1. Introduction
In the context of information geometry for finite-dimensional quantum systems, it is well-known that the canonical action
, where † denotes the usual adjoint of an operator, of the unitary group
on the manifold
of faithful quantum states provides symmetry transformations for every monotone quantum metric tensor on
pertaining to Petz’s classification [
1]. Therefore, the fundamental vector fields generating the canonical action of
are Killing vector fields for every quantum monotone metric tensor.
It is also known that the canonical action of
on
can be seen as the restriction to
of a nonlinear action of the general linear group
given by
The action
is transitive on
and turns it into a homogeneous manifold [
2,
3,
4,
5]. Therefore, the fundamental vector fields of the canonical action of
form a Lie-subalgebra of the algebra of fundamental vector fields of the action of
.
In [
6], it is shown that, in order to describe the fundamental vector fields of
, it is sufficient to consider the fundamental vector fields of the canonical action of
on
together with the gradient vector fields associated with the expectation-value functions
—where
is any self-adjoint element in the space
of bounded linear operators on
—by means of the so-called Bures–Helstrom metric tensor [
7,
8,
9,
10,
11,
12]. This instance provides an unexpected link between the unitary group
, the
-homogeneous manifold structure of
, the Bures–Helstrom metric tensor, and the expectation value functions.
However, this is not the only example in which a monotone metric tensor “interacts” with the general linear group
. Indeed, again in [
6], it is also shown that fundamental vector fields of the canonical action of
together with the gradient vector fields associated with the expectation value functions by means of the Wigner–Yanase metric tensor [
13,
14,
15,
16,
17,
18,
19] close on a representation of the Lie algebra of
that integrates to a group action given by
Of course, the action is different from the action , but it is still a transitive action so that is a homogeneous manifold also with respect to this action, and the underlying smooth structure coincides with the one related with . Moreover, a direct inspection shows that can be thought of as a kind of deformation of by means of the square-root map and its inverse on positive operators. This instance is better described and elaborated upon in the rest of the paper.
Finally, again in [
6], it is proved that there is another Lie group “extending” the unitary group
and for which a construction similar to the one discussed above is possible. This Lie group is the cotangent bundle
of
endowed with its canonical Lie group structure [
20,
21]. In this case, the gradient vector fields of the expectation value functions are built using the Bogoliubov–Kubo–Mori metric tensor [
22,
23,
24,
25,
26], and the action is given by
where
is a self-adjoint element which is identified with a cotangent vector at
U. Once again, we obtain a transitive action on
associated with a homogeneous manifold structure whose underlying smooth structure coincides with the two other smooth structures previously mentioned.
It is important to note that, when we restrict to the unitary group
, all the group actions we considered reduce to the canonical action
of the unitary group whose importance in quantum theories is almost impossible to overestimate.
Once we have these three “isolated” instances, it is only natural to wonder if they are truly isolated cases, or if there are other monotone metric tensors for which a similar construction is possible. In [
27], this problem is completely solved in the two-level case in which a direct, coordinate-based solution is possible. The result is that the only two groups for which the aforementioned construction works are precisely the general linear group
and the cotangent group
. Moreover, in the case of
, the only compatible action is the action
already described in [
6], while, for
, there is an entire family of compatible smooth actions parameterized by a real number
and given by
All these actions are connected with a different quantum metric tensor. For instance, when the Bures–Helstrom metric tensor is recovered, while the Wigner–Yanase metric is recovered when . All the other cases correspond to Riemannian metric tensors on which are invariant under the standard action of .
In this work, we further investigate the problem by showing that all the group actions and metric tensors found in [
27] for a two-level system actually appear also for a quantum system with an arbitrary, albeit finite, number of levels. Moreover, we characterize all the values of
for which the Riemannian metric tensor associated with the action
in Equation (
5) is actually a quantum monotone metric tensor (
cfr. Proposition 1).
The work is structured as follows. In
Section 2, we discuss those differential geometric properties of the manifold of normalized and un-normalized quantum states that are necessary to the proof of our main results. In
Section 4, we set up the problem and prove our main results for
and
, namely, Propositions 2 and 3. In
Section 5, we discuss our results and some possible future directions of investigation.
2. Geometry of (Un-Normalized) Quantum States
In this section, the construction of the space of quantum states [
7,
28] is briefly described and some of its geometric features are recalled; this gives the setting for our discussion. Then, we give a hint to the role played by group actions in the context of Quantum Mechanics and introduce some particular group actions that will be needed in order to get to the main result of this work. Finally, the concept of
monotone metric, which is crucial in the context of Quantum Information Geometry, is introduced.
In standard quantum mechanics [
29,
30], a quantum system is mathematically described with the aid of a complex Hilbert space
. The bounded observables of the system are identified with the self-adjoint elements in the algebra
of bounded linear operators on
, and the set of all such elements is denoted with
. The physical states of the system are identified with the so-called
density operators on
. In order to define what a density operator is, we start recalling that
is said to be positive semi-definite if
and its customary to write
and
when it is also invertible. The space of positive semi-definite operators is denoted by
so that
and its elements may be referred to as
un-normalized quantum states for reasons that are clarified below. From a geometrical point of view,
is a convex cone. A
density operator is just an element in
satisfying the normalization condition
. This linear condition defines a hyperplane
in
. As anticipated before, physical states are identified with density operators, and thus, the space of quantum states reads
and thus, the nomenclature “un-normalized quantum states” for elements in
appears justified. Clearly,
is given by the intersection between the convex cone
and the hyperplane
and thus is a convex set.
Remark 1. It is worth noting that there is a very deep analogy between the space of positive semi-definite operators and the space of classical measures on the measurable space M which are absolutely continuous with respect to the reference measure μ. This analogy finds the perfect mathematical formalization in the context of the theory of -algebras [31,32,33,34], where it turns out that both and arise as the space of normal positive linear functionals on suitable -algebras. This parallel is being exploited to give a unified account of some aspects of classical and quantum information geometry [3,22,24,35,36]. In the rest of this work, as it is often done in the context of quantum information theory [
37], we restrict our attention to the finite-dimensional case in which
has complex dimension
. Then, it is proved that both
and
may be endowed with the structure of stratified manifold whose underlying topological structure coincides with the topology inherited from
[
38]. It turns out that the
strata of these stratified manifolds can be described in terms of a particular action of the general linear group
[
2,
3,
4,
5,
39]. Specifically,
acts on the whole
according to
It is important to note that, when we restrict the action
in such a way that it acts only on positive elements and only by means of elements in
, it reduces to the canonical action
of the unitary group. The action
is linear, and it is a matter of direct inspection to check that it preserves both
and
and that the orbits through
are made of positive semi-definite operators of the same rank, denoted by
where
is the rank. These orbits thus become homogeneous manifolds and their underlying smooth structures agree with those associated with the stratification of
[
38].
In particular, we are interested in the maximal stratum
i.e., the space of invertible elements in
, which forms the open interior of
. The tangent space
of
at
is isomorphic to
, since
is an open set in
. Since
is a homogeneous manifold, the tangent space at each point can be described in terms of the fundamental vector fields of the action
evaluated at a point
[
40]. Recalling that the Lie algebra
of
is essentially
endowed with the standard commutator, a curve in the group
can be written as
with
and
self-adjoint operators. Therefore, the fundamental vector field associated with
at the point
reads
where we have used the notation
and we have set
and
As mentioned before, when we restrict
to
we obtain the canonical action
of
. Therefore, since the Lie algebra
is just the space of skew-adjoint elements in
, setting
in Equation (
13), we immediately obtain that the fundamental vector fields of
are recovered as the fundamental vector fields
of the action
. Concerning the vector fields of the form
, taking
, we obtain the vector field
which represents the infinitesimal generator of the Lie group
acting on
by dilation. However, in general, it turns out that
[
2,
3] so that they do not form a Lie subalgebra.
Besides
, also the cotangent Lie group
acts on
in such a way that the latter becomes a homogeneous manifold of
. Specifically, the action is given by
where we used the canonical identifications
. It is not hard to check that, if we restrict to
by considering only elements of the type
, the action
reduces to the action
of
on
. The action
is smooth with respect to the previously mentioned smooth structure on
associated with the action
, and it is a transitive action. Therefore, we conclude that the smooth structure underlying
when thought of as a homogeneous manifold for
coincides with the smooth structure underlying
thought of as a homogeneous manifold for
. The action
is basically related with the isomorphism
given by
and its inverse
given by
. Indeed, it is clear that
acts on
through
and it is a matter of direct inspection to show that
Thinking of
as a subgroup of the rotation group of the vector space
, it follows that the action
coincides with the restriction to
of the standard action of the affine group on
. The action
cannot be extended to the whole
essentially because
and its inverse cannot be extended. Concerning the fundamental vector fields
of
, we have that
as in Equation (
16), while
where we used the well-known equality
which is valid for every smooth curve
inside
(remember that the canonical immersion of
inside
is smooth) [
41].
We turn now our attention to faithful quantum states. The action in Equation (
10) does not preserve the hyperplane
in Equation (
8), and thus, it also does not preserve
. However, as already anticipated in Equation (
1), it is possible to suitably renormalize
to obtain the action
The normalization is recovered at the expense of the linearity/convexity of the action. However, when we restrict to the unitary group
, the action
reduces to the canonical action
of the unitary group on the space of states which does preserve convexity. Analogously to what happens for the action
on
, the orbits of
are made up of quantum states with the same fixed rank, and any such orbit is denoted as
where
k is the rank. These orbits thus become homogeneous manifolds and their underlying smooth structures agree with those associated with the stratification of
[
38]. Moreover, each manifold
can be seen as a submanifold of
singled out by the intersection with the affine hyperplane
. It is worth mentioning that the partition of
in terms of manifold of quantum states of fixed rank was also exploited in [
42,
43]; however, as far as the authors know, the homogeneous manifold structures was firstly understood in [
4,
5] and the stratified structure in [
38].
Remark 2. Building on Remark 1, for a reader familiar with Classical Information Geometry, it may be useful to think of the space of quantum states of an n-level quantum system as the quantum analogue of the -simplex, with thestrata of the space of quantum states taking the place of the faces of the simplex. A thorough discussion of this analogy can be found in [44,45,46]. In particular, we focus on the stratum of maximal rank, i.e., invertible, or
faithful states
The tangent space
of
at
is given by self-adjoint operators with the additional property of being traceless, i.e., we have
Since
is a homogeneous manifold, its tangent space can be described using the fundamental vector fields of the action
following what is done for
. The fundamental vector fields of the action
evaluated at a point
are given by
where
is defined as in (
13) and, now, we have set
and
Again in analogy with what happens on , the fundamental vector fields of the action of are identified with the vector fields .
As anticipated in the Introduction, the cotangent Lie group
also acts on
through the action
given in Equation (
3). This action is smooth with respect to the previously mentioned smooth structure on
associated with the action
, and it is a transitive action. Therefore, we conclude that the smooth structure underlying
when thought of as a homogeneous manifold for
coincides with the smooth structure underlying
thought of as a homogeneous manifold for
. Moreover, when restricting to
, a direct computation shows that the action
reduces to the standard action
of
on
. The fundamental vector fields
of
are then easily found. In particular,
as in Equation (
29), and
where we again exploited Equation (
23) (remember that the canonical immersion of
inside
is smooth). Concerning the vector fields in Equation (
31), it is worth mentioning that they already appeared in [
25] in connection with the Bogoliubov–Kubo–Mori metric tensor, and then, in the recent work [
47], where the finite transformations they induced are exploited in the definition of a Hilbert space structure on
, which is the quantum counterpart of a classical structure relevant in estimation theory. However, as far as the author know, the group-theoretical aspects relating the vector fields in Equation (
31) with the action
of
were first investigated in [
6].
Despite the lack of a universally recognized physical interpretation for un-normalized quantum states in
, it turns out that they provide a more flexible environment in which to perform the mathematics needed to prove the main result of this work. Intuitively speaking, it is already clear from the very definition of the actions
and
that imposing the linear normalization constraint needed to pass to (normalized) quantum states leads to the emergence of nonlinear aspects which destroy the inherent convexity of the space of quantum states. In fact, following the ideology expressed in [
48], it can also be argued that the choice of a normalization has a somewhat arbitrary flavor that does not really encode physical information, because basically nothing really serious happens if we decide to normalize to
rather than to 1. Following this line of thought, we will always work on
making sure that all the structure and results may be appropriately “projected” to
. For this purpose, it is relevant to introduce a projection map from
as
and an associated section given by the natural immersion map
reading
It is not hard to show that j is an embedding, while is a surjective submersion. Moreover, it is also possible to “extend” these maps to the whole and in the obvious way, thus obtaining a continuous projection map and a continuous immersion map that preserve the stratification of and and are smooth on each strata.
As mentioned before, bounded physical
observables are described by means of self-adjoint operators in
. Then, to any observable
, it is possible to associate a smooth function
given by
this is referred to as
expectation value function of the observable
. Of course, expectation value functions can also be defined on the space of quantum states
setting
and it turns out that
is connected to
by means of the pull-back with respect to
j, i.e., it holds
By relaxing smoothness to continuity, it is possible to extend the expectation value functions to the whole and the whole .
It is a matter of direct calculation using the very definition of fundamental vector fields for both
and
(
cfr. Equations (
4) and (
11)) to show that
where
is as in Equation (
15).
By direct computation, it is possible to spot an interesting intertwine between the maps
and
i and the actions
and
given by
where
is the identity map on
. Analogously, we obtain
where
is the identity map on
. Equations (
38) and (
39) explain in which sense
and
are a kind of normalized version of the actions
and
, respectively. The immersion map
j also allows us to obtain a pointwise relation between the fundamental vector fields
and
and between the fundamental vector fields
and
in terms of the tangent map
to
j at
. Indeed, from Equations (
16)–(
18), (
29), and (
30), it follows that
Accordingly, we conclude that
is
j-related with
while
is
j-related with
. Analogously, from Equations (
22), (
18) and (
31), it follows that
which means that
is
j-related with
.
3. Quantum Monotone Metric Tensors
In the classical case, the Riemannian aspects of most of the manifolds of probability employed in statistics, inference theory, information theory, and information geometry are essentially encoded in a single metric tensor (we are here deliberately “ignoring” all those Wasserstein-type metric tensors simply because their very definition depends on the existence of additional structures on the sample space), namely, the Fisher–Rao metric tensor [
49,
50,
51]. In the case of finite sample spaces, Cencov’s pioneering work [
52] investigated the Fisher–Rao metric tensor from a category-theoretic perspective and uncovered the uniqueness of this metric tensor when some invariance conditions are required. Specifically, let
denote the n-dimensional simplex in
, i.e., the space of probability distributions on a discrete sample space with n elements, and let
denote the interior of
, the space of probability distributions with full support. Note that
is a smooth,
-dimensional manifold while
is a smooth manifold with corners.
A linear map
is called a
Markov morphism if
, and a Markov morphism
F is called a
congruent embedding if
is diffeomorphic to
. Congruent embeddings where studied by Cencov who characterized the most general form of these maps (
cfr. [
53] for yet another characterization of congruent embeddings).
According to Cencov, the relevant geometrical structures on
must all be left unchanged when suitably acted upon by congruent embeddings. For instance, setting
, a family
with
a smooth Riemannian metric tensor on
is called
invariant if
for every congruent embedding
. Cencov’s incredible result was to show that, up to an overall multiplicative positive constant, there is only one invariant family of Riemannian metric tensor for which
coincides with the Fisher–Rao metric tensor. Then, much effort has been devoted to extend Cencov’s uniqueness result from the case of finite sample spaces to the case of continuous sample spaces leading, for instance, to a formulation on smooth manifolds [
54] and a very general formulation valid for very general parametric models [
48].
As already hinted at in Remarks 1 and 2, the manifold
may be thought of as the quantum analogue of
in the case of finite-level quantum systems. Then, the quantum analogue of a Markov morphism is a completely-positive and trace-preserving linear (CPTP) map
(
cfr. [
55] for the precise definition of CPTP maps and [
56,
57] for their role in quantum information). Quite trivially, a
quantum congruent embedding could be defined as a CPTP map
such that
is diffeomorphic to
. A typical example of quantum congruent embedding is given by
. As far as the authors know, there seems to be no general characterization of these maps at the moment as there is in the classical case.
Inspired by Cencov’s work, Petz investigated the following problem: to characterize the families
with
a smooth Riemannian metric tensor on
satisfying the monotonicity property
for every CPTP map
and for all
. He was able to prove [
1] that, up to an overall multiplicative positive constant, these families of
monotone quantum metric tensors are completely characterized by operator monotone functions
[
58] satisfying
In particular, if
is a family of monotone metric tensors, then
where
is a constant,
are vectors in
,
is a superoperator on
given by
with
f the operator monotone function mentioned before, and
and
are two linear superoperators on
whose action is given by the left and right multiplication by
.
We briefly mention a recent development towards the use of non-monotone metric tensors in quantum information theory [
59].
Since every
n-dimensional complex Hilbert space
is isomorphic to
, we can almost immediately generalize Equation (
44) to define a quantum monotone metric tensor
on
setting
In the following, for the sake of notational simplicity, we often simply write instead of because the Hilbert space is already clear from the context.
If we introduce the operators
diagonalizing
, that is, such that
we can also introduce the superoperators
acting on
according to
and it is then a matter of straightforward computation to check that
where
are the eigenvalues of
. Now, whenever
, from Equations (
46) and (
49), it follows that
where
and
are the diagonal elements of
and
with respect to the basis of eigenvectors of
. It is relevant to note then that in this case, we have
where
is the classical Fisher–Rao metric tensor on
, and we have set
,
, and
. Equation (
51) holds for every choice of the operator monotone function
f.
As mentioned before, the action
of
in (
4) gives rise to CPTP maps from
into itself. Moreover, these maps are invertible and their inverses are again CPTP maps from
to itself. Therefore, the monotonicity property in Equation (
42) becomes an invariance property, and we conclude that the fundamental vector fields
of the action
(
cfr. Equation (
29)) are Killing vector fields for every monotone quantum metric tensor
. Consequently, the unitary group
acts as a sort of universal symmetry group for the metric tensors classified by Petz and thus occupies a prominent role also in the context of Quantum Information Geometry.
To explicitly prove our main results, it is better to work first on
and then “project” the results down to
. Accordingly, we need a suitable extension of the monotone quantum metric tensors to
, very much in the spirit of Campbell’s work on the extension of the Fisher–Rao metric tensor to the non-normalized case of finite measures [
53]. Kumagai already investigated this problem and provided a complete solution of Petz’s problem when the normalization condition on quantum states is lifted [
60]. Quite interestingly, the result very much resembles Campbell’s result in the sense that the difference with the normalized case is entirely contained in a function
and a family
of operator monotone functions satisfying
.
In our case, however, it is not necessary to exploit the full level of generality of Kumagai’s work. It suffices to find a Riemannian metric tensor
on
such that
where
is the canonical immersion and
is a monotone quantum metric tensor as in Equation (
46). Accordingly, we consider
as given by
where
f is the operator monotone function appearing in Equation (
46) (and thus satisfying Equation (
43)),
,
, and
is as in Equation (
45). Equation (
53) corresponds to the choice
and
in Kumagai’s classification.
If we introduce the operators
diagonalizing
, we can proceed as in the normalized case to obtain an equation analogous to Equation (
49) so that, recalling Equation (
43), we immediately obtain
4. Lie Groups and Monotone Quantum Metric Tensors
We are interested in classifying all those actions of
and
on
that behave in the way described in the Introduction with respect to suitable monotone quantum metric tensors. Specifically, we want to find all those actions, say
, of either
or
on
for which there is a monotone metric tensor
on
such that the fundamental vector fields
of the standard action
of
on
together with the gradient vector fields
associated with the expectation value functions
close on a representation of the Lie algebra of either
or
that integrates to the action
. From the results in [
6], we know that there are at least 3 monotone metric tensors for which this construction is possible for any finite-level quantum system. Moreover, from the results in [
27], we know that in the case of two-level quantum systems, the Lie groups
and
are the only Lie groups for which the construction described above is actually possible. Here, we want to understand if the group actions of
and
found in [
27] can be extended from a 2-level quantum system to a system with an arbitrary, albeit finite, number of levels.
For this purpose, it is important to recall all those properties, shared by
and
and by their actions, that are at the heart of the results of [
6,
27]. First of all, both
and
contain the Lie group
as a Lie subgroup, and contain the elements
with
and
the identity operator on
. Then, all the (transitive) actions of both
and
on
appearing in the analysis of [
6,
27] arise as a sort of normalization of suitable (transitive) actions on
. Specifically, if
G denotes either
or
, then every
G-action
on
can be written as
with
a
G-action on
satisfying
for every
, for every
, and for every
. Moreover, among all those actions
satisfying the properties discussed above, there is a preferred action
(the action
in Equation (
10) for
, and the action
in Equation (
19) for
) such that every relevant action
can be written as
with
a smooth diffeomorphism arising from a smooth diffeomorphism
by means of functional calculus and such that
and
where
is a basis of
made of eigenvectors of
.
Equation (
57) implies that the map
is equivariant with respect to the action
and
, which in turn implies that the fundamental vector fields of
are
-related with that of
(
cfr. chapter 5 in [
40]). By the very definition of
-relatedness (
cfr. Chapter 4 in [
40]), denoting with
a fundamental vector field of
and with
a fundamental vector field of
, it follows that
We exploit Equation (
60) to explicitly describe how the fundamental vector fields
of the action
of
on
(
cfr. Equations (
10) and (
16)) transform under
. We then equate the result with the gradient vector field associated with the expectation value function
by means of the metric tensor
as in Equation (
53), thus obtaining an explicit characterization of the diffeomorphism
and the operator monotone function
f compatible with the equality. Finally, with this choice of
and
f, we prove that the gradient vector fields
associated with the expectation value functions
on
by means of the monotone quantum metric
as in Equation (
46) correspond to the fundamental vector fields
of the action
of
on
associated with the action
on
.
A similar procedure is then applied to the fundamental vector fields
of the action
of
on
(
cfr. Equations (
19) and (
22)).
4.1. The General Linear Group
Following [
6,
27], when considering the general linear group
, the reference action
appearing in Equation (
57) is the action
in Equation (
10). Therefore, denoting with
a fundamental vector field of
and with
a fundamental vector field of
, from Equations (
14) and (
60), and [
61] (Theorem 5.3.1), it follows that
where □ denotes the Schur product with respect to the basis of eigenvectors of
, and
with
the eigenvalues of
and with
the basis of
of eigenvectors of
and
(
cfr. Equation (
59)). Moreover, a direct computation shows that
where
are the components of
in the basis given by the eigenvectors of
. We thus conclude that
Now, we require that
is the gradient vector field of the expectation value function
with respect to the metric tensor
defined as in
Section 3 in order to characterize the function
f. From the very definition of the gradient vector field, it follows that
holds for any vector field
on
. On the other hand, it also holds that
so that, comparing Equation (
65) with Equation (
66), we obtain
Exploiting Equation (
49), it follows that Equation (
67) becomes
Comparing Equation (
64) with Equation (
68), we obtain
and
Equation (
69) implies
with
, so that, because of Equation (
70), the function
f in
must be of the form
A direct check shows that the function
f in Equation (
72) satisfies the properties listed in Equation (
43) for all
, but we do not know if it is operator monotone for every
. The following proposition shows that
f is operator monotone if and only if
.
Proposition 1. The function f in Equation (72) is operator monotone if and only if . Proof. When
it is
which is known to be operator monotone and to be associated with the Bures–Helstrom metric tensor [
1].
The function
f as in Equation (
72) is clearly
in
and it is continuous in
. When
, it holds
which means that there is
such that
is decreasing for
, and thus,
f cannot be operator monotone. Note that (
73) is no longer valid when
because of the term
When
, we consider the rational case
with
since the passage to an irrational
is obtained by continuity just as in [
62] (Proposition 3.1). Following [
62] (Proposition 3.1), we write
so that
Since
, the functions
are operator monotone according to [
62] (Theorem LH-1), and thus, the function
f in Equation (
75) is operator monotone because it is the sum of operator monotone functions. □
Finally, when
is as in Equation (
69) and
f is as in Equation (
72), we prove that the fundamental vector fields
of the normalized action
of
on
associated with
by means of Equation (
55) are indeed the gradient vector fields associated with the expectation value functions
by means of the monotone metric tensor
. Indeed, from Equation (
55), it follows that
Equation (
77) is equivalent to
for all
and all
, and this last instance is equivalent to the fact that
is
i-related with the vector field
for all
.
To finish the proof of the proposition, we need to prove that
is actually the gradient vector field of the expectation value function
for every
. For this purpose, we compute
Since we proved that
is be the gradient vector field associated with
by means of
, Equation (
80) becomes
The second term on the right-hand-side of Equation (
81) vanishes. Indeed, Equation (
53) implies that
From Equation (
54), we conclude that Equation (
82) becomes
where the last equality follows from Equation (
27). Inserting Equation (
83) in Equation (
81), we obtain
for every fundamental vector field of
of the type
, for every vector field
V on
, and for and every
. Equation (
84) is equivalent to the fact that
is the gradient vector field associated with the expectation value function
by means of
for every
as desired.
Collecting the results proved in this subsection, we obtain the following proposition.
Proposition 2. The function f given byis operator monotone and satisfies Equation (43) if and only if . In these cases, denoting with the fundamental vector fields of the canonical action α of on as in Equation (4), if is the associated monotone quantum metric tensor on as in Equation (46) and is the gradient vector field associated with the expectation value function with , the family of vector fields on close an anti-representation of the Lie algebra of the general linear group integrating to the group actionThe action in Equation (86) is transitive on for every . In particular, when , we recover the Bures–Helstrom metric tensor and the action β in Equation (1), while when , we recover the Wigner–Yanase metric tensor and the action in Equation (2). 4.2. The Cotangent Group of the Unitary Group
Following what is done in
Section 4.1, we consider an action
associated with the action
(
cfr. Equation (
19)) by means of Equation (
55) with
. The fundamental vector fields
of
are obtained as follows. From Equations (
22) and (
60), and [
61] (Theorem 5.3.1), it follows that
where □ denotes the Schur product with respect to the basis of eigenvectors of
, and
with
the eigenvalues of
and with
the basis of
of eigenvectors of
and
(
cfr. Equation (
59)). On the other hand, from Equations (
22) and (
59), it follows that
so that, exploiting Equations (
88) and (
89), Equation (
87) becomes
In analogy with what is done in
Section 4.1, we now require that
is the gradient vector field of the expectation value function
with respect to a metric tensor
defined as in
Section 3 in order to characterize the function
f. From the very definition of gradient vector field, it follows that
holds for any vector field
on
. On the other hand, it also holds that
so that, comparing Equation (
91) with Equation (
92), we obtain
Exploiting Equation (
49), it follows that Equation (
93) becomes
Comparing Equation (
90) with Equation (
94), we obtain
and
Equation (
95) implies
with
, and it is worth noting that the family of diffeomorphisms found here is the same as that found in
Section 4.1 in the case of the general linear group
(
cfr. Equation (
71)). Because of Equations (
96) and (
97), the function
f in
must be of the form
which is precisely the operator monotone function associated with the Bogoliubov–Kubo–Mori metric tensor up to the constant
[
1]. Note that the positive constant
is here arbitrary differently from what happens for
(
cfr. Section 4.1).
It is a matter of direct computation to check that the form of
in Equation (
97) implies that the action
associated with the action
(
cfr. Equation (
19)) by means of Equation (
55) with
reads
so that
(
cfr. Equations (
89), (
90) and (
97)). Consequently, the fundamental vector fields
of the normalized action
associated with
by means of Equation (
55) read
Equation (
101) is equivalent to
for all
and all
, and this last instance is equivalent to the fact that
is
j-related with the vector field
for all
.
Now, proceeding in complete analogy with what is done in
Section 4.1, it is possible to prove that, when
and
f are as in Equations (
97) and (
98), respectively, then the fundamental vector field
is the gradient vector field associated with the expectation value function
by means of the monotone quantum metric tensor
(coinciding with the Bogoliubov–Kubo–Mori metric tensor up to the constant
) for all
. Collecting the results in this subsection, we obtain the following proposition.
Proposition 3. Given the operator monotone functionsatisfying Equation (43) and associated with the Bogoliubov–Kubo–Mori metric tensor (up to the constant factor ) through Equation (46) [1], denoting with the fundamental vector fields of the canonical action α of on as in Equation (4), and denoting with the gradient vector field associated with the expectation value function with by means of , the family of vector fields on closes an anti-representation of the Lie algebra of the cotangent group , integrating to the group actionThe action in Equation (104) is transitive on for every . 5. Conclusions
There are several ways in which the results presented here can be further developed in order to fully understand how the 2-dimensional picture discussed in [
27] extends to arbitrary finite dimensions.
First of all, concerning the Lie group
, it is necessary to understand if there exist smooth transitive actions on
that are not of the form
(
cfr. Equations (
10) and (
57)). Then, it is necessary to understand if there exist smooth transitive actions on
that do not arise from smooth actions of
on
as in Equation (
55). If the answer to both these questions are negative, then it follows that the only actions of
on
whose associated Lie algebra anti-representations can be described in terms of the fundamental vector fields of the standard action of
on
(
cfr. Equations (
4) and (
29)) and the gradient vector fields
associated with the expectation value functions
by means of a suitable monotone quantum metric tensor are those found in this work.
Concerning the group
, it is necessary to understand if there exist smooth transitive actions on
that do not arise from smooth actions of
on
as in Equation (
55). If the answer to this question is negative, then it follows that the only action of
on
whose associated Lie algebra anti-representations can be described in terms of the fundamental vector fields of the standard action of
on
(
cfr. Equations (
4) and (
29)) and the gradient vector fields associated with the expectation value functions
by means of a suitable monotone quantum metric tensor are the ones found in this work, that is, the one associated with the Bogoliubov–Kubo–Mori metric tensor.
Besides the cases involving the Lie groups
and
, it is also necessary to understand if, for a quantum system whose Hilbert space
has dimension greater than 2, there exists other Lie groups acting smoothly and transitively on
and whose Lie algebra anti-representation can be described in terms of the fundamental vector fields of the standard action of
on
(
cfr. Equations (
4) and (
29)) and the gradient vector fields associated with the expectation value functions
by means of suitable monotone quantum metric tensors. Concerning this instance, something can be said on some general properties any such Lie group
must possess. First of all, the unitary group
must appear as a subgroup of
and
. This last condition follows from the fact that the gradient vector fields associated with the expectation value functions
are labeled by elements in
, and thus, the dimension of the Lie algebra
of
is twice that of the Lie algebra of
. From this last observation, it also follows that
as a vector space. Moreover, since
must be a subgroup of
, there must be a decomposition of
as in Equation (
105) for which
is a Lie subalgebra isomorphic to
. Then, as already argued in [
6], the requirement that the fundamental vector fields
of the standard action
of
on
are Killing vector fields for every monotone quantum metric tensor
imposes additional constraints on the possible commutator between these vector fields and the gradient vector fields
associated with the expectation value functions
. Specifically, since
is the gradient vector field associated with the expectation value function
for every
, it follows that
where we used Equation (
37), and the fact that
because the fundamental vector fields of the action
of
are Killing vector fields for all monotone quantum metric tensors. From Equation (
106), we conclude that
for every
. Then, since the differential of the expectation value functions provide a basis for the differential forms on
, Equation (
107) is equivalent to
Equation (
108) fixes the Lie bracket between elements of
and its complement, thus leaving us with the freedom to only define the bracket among elements that lies in the complement of
inside the Lie algebra
of
.
We are currently investigating all the problems discussed in this section and we plan to address them in detail in the (hopefully not too distant) future.