Skip to main content

ORIGINAL RESEARCH article

Front. Phys., 07 February 2023
Sec. Statistical and Computational Physics
This article is part of the Research Topic Advances in Information Geometry: Beyond the Conventional Approach View all 4 articles

Exponential arcs in manifolds of quantum states

  • Physics Department, Universiteit Antwerpen, Antwerp, Belgium

The manifold under consideration consists of the faithful normal states on a sigma-finite von Neumann algebra in standard form. Tangent planes and approximate tangent planes are discussed. A relative entropy/divergence function is assumed to be given. It is used to generalize the notion of an exponential arc connecting one state to another. The generator of the exponential arc is shown to be unique up to an additive constant. In the case of Araki’s relative entropy, every self-adjoint element of the von Neumann algebra generates an exponential arc. The generators of the composed exponential arcs are shown to add up. The metric derived from Araki’s relative entropy is shown to reproduce the Kubo–Mori metric. The latter is the metric used in linear response theory. The e- and m-connections describe a dual pair of geometries. Any finite number of linearly independent generators determines a submanifold of states connected to a given reference state by an exponential arc. Such a submanifold is a quantum generalization of a dually flat statistical manifold.

1 Introduction

The goal of the present paper is to show that the theory of quantum statistical manifolds can be formulated without reference to density matrices. It is tradition to describe the statistical state of a quantum model by a density matrix. In many cases this suffices, in particular when the Hilbert space of wave functions is finite-dimensional. However, even simple models such as the quantum harmonic oscillator or the hydrogen atom require an infinite-dimensional Hilbert space. This involves handling of unbounded operators which cause considerable technical complications. These complications are avoided in the present work.

A one-to-one correspondence between density matrices and quantum states is usually accepted. The quantum states form the sample space of the statistical description. An alternative description emerged in the past century, which introduced the notion of a mathematical state on an algebra of observables which can be realized as an algebra of bounded operators on Hilbert space. See for instance [15].

Equilibrium states of quantum statistical mechanics are described by the quantum analogue of the probability distribution of Gibbs, which is a density matrix ρ of the form

ρ=1ZeβH,

with H a Hermitian matrix, β a parameter the inverse temperature, and Z a function of β used to normalize density matrix ρ so that its trace equals 1. Models described in this way can belong to a quantum exponential family. They possess an intriguing property called the Kubo–Martin–Schwinger (KMS) condition [6]. The KMS condition describes a symmetry property of the time evolution of quantum states. This symmetry coincides with the symmetry between left and right multiplication of operators, which is studied in the Tomita–Takesaki theory [7]. [5] can be used as a reference text for this theory.

The notion of a statistical manifold is studied in information geometry ([812]). It is a manifold of probability distributions. The quantum analogue is described in Chapter 7 of [11] as a manifold of k by k density matrices. The book of Petz [13] reviews several aspects of quantum statistics, including the basics of quantum information and quantum information geometry.

The generalization of Amari’s dually flat geometry from statistical models with a finite number of parameters to Banach manifolds of mutually equivalent probability measures started with the work of [14]. Non-commutative versions were formulated by [1519].

The convex set M of faithful normal states on a σ-finite von Neumann algebra is in general not a Banach manifold. The point of view taken in the present work is that the set M should, by definition, be a quantum statistical manifold. This raises the question of how to transfer common notions of differential geometry and of Banach manifolds to this quantum setting. The present work contributes to this effort.

The relative entropy of Umegaki [20] is the starting point to implement Amari’s dually flat geometry on the quantum manifold. It should be noted that relative entropy is called a divergence function in mathematical literature. Araki [2123] generalizes Umegaki’s relative entropy to the context of mathematical states on an algebra of bounded operators on a Hilbert space. The use of Araki’s relative entropy replacing that of Umegaki’s is the core of the present work.

Exponential arcs were introduced in [24, 25] and used in [26]. These arcs can be considered one-parameter exponential families embedded in the manifold. The maximal exponential model centered at a given probability distribution p equals the set of all probability distributions connected to p by an open exponential arc. Exponential arcs were studied in the quantum setting by [27]. Here, the definition is generalized. The exponential arcs are used to define quantum statistical manifolds as submanifolds of the manifold of all quantum states.

The Radon–Nikodym Theorem plays an important role in probability theory. For each measure absolutely continuous with respect to the reference measure, there exists an essentially unique probability distribution function. The problem that arises in the non-commutative context is the non-uniqueness of the Radon–Nikodym derivative. This leads to different definitions of the relative entropy and of the exponential arcs. First attempts to reformulate the theory of the quantum statistical manifold in terms of states on a C*-algebra are found in [28,29] and in [27]. These two approaches differ in the choice of the Radon–Nikodym derivative. In the present work, the definition of an exponential arc is generalized so that it depends explicitly on the choice of relative entropy and in that way on the choice of the Radon–Nikodym derivative.

The alternative approach of [30] relies on the Lie Theory for the group of bounded operators with bounded inverse. The state space is partitioned into the disjoint union of the orbits of an action of the Lie group. Under mild conditions, it is shown that the orbits are Banach manifolds. The restriction to bounded operators implies that the orbits do not connect quasi-equivalent states when the Radon–Nikodym derivatives are unbounded operators.

Sections 2–4 give a short introduction on KMS states, on the theory of the modular operator, and on positive cones. Section 5 gives a definition of the manifold M under study as the convex set of faithful normal states on a sigma-finite von Neumann algebra. The tangent space consists of linear functionals on the algebra. Its extent depends on the chosen topology, and it is not obvious how to find a good compromise. Therefore, the notion of approximate tangent vectors is considered in Section 6.

A dense subset of the manifold M consists of states majorized by a multiple of the reference state. This subset of states is mentioned in Section 7 because it is easier to handle.

Section 8 gives a new definition of exponential arcs. It generalizes existing concepts and is broad enough to cover different approaches. The definition depends on the choice of a relative entropy/divergence function. Such an exponential arc can be seen as a one-dimensional sub-manifold and as a straightforward example of a quantum statistical manifold. Duality properties well-known for models of information geometry are elaborated in Section 9.

The important example of the algebra of n-by-n matrices is considered in Section 10.

Starting with Section 11 the paper specializes to the case of Araki’s relative entropy. It is shown in Section 13 that each self-adjoint element h of the von Neumann algebra defines an exponential arc defined relative to Araki’s relative entropy and starting at the reference state ω. The initial derivative of the arc exists as a Fréchet derivative and belongs to the tangent plane TωM. The inner product between two such tangent vectors reproduces the metric which is used in the Kubo–Mori Theory of linear response. This is shown in Section 14. The exponential arcs are shown to be geodesics for the e-connection which is, by definition, the dual of the m-connection.

Section 16 applies the results obtained so far to show that manifolds generated by a finite number of exponential arcs have the properties one expects from a quantum statistical manifold.

A few points of concern are discussed in the final Section 17.

2 KMS states

Equilibrium states of quantum statistical mechanics satisfy the KMS condition. In the GNS representation, an equilibrium state becomes a faithful state on a σ-finite von Neumann algebra of operators on a complex Hilbert space. The state is defined by a normalized cyclic and separating vector in the Hilbert space.

The state of a model of statistical physics can be described by a mathematical state on a C*-algebra A. It can be represented by a normalized vector Ω (a wave function) in a Hilbert space H. This is known as the GNS (Gelfand–Naimark–Segal) representation theorem. Observable quantities are represented by self-adjoint operators on H. The quantum expectation ⟨x⟩ of operator x is then given by

x=xΩ,Ω,(1)

with in the right-hand side the scalar product of the two vectors xΩ and Ω. It should be noted that the mathematical convention is followed that the scalar product (inner product) is linear in its first argument and conjugate-linear in the second argument. In Dirac’s bra-ket notation, it reads

x=Ω|xΩ.

For convenience, one works with a von Neumann algebra M of bounded operators on the Hilbert space H. Observables of interest, when unbounded, are represented by operators affiliated with M. The state ω on the C*-algebra extends to a vector state on M again denoted ω. It is given by

ωx=xΩ,Ω,xM.

The vector Ω is cyclic for M, which means that the subspace MΩ is dense in the Hilbert space H. It is also assumed in what follows that the state ω is faithful, i.e., ω(x*x) = 0 implies x = 0. This implies that Ω is a separating vector for M, i.e., xΩ = 0 implies x = 0 for any x in M, and hence it is a cyclic vector for the commutant M of M, the algebra of all operators commuting with all of M.

Equilibrium states of statistical mechanics are characterized by the KMS (Kubo–Martin–Schwinger) condition [6]. Roughly speaking, this condition states that the quantum time evolution of the model has an analytic extension into the complex plane. This is made more precise in what follows.

The time evolution is described by a strongly continuous one-parameter group tRut of unitary operators which leave the algebra M unchanged, i.e., xM implies that xt=ut*xut belongs to M for all t. The operators ut are determined by a self-adjoint operator H

ut=eitH,

which is the generator of the time evolution in the GNS representation. The time derivative of xt satisfies

iddtxt=xt,H.

This equation has the same form as Heisenberg’s equation of motion.

The KMS condition requires that for any pair x, y of operators in M, there exists a complex function F(w), defined and continuous on the strip −β ≤ Im w ≤ 0 and analytical inside with boundary values

Ft=xtyΩ,Ω and Ftiβ=yxtΩ,Ω,tR.

In the mathematics literature, the parameter β, which is the inverse temperature of the model, is usually taken equal to 1 or -1.

An immediate consequence of the KMS condition being satisfied is that the state ω is invariant. Indeed, take y equal to the identity operator. Then, one has F (t) = F(t) for all t in R. If in addition, x is self-adjoint, then F(t) is a real function. From the Schwarz reflection principle, one then concludes that F(w) is a constant function. This implies ω(xt) = ω(x) for all self-adjoint x and hence for all x. The GNS theorem then guarantees that the vector Ω can be taken to be invariant, i.e., utΩ = Ω for all t.

3 The modular operator

The quantum-mechanical time evolution coincides with the modular automorphism group of Tomita–Takesaki theory.

The KMS condition, when satisfied, expresses a symmetry which is present in the context of non-commuting operators. The symmetry is the inversion of the order of multiplication of operators. In non-commutative groups, the modular function links left and right Haar measures. The analogue in functional analysis is studied in the theory of the modular operator, also called the Tomita–Takesaki theory [7].

The operator eβH with H the generator of the quantum-time evolution is traditionally denoted as ΔΩ. It is the modular operator of the Tomita–Takesaki theory. It is in general an unbounded operator such that MΩ is in the domain of the definition of the square root ΔΩ1/2 of ΔΩ. Hence, the expression

Fw=xΔΩiw/βyΩ,Ω,x,yM,(2)

is well-defined for 0 ≥ Im w ≥−β/2. The other half of the strip 0 ≥ Im w ≥−β is covered by the Schwarz reflection principle. Indeed, if x and y are self-adjoint, then one can show with the Tomita–Takesaki theory that the map tF (t/2) is a real function. Hence, the principle can be applied to obtain F(w)=F(wiβ)̄.

The unitary time evolution operator ut can be written as

ut=ΔΩit/β.

The time evolution of an operator x in the Heisenberg picture is then given by

xt=τtΩx=ΔΩit/βxΔΩit/β.

The action tτtΩ of the group R,+ is called the modular automorphism group.

The modular conjugation operator J of the Tomita–Takesaki Theory represents the symmetry which is at the basis of the theory. It is a conjugate-linear operator satisfying J = J* and J2=I. Operator x belongs to the von Neumann algebra M if and only if JxJ belongs to the commutant algebra M. The latter is the space of operators commuting with all operators in M. The product JΔΩ1/2 is denoted as SΩ and has the property of

SΩxΩ=x*Ω,xM.

4 Dual cones

The natural positive cone PΩ is needed in subsequent sections. One reason for making use of it is that there exists a one-to-one correspondence between normal states on M and normalized vectors in PΩ.

Section 4 of [22]introduces the cones VΩα, 0 ≤ α ≤ 1/2, of the vectors in H. The self-dual cone VΩ1/4 is called the natural positive cone and is denoted as PΩ.

By definition, VΩα is the closure of the cone

ΔΩαxΩ:xM,x0.

The cone VΩ1/2 is used in [27] to introduce exponential arcs. It is equal to the closure of the set

yΩ:yM,y0.

To see this note that

ΔΩ1/2xΩ=JSΩxΩ=Jx*Ω=yΩ

with y = Jx*J. The latter is an arbitrary element of the commutant M.

The following characterization of the natural positive cone PΩ is found in Section 2.5 of [5].

Proposition 1: The cone PΩ=VΩ1/4 equals the closure of the set of vectors

xJxΩ:xM.(3)

This result can be understood as follows. Take Φ in P of the form (3), i.e., Φ = xJxΩ with x in M. Let

y=ΔΩ1/4xΔΩ1/4.(4)

This expression can be inverted to

x=ΔΩ1/4yΔΩ1/4

so that

Φ=xJxΩ=ΔΩ1/4yΔΩ1/4JΔΩ1/4yΔΩ1/4Ω=ΔΩ1/4yJΔΩ1/2yΩ=ΔΩ1/4ySΩyΩ=ΔΩ1/4yy*Ω.

Assume now that one could prove that the operator y defined by (4) belongs to M; then, the above calculation would show that Φ is of the form Φ=ΔΩ1/4aΩ with a = yy* a positive element of M. The actual proof of the proposition uses that τtΩx=ΔΩitxΔΩit belongs to M.

The cone PΩ is independent [22] of the choice of the cyclic and separating vector Ω in PΩ, and the isometry J is the same for all these choices. For this reason, it is said to be universal.

From (3), it is easy to see that each vector in PΩ is an eigenvector with eigenvalue 1 of the modular conjugation operator J. Indeed, one has

xJxΩ=xJxJΩ=JxJxΩ=JxJxΩ.

Here, use is made of JΩ = Ω and the fact that the operators x and JxJ commute with each other.

5 A manifold of quantum states

A manifold M of vector states on the von Neumann algebra M is defined. Tangent vector fields are Fréchet derivatives of paths in M.

Introduce the notation ωΦ for the vector state defined by the normalized vector Φ in H. It is given by

ωΦx=xΦ,Φ,xM.

A manifold M of states on the von Neumann algebra M is defined by

M=ωΦ:ΦPΩ, normalized, cyclic and separating for M.

The equilibrium state ω = ωΩ is taken as a reference point in M. The subset PΩ of H is the natural positive cone introduced in the previous section.

The topology on the manifold M is that of the operator norm. One has

ωΦωΨ=sup|ωΦxωΨx|:xM,x1.

Several topologies can be defined on the algebra M. Particularly relevant is the σ-weak topology. For what follows, it is important to know that in the present context, a state ω on M is said to be normal if and only if it is σ-weakly continuous and if and only if it is a vector state. See for instance, Theorems 2.4.21 and 2.5.31 of [5].

Any tangent vector is a σ-weakly continuous linear functional on the von Neumann algebra M. Let tγt be a Fréchet differentiable map defined on an open interval of R with values in the manifold M. The derivative

γ̇t=ddtγt

is required to exist as a Fréchet derivative, i.e., it satisfies

γtγstsγ̇t=ots.

From the normalization, γt (1) = 1 for all t in the domain of the map, one obtains γ̇t(1)=0. From γt(x*)=γt(x)̄, one obtains γ̇t(x*)=γ̇t(x)̄. Hence, the linear functional γ̇t is Hermitian.

There are several ways to define the tangent space TωM at the point ω in M. Intuitively, a tangent vector is a derivative, defined in some sense, of a path tγt in M passing through the point ω. The states of the manifold M belong to the space M* of all σ-weakly continuous linear functionals on the algebra M (see Proposition 2.4.18 of [5]). Hence, one expects that tangent vectors belong to M* as well.

In this section, the requirement is made that the path tγt is Fréchet-differentiable. This may be too restrictive. In what follows, we adopt the definition that the tangent space TωM consists of all Hermitian χ in M* , satisfying χ(1) = 0. It should be noted that it is well-possible that for certain elements χ of this space, there is no smooth curve passing through ω with the property that the derivative at ω equals χ.

6 Approximate tangents

Approximate tangent vectors can be defined in an intrinsic manner.

An alternative definition of the tangent space starts from the following observation.

Proposition 2: The set Tω defined by

Tω=λϕψ:ϕ,ψM,λR and ω=12ϕ+ψ.

is a linear subspace of the tangent space TωM.

Proof:

Let ϕ and ψ be two states in M such that ω=12(ϕ+ψ). Construct a Fréchet-differentiable path γ by

γt=1tψ+tϕ,t0,1.

The state γt belongs to the manifold M because the latter is a convex set. In particular, one has ω = γ1/2 and ϕψ=γ̇1/2 is a tangent vector. This shows that ϕψ and hence also λ(ϕψ) belongs to TωM. One concludes that TωTωM.

Assume now that λ(ϕψ) and λ′(ϕ′ − ψ′) both belong to Tω. We have to show that

λϕψ+λϕψ

belongs to Tω. If λ = 0 or λ′ = 0, then the claim is clearly satisfied. Without restriction, assume λ = 1.

If λ′ > 0, then choose

ϕ=11+λϕ+λϕ and ψ=11+λψ+λψ.

Both ϕ and ψ belong to M because the latter is a convex set. One verifies that ϕ + ψ = 2ω and

1+λϕψ=ϕψ+λϕψ.

This shows that the latter sum belongs to Tω.

In the case that λ′ < 0, one chooses

ϕ=11λϕλψ and ψ=11λψλϕ

to reach the same conclusion. This finishes the proof that Tω is a linear subspace of TωM.

We introduce the notations

Rω,ϵ=Tϕ with ϕM such that ϕω<ϵ.

and

Tωapprox=ϵ>0Rω,ϵ̄.

The construction of Tωapprox is analogous to the construction of the approximate tangent space in Chapter 3 of [31]. Clearly, TωTωapprox. Further properties are derived below.

Proposition 3: If γ is a Fréchet-differentiable path in M, then γ̇t belongs to Tωapprox with ω = γt.

Proof:

Let γ be a Fréchet-differentiable path in M. Without restriction of generality, assume that γ0 = ω. For any ϵ > 0 and δ > 0, there exists t ≠ 0 such that

γt+γt/2γ0<ϵ.

and

γtγt/2tγ̇0<δ.(5)

Then, ϕ defined by

ϕ=12γt+γt

satisfies ‖ϕω‖ < ϵ, and (γtγt)/2t belongs to Rω,ϵ. Hence, (5) shows that the tangent vector γ̇0 belongs to the closure of Rω,ϵ. Because ϵ > 0 is arbitrary, it also belongs to the intersection, which is Tωapprox.

Lemma 1: Rω,ϵ is a linear subspace of TωM.

Proof:

Take χ and ξ in Rω,ϵ. There exist ϕ and ψ in M such that χTϕ and ξTψ with ‖ϕω‖ < ϵ and ‖ψω‖ < ϵ. Therefore, there exist real λ, μ and states ϕ1, Φ2, Ψ1, Ψ2 in M such that

χ=λϕ1ϕ2 and ϕ=12ϕ1+ϕ2

and

ξ=μψ1ψ2 and ψ=12ψ1+ψ2.

If λ = 0 or μ = 0, then χ + ξ belongs to Rω,ϵ without further argument. Assume, therefore, that λ ≠ 0 and μ ≠ 0. If λμ > 0, then χ + ξ belongs to Tπ, with π = (1 − α)ϕ + αψ and α given by

α=μλ+μ.

Indeed, let

π1=1αϕ1+αψ1,π2=1αϕ2+αψ2.

Then both π1 and π2 belong to M and satisfy

π1+π2=21αϕ+2αψ=2π

and

λ+μπ1π2=λ+μ1αλχ+αμξ=χ+ξ.

In addition,

πω=1αϕω+αψω1αϕω+αψω<ϵ.

One concludes that in this case, χ + ξ belongs to Rω,ϵ.

The case that λμ < 0 is similar. That χRω,ϵ implies λχRω,ϵ is straightforward. One can conclude that Rω,ϵ is a linear space. It clearly is a subspace of TωM.

Proposition 4: Tωapprox is a closed linear subspace of TωM.

Proof:

The lemma shows that Rω,ϵ is a linear subspace of TωM, which is a space closed in norm. Hence, also the norm closure of Rω,ϵ is a subset of this space and therefore also of TωM.

7 Majorized states

The subset of states majorized by a multiple of the reference state ω is considered.

Definition 1: A state ϕ on M is said to be majorized by a multiple of the state ω if there exists a positive constant λ such that

ϕx*xλωx*x for all xM.

Take a′ ≠ 0 in the commutant algebra M and let

Φ=1aΩaΩ.

Then, the state ωΦ is majorized by a multiple of the state ω. Indeed, one has for any positive x in M

ωΦx=xaΩ,aΩaΩ,aΩ=a*ax1/2Ω,x1/2ΩaΩ,aΩa*aaΩ,aΩωx.

It is well-known that all states majorized by a multiple of the state ω are obtained in this way. This is the content of the following proposition.

Proposition 5: If the vector state ωΦ is majorized by a multiple of the state ω, then there exists a unique element a′ of the commutant M such that Φ = a′Ω.

Proof:

An operator a′ is densely defined by

axΩ=xΦ,xM.

It satisfies a′Ω = Φ. It is well-defined because xΩ = 0 implies

xΦ2=ωΦx*xconstant ωx*x=constant xΩ2=0

so that xΦ = 0.

The operator a′ is bounded because

axΩ2=ϕx*xconstant ωx*x=constant xΩ2.

The operator a′ commutes with any x in M because

axyΩ=xyΦ=xayΩ=xayΩ

and Ω is cyclic for M.

The operator a′ is unique. Indeed, assuming b′ in M satisfies Φ = b′Ω. Then, one has for all x in M

0=xabΩ=abxΩ.

Hence, a′ − b′ vanishes on MΩ which is dense in the Hilbert space because Ω is cyclic for M. Because a′ − b′ is a bounded and hence continuous operator, it vanishes everywhere so that a′ = b′.

Item (8) of Theorem 3 of [22] implies the following.

Proposition 6: If a vector state ωΦ, defined by a vector Φ in the natural positive cone PΩ, is dominated by a multiple of the state ω, then there exists a unique element a in the algebra M such that Φ = aΩ and

ωΦx=ωa*xa,xM.

Proof:

Proposition 5 shows that a′ in the commutant M exists such that Φ = a′Ω. Because Φ and Ω both belong to PΩ, one has Φ = JΦ = JaJΩ.

Let a = JaJ. From JMJ=M, it follows that a belongs to M. This shows the existence.

The element a is unique because the correspondence between vector states on M and vectors in PΩ is one-to-one and Ω is a separating vector for M.

If M is a commutative algebra, then a*a is the Radon–Nikodym derivative of the state ωΦ with respect to the reference state ω.

The subset of states of M majorized by a multiple of the state ω is dense in M in the sense that for any state ϕ in M, there exists a sequence (an)n of elements of M with the property that anΩ is a Cauchy sequence and

ϕx=limnωan*xan,xM.

See Propositions 1.5 and 2.5 of [32].

Proposition 7: A tangent vector χ belongs to the subspace Tω of the tangent space TωM if and only if it is proportional to the difference of two states ϕ and ψ in M, both majorized by a multiple of the state ω.

Proof:

If χ belongs to Tω, then by definition, there exist states ϕ and ψ in M such that χ = λ(ϕψ) and ϕ + ψ = 2ω. The latter implies that both ϕ and ψ are majorized by 2ω.

Conversely, assume that ϕ and ψ in M are both majorized by a multiple of the state ω and let χ = λ(ϕψ). This implies the existence of μ ≥ 1 and ν ≥ 1 such that ϕμω and ψνω.

Without restriction, assume that λ > 0.

Introduce

ϕ=ω+ρχ and ψ=ωρχ

with ρ still to be chosen. By construction, it holds that ϕ′ + ψ′ = 2ω and ϕ′ − ψ′ = 2ρχ. Hence, if ϕ′ and ψ′ are states in M and ρ ≠ 0, then one can conclude that χ belongs to Tω.

From

χx*xλϕx*xλμωx*x

and

χx*xλψx*xλνωx*x,

one obtains

ϕx*x1ρλνωx*x,ψx*x1ρλμωx*x.

Let ρ be equal to the inverse of the maximum of λμ and λν to prove the positivity of the functionals ϕ′ and ψ′. Normalization ϕ′(1) = ψ′(1) = 1 follows from χ(1) = 0. The functions are σ-weakly continuous as well. Hence, they are states in M. This ends the proof that χ belongs to Tω.

8 Exponential arcs

[27] introduces the notion of an exponential arc in the Hilbert space, inspired by the notion of exponential arcs in probability space as introduced by [24, 25]. Here, a definition is given which depends on the choice of a relative entropy.

In the present context, a divergence function D (ϕψ) is a real function of two states ϕ and ψ in the manifold M. It cannot be negative, and it vanishes if and only if the two arguments are equal. A value of + is allowed. An energy function is an affine function h defined on a convex subset of the set of normal states on the algebra M.

The following definition of an exponential arc in the manifold M assumes that a divergence function D (ϕψ) is given.

Definition 2: An exponential arc γ is a path in the manifold

t0,1γtM

for which there exists an energy function h such that

γt is in the domain of h;

• The divergence D (γsγt) between any two points of the arc is finite;

• For any state ψ in the domain of h, one has

Dψγt=Dψγ0+Dγ0γt+thγ0hψ,0t1.(6)

The energy function h is the generator of the exponential arc. The arc is said to connect the state γ1 to the state γ0.

A subclass of energy functions is formed by the functions h for which there exists a self-adjoint operator h in the von Neumann algebra M so that

hψ=ψh,ψM.(7)

In such a case, h is called the generator as well. The exponential arcs defined in [27] agree with the above definition with a generator defined by an unbounded operator affiliated with the commutant algebra M.

Proposition 8: Expression (6) implies

Dγsγ0+Dγ0γs=shγshγ0.(8)

and

Dψγt=Dψγs+Dγsγt+tshγshψ.(9)

It should be noted that with s = 0, expression (9) reduces to (6).

Proof:

Take ψ = γs in (6) to find

Dγsγt=Dγsγ0+Dγ0γt+thγ0hγs,0s,t1.(10)

In particular, with s = t, this implies (8).

To prove (9), use (10) to write the right-hand side as

r.h.s.=Dψγs+Dγsγ0+Dγ0γt+thγ0hψshγshψ.

Next, eliminate D (γ0γt) and D (ψγs) with the help of (6). This gives

r.h.s.=Dψγs+Dγsγ0+DψγtDψγ0shγshψ=Dγsγ0+Dψγt+Dγ0γs+shγ0hγs=Dψγt.

To obtain the last line, use (8).

Corollary 1: If tγt is an exponential arc with generator h that connects γ1 to γ0, then for any s, t in [0, 1], the map ϵγ(1−ϵ)s+ϵt is an exponential arc with generator (ts)h that connects γt to γs.

Corollary 2: If tγt is an exponential arc with generator h that connects γ1 to γ0, then tγ1−t is an exponential arc with generator h, connecting the state γ0 to the state γ1.

The following two propositions deal with the uniqueness of an exponential arc and of its generator.

Proposition 9: Let ω and ϕ be two states in M. Fix an energy function h. There is at most one exponential arc tγt with generator h that connects ϕ to ω.

Proof:

Assume both tγt and tδt are exponential arcs connecting the state ϕ to the state ω. Subtract (6) from the same expression with γt replaced by δt and take s = 0. This gives

DψδtDψγt=DωδtDωγt.(11)

Take ψ equal to δt. Then, one obtains

0Dδtγt=DωδtDωγt.

On the other hand, with ψ = γt, one obtains

0Dγtδt=DωδtDωγt.

The two expressions together yield

DωδtDωγt=0.

This implies D (γtδt) = 0. By the basic property of a divergence, one concludes that γt = δt.

Proposition 10: If the exponential arc tγt has two generators h and k, then these generators differ by a constant on their common domain of definition.

Proof:

It follows from (6) that

hγshψ=kγskψ,s0,1(12)

for all states ψ in the intersection of the domains of h and k. This implies that a constant c exists so that

kψ=hψ+c

for all ψ in the common domain.

The requirement (6) is a stability condition. The generator h is a perturbation which shifts the state γ0 to the state ψ. This interpretation will become clear further on. The effect on the relative entropy of the shift along the arc tγt is linear. In the standard case, the relative entropy is based on the logarithmic function. This justifies calling the path tγt an exponential arc.

It should be noted that the Pythagorean relation [33, 34]

Dψγt=Dψγs+Dγsγt

is satisfied for all ψ with the same energy as the state γs, i.e., with

hψ=hγs.

If the divergence function is interpreted as the square of a pseudo-distance, then the aforementioned relation states that for an arbitrary state ψ, the point γs of the arc which has the same energy is the point with minimal distance.

9 The scalar potential

The exponential arc has a dual structure similar to that found in information geometry [10, 11].

Given an exponential arc tγt with generator h, introduce the potential Φγ defined by

Φγt=Dγ0γt+thγ0.

Its Legendre transform is given by

Φγ*α=supαtΦγt:0t1.

Proposition 11: For any exponential arc tγt with generator h, one has

(a) The function th(γt) is strictly increasing;

(b) Φγ(t)=Φγ(s)+D(γsγt)+(ts)h(γs);

(c) The line tΦγ(s)+(ts)h(γs) is tangent to the potential Φγ at the point t = s; this implies that the potential Φγ(s) is a strictly convex function, continuous on the open interval (0, 1);

(d) The following identity holds:

Φγs+Φγ*hγs=shγs,s0,1.

Proof:

(a) Take ψ = γt in (6). This gives

0=Dγtγs+Dγsγt+tshγshγt.

Because divergences cannot be negative, this implies that th(γt) is non-decreasing. Assume now that h(γs)=h(γt). Then, it follows that

0=Dγtγs=Dγsγt.

The latter implies that s = t. One concludes that s < t implies a strict inequality h(γs)<h(γt).

(b) From the definition of the exponential arc, one obtains

Dγsγt+tshγs=DψγtDψγs+tshψ.

Take ψ = γ0 in this expression to find

Dγsγt+tshγs=Dγ0γtDγ0γs+tshγ0=ΦγtΦγs.

(c) From (b), one obtains

ΦγtthγsΦγsshγs,0t1(13)

because D (γsγt) ≥ 0 with equality if and only if s = t. This implies that tΦγ(s)+(ts)h(γs) is a line tangent to the potential Φγ(s). By (a), the slope of this line is a strictly increasing function of s. Hence, the potential Φγ(s) is a strictly convex function, continuous on the open interval (0, 1).

(d) (13) implies that

Φγ*hγs=suptthγsΦγtshγsΦγs.

On the other hand, one can use (b) to obtain

Φγs+Φγ*hγsΦγs+thγsΦγt=thγsDγsγt+tshγs=Dγsγt+shγs.

The optimal choice t = s yields the lower bound sh(γs).

A dual parameter η of the exponential arc γ, dual to the parameter t, is the value h(γt) of the generator h. By item (a) of the proposition, it is a strictly increasing function of t. It is almost equal everywhere to the derivative Φ̇γ(t) of the value of the potential along the path.

10 The matrix case

If ρ and σ are two density matrices, then the obvious definition of an exponential arc connecting σ to ρ is tσt=explogρ+tlogσlogρζtwith normalization ζ(t) given by ζt=log Tr explogρ+tlogσlogρ.It is shown below that the corresponding states given by ϕtx= Tr σtx,xAform an exponential arc for the relative entropy of Umegaki [20] in the GNS-representation of the state σ0.

Fix a non-degenerate density matrix ρ of size n-by-n. It is a positive-definite matrix with trace Tr ρ equal to 1.

Umegaki’s relative entropy for the pair of density matrices σ, τ is given by

Dστ= Tr σlogσlogτ.

Assume now a map

tσt=explogρ+thζt(14)

with normalization ζ(t) and with h given by

h=logσlogρ.

This is the obvious definition of an exponential arc in terms of density matrices. The corresponding potential is

Φσt=Dσ0σt+thσ0=ζt

with

hτ= Tr τh= Tr τlogσlogρ.

The map (14) is also an exponential arc in the sense of Definition 2. To see this, consider any density matrix τ and calculate

DτσtDτσsDσsσt= Tr τlogτlogσt Tr τlogτlogσs Tr σslogσslogσt=ts Tr τσsh=tshγshτ.

This is of the form (6) except that the relative entropy is expressed in terms of density matrices in M instead of vector states in the GNS representation of the state defined by the density matrix ρ.

An explicit construction of the GNS representation is possible. See for instance, the appendix of [28]. Let ω = σ0 denote the state determined by the density matrix ρ

ωA= Tr ρA

for any n-by-n matrix A with entries in C. Such a matrix A is represented on the Hilbert space H=CnCn by the operator AI, where I is the n-by-n identity matrix. The von Neumann algebra M is the space of operators AI.

The matrix ρ can be diagonalized. This gives the spectral representation

ρ=ipiei,

where (ei)i is an orthonormal basis in Cn. Let

Ω=ipieiei.

It is a normalized vector in H. One readily verifies that

ωA=AIΩ,Ω

for any n-by-n matrix A. In this way, any density matrix ρ defines a vector Ω in H. The vector Ω is cyclic and separating for M if ρ is non-degenerate. Hence, there is a one-to-one correspondence between non-degenerate density matrices and states in the manifold M. It is then straightforward to replace the density matrices by states in the expressions obtained in the first part of this section.

11 The relative modular operator

Araki [35] introduces the relative modular operator ΔΦ,Ψ for any pair of vectors Φ and Ψ in the natural positive cone P.

Assume that Φ and Ψ are vectors in P which are separating for the algebra M. Then, a conjugate–linear operator is defined by

xΨx*Φ,xM.

It is well-defined because by assumption, xΨ = 0 implies that x = 0 so that also x*Φ = 0. It is a closable operator. Indeed, assume the sequence xnΨ converges to 0. Then, one has for any y in the commutant M that

xnΨ,yΦ=yΨ,xn*Φ

converges to 0. By assumption, Ψ is separating for M so that it is cyclic for the commutant M. Hence, if the sequence xn*Φ converges, then it converges to 0. This shows the closability of the operator.

Let SΦ,Ψ denote the closure of this operator. It satisfies

SΦ,ΨxΨ=x*Φ,xM.

Its inverse equals SΨ,Φ.

The relative modular operator ΔΦ,Ψ is defined by

ΔΦ,Ψ=SΦ,Ψ*SΦ,Ψ.

Important properties of the relative modular operator are

ΔΦ,Φ=ΔΦ and SΦ,Ψ=JΔΦ,Ψ1/2.

where J is the modular conjugation operator for the vector Φ.

12 Araki’s relative entropy

Araki [22, 23] uses the relative modular operator ΔΦ,Ψ to define the relative entropy/divergence D (ϕψ) of the corresponding states ϕ = ωΦ and ψ = ωΨ by Dϕψ=logΔΦ,ΨΦ,Φ.

Proposition 12: The divergence D (ϕψ) satisfies D (ϕψ) ≥ 0 with equality if and only if ϕ = ψ.

Proof:

Let

ΔΦ,Ψ=λdEλ

denote the spectral decomposition of the operator ΔΦ,Ψ. From the concavity of the logarithmic function, it follows that

Dϕψ=logΔΦ,Ψ1Φ,Φ=logλ1dEλΦ,Φlogλ1dEλΦ,Φ=logΔΦ,Ψ1/2Φ,ΔΦ,Ψ1/2Φ=logΨ,Ψ=0.

This shows that the divergence cannot be negative.

If ϕ = ψ, then one has

Dϕϕ=logΔΦΦ,Φ=0

because ΔϕΦ = Φ.

Finally, D (ϕψ) = 0 implies that Φ is in the domain of log  ΔΦ,Ψ and that log  ΔΦ,ΨΦ = 0. The latter implies that

Ψ=ΔΦ,Ψ1Φ=Φ.

This shows that D (ϕψ) = 0 vanishes only when Φ = Ψ.

Theorem 2.4 of [35] shows that

logΔΦ,Ψ+JlogΔΨ,ΦJ=0.

Because Φ belongs, by assumption, to the natural positive cone P, it satisfies Φ = JΦ. Hence, one has also

Dϕψ=logΔΨ,ΦΦ,Φ.

13 A theorem

Each self-adjoint element h of the von Neumann algebra M defines an exponential arc with a generator equal to the energy function defined by h.

[21] constructs for each self-adjoint operator h in M a vector Φh in the natural positive cone P and calls h the relative Hamiltonian. Inspection of the explicit expression used in [21] shows that

Φh=Ω+XhΩ+Oh2(15)

with operator X given by

X=01/2duΔΩu.

The vector Φh defines a state ϕh by

ϕhx=eξhxΦh,Φh,xM.

Here, ξ(h) is the normalization

ξh=logΦh,Φh.

Theorem 3.10 of [35] implies that the state ϕh obtained in this way satisfies for all ψ in M

Dψϕh=Dψωψh+ξh.(16)

Take ψ = ϕh and ψ = ω to find that the normalization ξ(h) is given by

ξh=ϕhhDϕhω=ωh+Dωϕh.

Consider now the path γ defined by γt = ϕth. Then, (16) becomes

Dψγt=Dψωtψh+ζt.(17)

with

ζt=tγthDγtω=tωh+Dωγt=ξth.

From this last expression, one obtains

0Dγtω+Dωγt=tγthωh.

From (15), we infer that γt converges to ω as t ↓ 0. Hence, D (γtω) and D (ωγt) converge to 0 faster than t. This implies that the derivative ζ̇(0) exists and equals ω(h). This also implies that

ddtt=0Dψγt=ωhψh.(18)

Elimination of ζ(t) from (17) yields

Dψγt=Dψω+Dωγt+tωhψh.

This shows that γ is an exponential arc connecting γ1 to γ0 = ω.

Proposition 13: One has

γ̇0x=TΩhΩ,TΩxωx*Ω(19)

with the operator TΩ given by

TΩ=ΔΩ1logΔΩ1/2.(20)

It should be noted that this operator TΩ was introduced in [36].

Proof:

From (15), one obtains

γ̇0x=ddtt=0γtx=xXhΩ,Ω+xΩ,XhΩζ̇0ωx.(21)

Write

xXhΩ,Ω=01/2duxΔΩuhΩ,Ω=01/2duΔΩu/2hΩ,ΔΩu/2x*Ω

and

xΩ,XhΩ=01/2duxΩ,ΔΩuhΩ=01/2duΔΩu/2JΔΩ1/2x*Ω,ΔΩu/2JΔΩ1/2hΩ=01/2duJΔΩ1u/2x*Ω,JΔΩ1u/2hΩ=1/21duJΔΩu/2x*Ω,JΔΩu/2hΩ=1/21duΔΩu/2hΩ,ΔΩu/2x*Ω.

The two contributions to (21) can now be taken together. One obtains

γ̇0x=01duΔΩu/2hΩ,ΔΩu/2x*Ωζ̇0ωx=TΩhΩ,TΩx*Ωζ̇0ωx.

Take x = 1 to see that

ζ̇0=TΩhΩ,TΩΩ=ωh

so that it follows (19).

In summary, one can infer

Theorem 1: Let ω in M be a vector state with cyclic and separating vector Ω. Choose the divergence function equal to the relative entropy of Araki as defined by (15). For each self-adjoint element h in M, an energy function h is defined by h(ϕ)=ϕ(h) and there exists an exponential arc γ with generator h connecting some state γ1 of M to the state γ0 = ω. For any state ψ in M, the derivative of D (ψγt) at t = 0 exists and is given by ω(h) − ψ(h). The derivative of the exponential arc at t = 0 satisfies (19).

Further properties hold for the exponential arc of the above theorem.

Proposition 14: For any exponential arc γ constructed in Theorem 1, the derivative γ̇0 is a Fréchet derivative.

Proof:

Let Ξ(h) denote the remainder of order h2 in (15), i.e.,

Φh=Ω+XhΩ+Ξh.

Then one can use (19) for

γtxγ0xtγ̇0x=eζtxΦth,ΦthωxtxXhΩ,ΩtxΩ,XhΩ+tωhωx=eζt1+tωhωx+teζt1xXhΩ,Ω+xΩ,XhΩ+eζtxΞth,Ω+xΩ,Ξth+t2xXhΩ,XhΩ+txΞth,XhΩ+txXhΩ,Ξth+xΞth,Ξth.

This yields

γtγ0tγ̇0|eζt1+tωh|+2t|eζt1|XhΩ+2eζtΞth+eζtΞth+tXhΩ2.

Each of the terms in the right-hand side of this expression is of order less than t as t tends to 0. Hence, γ̇0 is a Fréchet derivative.

Proposition 15 (Additivity of generators): If the state ϕ is connected to the state ω by the exponential arc with generator h and ψ is connected to ϕ by the exponential arc with generator k, then ψ is connected to ω by the exponential arc with generator h + k and ω is connected to ψ by the exponential arc with generator −h.

For the proof, see Proposition 4.5 of [21].

14 The metric

Eguchi [37] introduced the technique of deriving the metric of the tangent space by taking two derivatives of the divergence. Application here yields the metric which is used in the Kubo–Mori theory of linear response [38, 39].

Consider two exponential arcs tγt and sηs with respective generators h and k. They connect the states γ1 and η1 to the reference state ω. The tangent vectors at s = t = 0 are γ̇0 and η̇0. They belong to the tangent space TωM. The scalar product between them is by definition given by

η̇0,γ̇0ω=sts=t=0Dηsγt.

Assume now that these exponential arcs are those constructed in Theorem 1. Then, one has

η̇0,γ̇0ω=ss=t=0ωhηsh=η̇0h=TΩkΩ,TΩhωhΩ=TΩkωkΩ,TΩhωhΩ,(22)

with the operator TΩ defined by (20). It should be noted that in most applications, one assumes that the expectations ω(h) of the generator h and ω(k) of the generator k vanish. Then, the result obtained here coincides with that used in [36]. In what follows, a non-vanishing expectation of the generators is taken into account.

Let us now discuss some technical issues. The scalar product is well-defined by (22). This follows from

Lemma 2: If two exponential arcs with initial point ω with generators h, respectively k, both in M, have the same initial tangent vector, then one has

TΩhωhΩ=TΩkωkΩ.

Proof:

Let γ and η be two exponential arcs with generators h, respectively k in M, such that γ0 = η0 = ω. Without restriction, assume that ω(h) = ω(k) = 0 and γ̇0=η̇0. Then, (19) implies that

TΩhkΩ,TΩx*Ω=0,xM.

Take x = hk. Then, it follows that TΩ(hk)Ω = 0.

This lemma shows that the map

γ̇0TΩhωhΩ(23)

is one-to-one and identifies the tangent vector γ̇0 with the vector TΩhΩ in the Hilbert space H.

Expression (22) defines a bilinear form. This follows from.

Lemma 3: The map (23) is linear.

Proof:

Let γ be an exponential arc with generator h in M. Then, tγϵt is an exponential arc with generator ϵh for any ϵ in [−1, 1] and the tangent vector is ϵγ̇0. Hence, (23) maps ϵγ̇0 onto ϵTΩhΩ.

Next, consider a pair of exponential arcs γ and η with generators k, and h, respectively, in M and with γ0 = η0 = ω. Let θ denote the exponential arc with generator h + k. It exists by Theorem 1. The state θt can then be written as

θtx=xΦth+tk,Φth+tk

with Φth+tk being the unique element in the natural positive cone representing the state θt. Now, use (15) to write

θtx=ωx+t201duxΔΩu/2h+kΩ,Ω+t201duxΩ,ΔΩu/2h+kΩ+ ot.

This implies

θ̇0=γ̇0+η̇0.

Both observations together prove the linearity of map (23).

Proposition 16: Expression (22) defines a non-degenerate scalar product on the space of tangent vectors of the form γ̇0 with γ an exponential arc as constructed in Theorem 1.

Proof:

The two lemmas show that (22) is a well-defined bilinear form. Positivity of the form is clear. The symmetry follows from (22). It remains to be shown that it is non-degenerate.

Assume that (γ̇0,γ̇0)=0. This implies

TΩhωhΩ=0,

with h the generator of γ. The operator TΩ is invertible—see the proof of Lemma II.2 of [36]. Hence, it follows that

hωhΩ=0.

Because O is separating for M, it follows that h is a multiple of the identity. The latter implies that γ̇=0.

15 Dual geometries

The geodesics of the e-connection are the exponential arcs. In the m-connection, the geodesics are made up by convex combinations of a pair of states. The m- and e-connections are each others’ dual with respect to the metric of Section 14.

Consider two states ω and ϕ in the manifold M. The tangent vector

γ̇t=ddtγt=ϕω,0<t<1,

is independent of t. Hence, it is a geodesic for the connection in which all parallel transport operators are taken equal to the identity operator. It should be noted that the tangent space TωM coincides with the space of σ-weakly continuous linear functionals χ, satisfying χ(1) = 0 and hence it is the same everywhere . This connection is by definition the m-connection.

For t in (0, 1), the tangent vector γ̇t belongs to the subspace Tγt of the tangent space TωM which is introduced in Section 6. Conversely, every vector χ in Tγt is the tangent vector of an m-geodesic passing through the point γt. However, this does not imply that through parallel transport Π(γtγs), the space Tγt maps onto the space Tγs.

The transport operators Π* of the dual geometry are defined by

ΠϕωV,Π*ϕωWω=V,Wϕ.

In this expression, V and W are vector fields and (⋅,⋅)ω is the scalar product defined in the previous section and evaluated at the point ω of the manifold M.

It can be shown that any exponential arc γ is a geodesic for this dual geometry. To do so, we have to show that

Π*γsγtγ̇s=γ̇t.

The tangent vector γ̇t at t = 0 is given by (19). Its value for arbitrary t is given by the following proposition.

Proposition 17: Let γ denote an exponential arc γ with generator h belonging to M. Let Φt be the normalized vector in the natural positive cone P representing the state γt. The derivative γ̇t is given by

γ̇tx=TΦthΦt,TΦtxγtx*Φt,xM.(24)

Proof:

The state γ1 is connected to ω by the exponential arc with generator h and γt is connected to ω by the exponential arc with generator th. Let

ψs=γ1st+s.

It follows from Proposition 8 that sψs is an exponential arc with generator (1 − t) h connecting γt to γ1. Application of (19) to the latter arc gives

ψ̇0x=ddss=0ψsx=1tTΨhΨ,Tψxωx*Ψ(25)

with Ψ = Φt. This implies (24) because ψ̇0=(1t)γ̇t.

Theorem 2: Any exponential arc γ with generator h in M is a geodesic for the dual of the m-connection with respect to the metric introduced in Section 14.

Proof:

Let tϕt be an exponential arc with generator k in M such that ϕ0 = γt. Fix t in [0, 1] and let Φt denote the normalized element of the natural positive cone P representing the state γt. Let η be an exponential arc with generator k starting at γt, i.e., η0 = γt. Because Π(γsγt) is the identity, the definition of the dual transport operator yields

η̇0,Π*γsγtγ̇sγt=η̇0,γ̇sγs=TΦtkγtkΦt,TΦtlγtlΦt,

with l the generator of the arc sγ(1−s)t+s. It equals l = (1 − t)h. This last expression equals

=η̇0,γ̇tγt.

By proposition 16, the scalar product (,)γt is non-degenerate. Therefore, one can conclude that

Π*γsγtγ̇s=γ̇t.

This shows that the exponential arc γ is a geodesic for the dual of the m-connection.

16 Finite-dimensional submanifolds

A finite set of linearly independent generators is shown to define a finite-dimensional submanifold in which all states are connected to the reference state by an exponential arc. The submanifold defined in this way is a dually flat quantum statistical manifold.

Let ω be the reference state of M. It is a vector state with a cyclic and separating vector Ω. Choose an independent set of self-adjoint operators h1, … , hn in M. By Theorem 1, there exists an exponential arc γ with generator h = θihi connecting some state γ1 in M to the state γ0 = ω. A parameterized family of states ωθ, θRn is now defined by putting ωθ = γ1. These states form a submanifold of M.

From the definition of an exponential arc, one obtains immediately that for any ψ in M

Dψωθ=Dψω+Dωωθ+θiωhiψhi.(26)

Take ψ = ωθ in this expression to find

θiωhiDωωθ+θiωhi=θiωθhiDωθωθiωθhi.(27)

Hence, the quantity θiω(hi) is maximal if and only if ωθ equals the reference state ω.

Proposition 18: Dual coordinates ηi are defined by

ηi=ωθhi.

They satisfy

ηiθj=i,jθ

with (⋅,⋅)θ equal to the scalar product (,)ωθ introduced in Section 14 and with basis vectors i equal to ∂ωθ/∂θi.

Proof:

Introduce the path γ(i) defined by

γi:tωθ+tgi.

It satisfies

θiωθ=i=γ̇i0.

By definition, ωθ+gi is the end point of the exponential arc with generator (θj+gij)hi. From Proposition 15, it then follows that γ(i) is an exponential arc with generator hi connecting ωθ+gi to ωθ. These arcs γ(i) are used in the calculation that follows.

The definition of the scalar product at the beginning of Section 14 gives

i,jωθ=γ̇i0,γ̇j0ωθ=sts=t=0Dγsiγtj=ss=0θjDγsiωθ=ss=0γsihj=ihj=ηjθi.

Corollary 3: There exists a potential Φ(θ) such that

ηi=Φθi.(28)

This follows because the scalar product is symmetric so that

ηjθi=i,jθ=j,iθ=ηiθj.

This symmetry is a sufficient condition for the potential Φ(θ) to exist.

Consider the following generalization of the potential introduced in Section 9.

Φθ=Dωωθ+θiωhi.(29)

Apply (18) to the exponential arc γ(i) which connects ωθ+gi to ωθ to find

θiDωωθ=ωθhiωhi.

This implies that Φ(θ) satisfies (28).

One can conclude that the selection of an independent set of self-adjoint operators h1, … , hn in M defines a parameterized statistical model θωθ of states on the von Neumann algebra M. An obvious basis in the tangent plane TωθM is formed by the derivative operators i. The scalar product (i,j)ωθ introduced in Section 14 starting from the relative entropy of Araki defines a Hessian metric on the tangent planes. Exponential arcs are geodesics for the e-connection which is the dual of the m-connection.

17 Discussion

• The manifold M under consideration consists of vector states on a sigma-finite von Neumann algebra M in its standard representation. Such a manifold has nice properties described by the Tomita–Takesaki Theory and hence is an obvious study object when exploring quantum statistical manifolds in an infinite-dimensional setting. Particular attention is given in the present work on the definition of the tangent planes. This is also a point of concern in the commutative context of manifolds of probability measures. See, for instance, the approach of [14]. A convenient choice for the tangent space TωM at the state ω in the manifold M is to take it equal to the space of all σ-weakly continuous Hermitian linear functionals χ on M vanishing on the identity operator I. However, it is well-possible that the equivalence class of smooth curves through ω with initial tangent equal to a given χ is empty. Approximate tangent planes are considered an alternative in Section 6. They form a subspace of TωM as defined previously. Nevertheless, the initial tangent vectors of Fréchet-differentiable paths starting at ω belong to the approximate tangent space. It is not clear whether the initial tangents of exponential arcs are dense in the approximate tangent space with respect to the inner product of Section 14. Further research is needed at this point.

• A new definition of exponential arcs is given. It depends on the choice of a divergence function/relative entropy defined on pairs of points in the manifold and on the choice of a generator which is a linear functional defined on a domain in the manifold. It is general enough to cover different approaches that one can follow to solve the non-uniqueness problem of the Radon–Nikodym derivative in the context of non-commutative probability. Nevertheless, one can prove in full generality nice properties such as uniqueness of the generator, existence of scalar potential, and Pythagorean relations. The additivity of generators when composing exponential arcs is shown in the specific context of Araki’s relative entropy. See Proposition 15.

• The second half of the paper focuses on the relative entropy of Araki. Only exponential arcs with bounded generators belonging to the von Neumann algebra are considered. This suffices to reach the goal of replacing the existing approach based on density matrices and Umegaki’s relative entropy. However, the solution of the problem mentioned previously regarding the extent of the tangent spaces most likely requires the handling of unbounded generators.

• The scalar product of Bogoliubov presented in Section 14 is used extensively in Linear Response Theory, also known as Kubo–Mori theory. Its link with the KMS condition of Section 2 is not highlighted in the present text. It is tradition in the Kubo–Mori theory and more generally in statistical mechanics to focus on a small number of variables. It is shown in Section 16 that the selection of a finite number of variables defines a quantum statistical manifold supporting Amari’s dually flat geometry.

Data availability statement

The original contributions presented in the study are included in the article/Supplementary Material: further inquiries can be directed to the corresponding author.

Author contributions

The author confirms being the sole contributor of this work and has approved it for publication.

Funding

The publication charges are covered by the Universiteit Antwerpen.

Conflict of interest

The author declares that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors, and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

1. Dixmier J. Les C*-algèbres et leurs représentations. Paris: Gauthier-Villars (1964).

Google Scholar

2. Dixmier J. Les algèbres d’operateurs dans l’espace Hilbertien. Paris: Gauthier-Villars (1969).

Google Scholar

3. Ruelle D. Statistical mechanics, Rigorous results. New York: W.A. Benjamin, Inc. (1969).

Google Scholar

4. Emch G. Algebraic methods in statistical mechanics and quantum field theory. New Jersey, United States: John Wiley & Sons (1972).

Google Scholar

5. Bratteli O, Robinson D. Operator algebras and quantum statistical mechanics. New York, Berlin: Springer (1979).

Google Scholar

6. Haag R, Hugenholtz NM, Winnink M. On the equilibrium states in quantum statistical mechanics. Commun Math Phys (1967) 5:215–36. doi:10.1007/bf01646342

CrossRef Full Text | Google Scholar

7. Takesaki M. Tomita’s theory of modular Hilbert algebras and its applications. In: Lecture notes in mathematics. Berlin: Springer-Verlag (1970).

Google Scholar

8. Chentsov NN. Statistical decision rules and optimal inference. In: Transl. Math. Monographs. Rhode Island, United States: American Mathematical Society (1972).

Google Scholar

9. Efron B. Defining the curvature of a statistical problem. Ann Stat (1975) 3:1189–242.

CrossRef Full Text | Google Scholar

10. Amari S. Differential-geometrical methods in statistics. In: Lecture notes in statistics. New York, Berlin: Springer (1985).

CrossRef Full Text | Google Scholar

11. Amari S, Nagaoka H. Methods of information geometry. In: Translations of mathematical monographs. Oxford, United Kingdom: Oxford University Press (2000).

Google Scholar

12. Ay N, Jost J, Vân Lê H, Schwachhöfer L. Information geometry. Berlin, Germany: Springer (2017).

Google Scholar

13. Petz D. Quantum information theory and quantum statistics. (Berlin: Springer-Verlag) (2008).

Google Scholar

14. Pistone G, Sempi C. An infinite-dimensional structure on the space of all the probability measures equivalent to a given one. Ann Stat (1995) 23:1543–61.

CrossRef Full Text | Google Scholar

15. Grasselli MR, Streater RF. On the uniqueness of the chentsov metric in quantum information geometry. Infin Dim Anal Quan Prob. Rel. Top. (2001) 4:173–82. doi:10.1142/s0219025701000462

CrossRef Full Text | Google Scholar

16. Streater RF. Duality in quantum information geometry. Open Syst Inf Dyn (2004) 11:71–7. doi:10.1023/b:opsy.0000024757.25401.db

CrossRef Full Text | Google Scholar

17. Streater RF. Quantum orlicz spaces in information geometry. Open Syst Inf Dyn (2004) 11:359–75. doi:10.1007/s11080-004-6626-2

CrossRef Full Text | Google Scholar

18. Jenčová A. A construction of a nonparametric quantum information manifold. J Funct Anal (2006) 239:1–20. doi:10.1016/j.jfa.2006.02.007

CrossRef Full Text | Google Scholar

19. Grasselli MR. Dual connections in nonparametric classical information geometry. Ann Inst Stat Math (2010) 62:873–96. doi:10.1007/s10463-008-0191-3

CrossRef Full Text | Google Scholar

20. Umegaki H. Conditional expectation in an operator algebra. IV. Entropy and information. Kodai Math Sem Rep (1962) 14:59–85. doi:10.2996/kmj/1138844604

CrossRef Full Text | Google Scholar

21. Araki H. Relative Hamiltonian for faithful normal states of a von Neumann algebra. RIMS (1973) 9:165–209.

CrossRef Full Text | Google Scholar

22. Araki H. Some properties of modular conjugation operator of von Neumann algebras and a non-commutative Radon–Nikodym theorem with a chain rule. Pac J Math (1974) 50:309–54. doi:10.2140/pjm.1974.50.309

PubMed Abstract | CrossRef Full Text | Google Scholar

23. Araki H. Relative entropy of states of von Neumann algebras. Publ RIMS Kyoto Univ (1976) 11:809–33.

Google Scholar

24. Pistone G, Cena A. Exponential statistical manifold. AISM (2007) 59:27–56. doi:10.1007/s10463-006-0096-y

CrossRef Full Text | Google Scholar

25. Pistone G. Nonparametric information geometry. In: F Nielsen, and F Barbaresco, editors. Geometric science of information. Berlin, Germany: Springer (2013). p. 5–36.

CrossRef Full Text | Google Scholar

26. Santacroce M, Siri P, Trivellato B. On mixture and exponential connection by open arcs. In: F Nielsen, and F Barbaresco, editors. Geometric science of information. Berlin, Germany: Springer (2017). p. 577–84.

CrossRef Full Text | Google Scholar

27. Naudts J. Exponential arcs in the manifold of vector states on a σ-finite von Neumann algebra. Inf Geom (2022) 5:1–30. doi:10.1007/s41884-021-00064-4

CrossRef Full Text | Google Scholar

28. Naudts J. Quantum statistical manifolds. Entropy (2018) 20:472. doi:10.3390/e20060472

PubMed Abstract | CrossRef Full Text | Google Scholar

29. Naudts J. Correction: Naudts, J. Quantum statistical manifolds. Entropy 2018, 20, 472. Entropy (2018) 20:796. doi:10.3390/e20100796

PubMed Abstract | CrossRef Full Text | Google Scholar

30. Ciaglia FM, Ibort A, Jost J, Marmo G. Manifolds of classical probability distributions and quantum density operators in infinite dimensions. Inf Geom (2019) 2:231–71. doi:10.1007/s41884-019-00022-1

CrossRef Full Text | Google Scholar

31. Simon L. Lectures on geometric measure theory. In: Proceedings of the centre for mathematical Analysis. Australia: Australian National University (1983).

Google Scholar

32. Niestegge G. Absolute continuity for linear forms on b*-algebras and a Radon-Nikodym type theorem (quadratic version). Rend Circ Mat Palermo (1983) 32:358–76. doi:10.1007/bf02848539

CrossRef Full Text | Google Scholar

33. Csiszár I. I-divergence geometry of probability distributions and minimization problems. Ann Probab (1975) 3:146–58.

Google Scholar

34. Cziszár I, Matús̆ F. Generalized minimizers of convex integral functionals, Bregman distance, Pythagorean identities. Kybernetika (2012) 48:637–89.

Google Scholar

35. Araki H. Relative entropy for states of von Neumann algebras II. Publ Rims, Kyoto Univ (1977) 13:173–92.

CrossRef Full Text | Google Scholar

36. Naudts J, Verbeure A, Weder R. Linear response theory and the KMS condition. Comm Math Phys (1975) 44:87–99. doi:10.1007/bf01609060

CrossRef Full Text | Google Scholar

37. Eguchi S. Information geometry and statistical pattern recognition. Sugaku Expositions (2006) 19:197–216.

Google Scholar

38. Kubo R. Statistical-Mechanical theory of irreversible processes. I General theory and simple applications to magnetic and conduction problems. J Phys Soc Jpn (1957) 12:570–86. doi:10.1143/jpsj.12.570

CrossRef Full Text | Google Scholar

39. Mori H. Transport, collective motion, and Brownian motion. Progr Theor Phys (1965) 33:423–55. doi:10.1143/ptp.33.423

CrossRef Full Text | Google Scholar

Keywords: exponential arcs, quantum statistical manifold, quantum divergence function, Araki’s relative entropy, dually flat geometry, Tomita–Takesaki theory, linear response theory, Kubo–Mori theory

Citation: Naudts J (2023) Exponential arcs in manifolds of quantum states. Front. Phys. 11:1042257. doi: 10.3389/fphy.2023.1042257

Received: 12 September 2022; Accepted: 06 January 2023;
Published: 07 February 2023.

Edited by:

Florio M. Ciaglia, Universidad Carlos III de Madrid de Madrid, Spain

Reviewed by:

Sorin Dragomir, University of Basilicata, Italy
Fabio Di Cosmo, Universidad Carlos III de Madrid de Madrid, Spain

Copyright © 2023 Naudts. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Jan Naudts, jan.naudts@uantwerpen.be

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.