (20)

Let us comment on this result in some detail: here GFRn1G_{FR}^{n-1} is the Fisher-Rao metric tensor defined on Δn1o\Delta^{\mathrm{o}}_{n-1}, πD\pi_{D} is the dequantization map defined as in (10), so that the last term of this expression can be considered as some sort of classical term of the metric GfG^{f}_{\mathcal{M}}. Moreover, the term containing GFRn1G^{n-1}_{FR} does not depend on the choice of the operator monotone function ff, meaning that this is a feature of all metrics that can be obtained via the unfolding procedure from the one introduced by Petz. The term G𝒰f(iH,iK)G^{f}_{\mathcal{U}}(iH,iK), on the other hand, depends on the choice of the operator monotone function ff and it cannot be seen as the pull-back of a tensor defined on 𝒰()\mathcal{U}(\mathcal{H}), since it depends also on 𝐩\mathbf{p}. Nonetheless, it can be shown that this term is equal to zero whenever HH and KK commute with π(U,𝐩)\pi(U,\mathbf{p}).

This shows how Classical Information Geometry can in a sense always be considered as contained in Quantum Information Geometry. More specifically, whenever one considers vectors in T(U,𝐩)T_{(U,\mathbf{p})}\mathcal{M} commuting with π(U,𝐩)\pi(U,\mathbf{p}) one recovers exactly the only tensor that is relevant in Classical Information Geometry.

As already recalled in the introduction, it is possible to regard the Fisher-Rao metric as a “second-order expansion” of Kullback-Leibler relative entropy. Now, adapting the results in [16] to our framework, we want to show that if we pull-back relative gg-entropies from the space of quantum states to the unfolding space and take some sort of “second-order expansion” of it, we recover equation (12), i.e. the unfolded version of Petz’s characterization of monotone metrics.

Relative gg-entropies are defined as two-point functions on the space of quantum states,

Hg:𝒮()×𝒮()(ρ,σ)Sg(ρ,σ)=Tr(ρg(𝐋σ𝐑ρ1)(ρ)),H_{g}:\mathscr{S}(\mathcal{H})\times\mathscr{S}(\mathcal{H})\ni(\rho,\sigma)\mapsto S_{g}(\rho,\sigma)=\textit{Tr}\left(\sqrt{\rho}g(\mathbf{L}_{\sigma}\mathbf{R}_{\rho^{-1}})(\sqrt{\rho})\right)\in\mathbb{R}, (21)

where gg is an operator convex function. It turns out that HgH_{g} satisfies the monotonicity condition

Hg(ρ,σ)=Hg𝒦(Φ(ρ),Φ(σ))H^{\mathcal{H}}_{g}(\rho,\sigma)=H^{\mathcal{K}}_{g}(\Phi(\rho),\Phi(\sigma)) (22)

for all Hilbert spaces \mathcal{H} and 𝒦\mathcal{K} and for all CPTP maps Φ:()(𝒦)\Phi:\mathcal{B}(\mathcal{H})\rightarrow\mathcal{B(K)} [10] Let now (U,𝐩)(U,\mathbf{p}) and (V,𝐪)(V,\mathbf{q}) be two points in ()\mathcal{M}(\mathcal{H}) such that

ρ=π(U,𝐩),σ=π(V,𝐪),\begin{split}\rho=\pi(U,\mathbf{p}),\\ \sigma=\pi(V,\mathbf{q}),\end{split} (23)

In order to pull-back two-point functions we need to define a new projection map that extends the projection map π\pi to the cartesian product of ()\mathcal{M(H)} with itself, i.e.

Π:()×()((U,𝐩),(V,𝐪))(π(U,𝐩),π(V,𝐪))𝒮()×𝒮().\Pi:\mathcal{M(H)}\times\mathcal{M(H)}\ni((U,\mathbf{p}),(V,\mathbf{q}))\mapsto(\pi(U,\mathbf{p}),\pi(V,\mathbf{q}))\in\mathscr{S}(\mathcal{H})\times\mathscr{S}(\mathcal{H}). (24)

In this way ΠSg\Pi^{*}S^{\mathcal{H}}_{g} is a two-point function on ()\mathcal{M(H)}, a short computation shows that

Π(Sg)((U,𝐩),(V,𝐪))=j,kg(qjpk)ek|UV|ejej|VU|ek.\Pi^{*}\left(S^{\mathcal{H}}_{g}\right)\left((U,\mathbf{p}),(V,\mathbf{q})\right)=\sum_{j,k}g\left(\frac{q^{j}}{p_{k}}\right)\bra{e_{k}}U^{\dagger}V\ket{e_{j}}\bra{e_{j}}V^{\dagger}U\ket{e_{k}}. (25)

Clearly the second derivatives will now be carried out using the tools of differential geometry and in particular, for the part containing the unitary operators, making use of the differential calculus on Lie groups. Here we will not retrace the computations and refer the reader to [9] for details. This gives rise to the expression

Gg(X,Y)=g′′(1)πrGFRn1(X,Y)+g(1)j=1kpjej|UdU(X)UdU(Y)+UdU(Y)UdU(X)|ej2g(1)j=1npjej|UdU(X)|ejej|UdU(Y)|ej2k>j(g(pjpk)pk+g(pkpj)pj)(ek|UdU(X)|ejej|UdU(Y)|ek).\begin{split}G^{\mathcal{M}}_{g}(X,Y)&=g^{\prime\prime}(1)\,\pi_{r}^{*}G_{FR}^{n-1}(X,Y)\\ &+g\left(1\right)\sum_{j=1}^{k}p^{j}\bra{e_{j}}U^{\dagger}\mathrm{d}U(X)U^{\dagger}\mathrm{d}U(Y)+U^{\dagger}\mathrm{d}U(Y)\,U^{\dagger}\mathrm{d}U(X)\ket{e_{j}}\\ &-2g\left(1\right)\sum_{j=1}^{n}p^{j}\,\bra{e_{j}}U^{\dagger}\mathrm{d}U(X)\ket{e_{j}}\bra{e_{j}}U^{\dagger}\mathrm{d}U(Y)\ket{e_{j}}\\ &-2\sum_{k>j}\left(g\left(\frac{p^{j}}{p^{k}}\right)p^{k}+g\left(\frac{p^{k}}{p^{j}}\right)p^{j}\right)\,\,\Re\left(\bra{e_{k}}U^{\dagger}\mathrm{d}U(X)\ket{e_{j}}\bra{e_{j}}U^{\dagger}\mathrm{d}U(Y)\ket{e_{k}}\right).\end{split} (26)

Where XX and YY are in T(U,𝐩)()T_{(U,\mathbf{p})}\mathcal{M(H)} and UdUU^{\dagger}\mathrm{d}U is the Maurer-Cartan one-form [5] on 𝒰()\mathcal{U}(\mathcal{H}).

It can be shown that this expression coincides with (12) iff g(1)=0g(1)=0 and

f(x)=(1x)2g(x)+xg(x1).f(x)=\frac{(1-x)^{2}}{g(x)+xg(x^{-1})}. (27)

holds. Using the last expression it is also easily seen that g′′(1)=f(1)=1g^{\prime\prime}(1)=f(1)=1. Equation (27) provides a relation between the operator convex function gg in (22) and the operator monotone function ff in (12).

This result allows to simplify significantly the expression in (26),

Gg(X,Y)=πDGFRn1(X,Y)2k>j(g(pjpk)pk+g(pkpj)pj)(ek|UdU(X)|ejej|UdU(Y)|ek),\begin{split}G^{\mathcal{M}}_{g}(X,Y)&=\pi_{D}^{*}G_{FR}^{n-1}(X,Y)\\ &-2\sum_{k>j}\left(g\left(\frac{p^{j}}{p^{k}}\right)p^{k}+g\left(\frac{p^{k}}{p^{j}}\right)p^{j}\right)\,\,\Re\left(\bra{e_{k}}U^{\dagger}\mathrm{d}U(X)\ket{e_{j}}\bra{e_{j}}U^{\dagger}\mathrm{d}U(Y)\ket{e_{k}}\right),\end{split} (28)

plugging (27) in the last expression we get

Gg=πDGFRn12k>j(pkpj)2(pkf(pjpk))1(ek|UdU|ejej|UdU|ek).\begin{split}G^{\mathcal{M}}_{g}=\pi_{D}^{*}G_{FR}^{n-1}-2\sum_{k>j}\left(p_{k}-p_{j}\right)^{2}\left(p_{k}\,f\left(\frac{p^{j}}{p^{k}}\right)\right)^{-1}\,\,\Re\left(\bra{e_{k}}U^{\dagger}\mathrm{d}U\ket{e_{j}}\otimes\bra{e_{j}}U^{\dagger}\mathrm{d}U\ket{e_{k}}\right).\end{split} (29)

This expression is in complete agreement with (12), for a complete proof of this statement see [9]. As anticipated, we can see that the expression exhibits a splitting in two terms, one containing Fisher-Rao metric tensor and the other is a term that doesn’t appear in the classical case and it is a two-contravariant tensor that depends on the choice of the operator monotone function ff and it is not the pull-back of some tensor defined on the unitary group 𝒰()\mathcal{U}(\mathcal{H}). This can be seen from the fact that also the second terms depends on the values of the components of 𝐩\mathbf{p}.

4 Conclusions

We showed how the shift from the state of quantum states 𝒮()\mathcal{S(H)} to the unfolding space ()\mathcal{M(H)} makes it strikingly easy to recognize at first glance the classical structures underlying the quantum world. Stated differently, some sort of dequantization process becomes possible by simply forgetting about the first argument in the couple (U,𝒑)(U,\boldsymbol{p}). In fact, we showed the expression that the metric tensors characterized by Petz assume in this framework, and we saw that in the unfolding space the metric tensors split in two terms. One contains information about the classical aspects and, as one would expect, is exactly Fisher-Rao metric. The other one is a term that doesn’t appear in the classical case and it is a two-contravariant tensor depending on the choice of the operator function ff and written containing the contribution given by the action of the unitary group.

Finally we recalled the procedure of obtaining a monotone quantum metric from a relative gg-entropy in this unfolded perspective. Again the results are in agreement both with the original result in [16] and with the unfolded version of Petz’s characterization. With this approach we also saw how the “non-classical term’ ’is a two-contravariant tensor written in terms of the Maurer-Cartan one form on 𝒰()\mathcal{U}(\mathcal{H}). This shows that, apart from the conceptual idea behind the unfolding, this can also be exploited for practical uses. In fact, this allows to set the whole discussion on a manifold which is the product of two manifolds whose geometry is well-studied. In particular, ideas from the differential calculus on Lie groups can be applied to perform calculations on 𝒰()\mathcal{U}(\mathcal{H}).

It is also true that one can obtain dually-related connection from third order derivative of relative gg-entropies. A possible next step would be to compute the family of dually-related connections determined by the relative g-entropies in this unfolded perspective. What one would expect is that also the skewness tensor describing this dual structure will split into a purely classical part and a quantum part given in terms of elements of the unitary group.

Another possible follow-up for this work would be to try to apply the unfolding procedure to the infinite-dimensional case. In this case the interior of the simplex will be replaced by the subset of the Banach space l1()l^{1}(\mathbb{R}) given by strictly positive sequences adding to one. It is clear that such a generalization brings all the technicalities related to infinite-dimensional analysis and geometry, but at the same time it is immediate to see that such a generalization would allow to apply this procedure to relevant quantum mechanical problem such as the harmonic oscillator or the hydrogen atom.

References

  • [1] R. Abraham, J. E. Marsden, and T. Ratiu. Manifolds, tensor analysis, and applications. Springer-Verlag, New York, third edition, 2012.
  • [2] S. I. Amari. Information Geometry and its Application. Springer, Japan, 2016.
  • [3] A. Ashtekar and T. A. Schilling. Geometrical formulation of quantum mechanics. In A. Harvey, editor, On Einstein’s Path: Essays in Honor of Engelbert Schucking, pages 23 – 65. Springer-Verlag, New York, 1999.
  • [4] I. Bengtsson and K. Życzkowski. Geometry of Quantum States: An Introduction to Quantum Entanglement. Cambridge University Press, New York, 2006. DOI: 10.1017/cbo9780511535048.
  • [5] Élie Cartan. Sur la structure des groupes infinis de transformation. In Annales scientifiques de l’École normale supérieure, volume 21, pages 153–206, 1904.
  • [6] N. N. Cencov. Statistical Decision Rules and Optimal Inference. American Mathematical Society, Providence, RI, 1982.
  • [7] M. Choi. Completely positive linear maps on complex matrices. Linear Algebra and its Applications, 10(3):285–290, 1975. DOI: 10.1016/0024-3795(75)90075-0.
  • [8] F. M. Ciaglia, F. Di Cosmo, M. Laudato, G. Marmo, G. Mele, F. Ventriglia, and P. Vitale. A Pedagogical Intrinsic Approach to Relative Entropies as Potential Functions of Quantum Metrics: the q-z family. Annals of Physics, 395:238–274, 2018. DOI: 10.1016/j.aop.2018.05.015.
  • [9] Florio M Ciaglia, Fabio Di Cosmo, Fabio Di Nocera, and Patrizia Vitale. Monotone metric tensors in quantum information geometry. arXiv preprint arXiv:2203.10857, 2022.
  • [10] Joel E Cohen, Yoh Iwasa, Gh Rautu, Mary Beth Ruskai, Eugene Seneta, and Gh Zbaganu. Relative entropy under mappings by stochastic matrices. Linear algebra and its applications, 179:211–235, 1993.
  • [11] I. Csizár. Eine informationstheoretische Ungleichung und ihre Anwendung auf den Beweis der Ergodizitat von Markoffschen Ketten. A Magyar Tudományos Akadémia Matematikaiés Fizikai Tudományok Osztályának Közleményei, 8:85–108, 1963.
  • [12] F. D’Andrea and D. Franco. On the pseudo-manifold of quantum states. Differential Geometry and its Applications, 78:101800, 2021. DOI: 10.1016/j.difgeo.2021.101800.
  • [13] J. Grabowski, M. Kuś, and G. Marmo. Symmetries, group actions, and entanglement. Open Systems & Information Dynamics, 13(04):343 – 362, 2006.
  • [14] Brian C Hall. Lie groups, lie algebras, and representations. In Quantum Theory for Mathematicians, pages 333–366. Springer, 2013.
  • [15] Fritz Kraus. Über konvexe matrixfunktionen. Mathematische Zeitschrift, 41(1):18–42, 1936.
  • [16] A. Lesniewski and M. B. Ruskai. Monotone riemannian metrics and relative entropy on noncommutative probability spaces. Journal of Mathematical Physics, 40(11):5702 – 5724, 1999.
  • [17] Karl Löwner. Über monotone matrixfunktionen. Mathematische Zeitschrift, 38(1):177–216, 1934.
  • [18] V. I. Man’ko, G. Marmo, F. Ventriglia, and P. Vitale. Metric on the space of quantum states from relative entropy. Tomographic reconstruction. Journal of Physics A: Mathematical and Theorerical, 50(33):335302, 2017. DOI: 10.1088/1751-8121/aa7d7d.
  • [19] E. A. Morozowa and N. N. Cencov. Markov invariant geometry on state manifolds. Journal of Soviet Mathematics, 56(5):2648–2669, 1991. DOI: 10.1007/BF01095975.
  • [20] D. Petz. Monotone metrics on matrix spaces. Linear Algebra and its Applications, 244:81–96, 1996. DOI: 10.1016/0024-3795(94)00211-8.
  • [21] C. R. Rao. Information and accuracy attainable in the estimation of statistical parameters. Bulletin of the Calcutta Mathematical Society, 37(3):81–91, 1945.