1. Introduction
Quantum information has the reputation of being a futuristic field full of far-reaching promises. The field has attracted researchers from many different branches of science and engineering, whose efforts have greatly improved our understanding of the physical nature of information, and hopefully will provide us with new cutting-edge technological innovations in the future. Quantum information theory has been applied to such diverse areas as bio-science, nano-technology, economics, and game theory [
1,
2,
3].
In recent years, new experimental results have shed light on some murky and hidden parts of quantum information, and have also opened up new opportunities beyond our expectations. Furthermore, new theoretical tools, mostly from geometry and topology, have been successfully applied to the field. For example, geometrical considerations led to the important characterization of entanglement and the development of efficient, error-prone systems for quantum computers.
Distance measures are some of the most basic geometrical tools used in quantum information theory. Such measures are, for example, used to compare the input and output of quantum channels and gates, and hence to quantify to what extent information is preserved, or altered, by quantum processes. Examples of well-known distance measures are the trace-distance, fidelity, and Bures distance [
4,
5,
6]. In this paper we propose a new distance measure that we call the dynamic distance measure. This distance measure is defined for all pairs of isospectral,
i.e., unitarily equivalent, mixed states. We show that the dynamic distance is a proper distance measure—a verification that, despite the naturalness of the definition, requires a surprisingly extensive geometric machinery.
Here is the outline of the paper. In
Section 2 we define the dynamic distance measure, and state the main result. In
Section 3 we introduce a geometric framework, designed and tailored to our needs, and in
Section 4 we in detail discuss properties of the dynamic distance measure and prove the main result.
Section 5 contains a derivation of the dynamic distance between distinguishable isospectral mixed states. Finally, in
Section 6 we compare the dynamic distance measure and the Bures distance.
2. Isospectral Mixed States and the Dynamic Distance Measure
Mixed quantum states can be represented by density operators, i.e., self-adjoint, positive, trace-class operators with unit trace. We denote the space of density operators for a quantum system modeled on a Hilbert space by , and its subspace of density operators with finite rank at most k by .
A density operator that evolves according to a von Neumann equation remains in a single orbit of the left conjugation action of the unitary group of on . The orbits are in one-to-one correspondence with the possible spectra for density operators on . By the spectrum of a density operator of rank k we mean its descending sequence of positive eigenvalues, repeated in accordance with their multiplicity. We henceforth assume σ to be fixed, and write for the corresponding orbit.
Suppose
and
are two density operators in
. Let
H be a Hamiltonian operator on
, and assume that a curve
ρ satisfies the boundary value von Neumann equation:
We then define the
H-distance from
and
to be the path integral of the uncertainty of
H along
ρ,
We also define the
dynamic distance between
and
to be
where the infimum is taken over all Hamiltonians
H for which the boundary value problem in Equation (
1) has a solution. The dynamic distance measure is defined for each pair of isospectral density operators because any two such can be connected by a solution to some von Neumann equation. The main result of this paper is that the dynamic distance measure is a proper distance measure on
:
Theorem 1. The dynamic distance measure is a proper distance measure.
Recall that a proper distance measure is a binary function dist that satisfies the following conditions:
Positivity: | |
Non-degeneracy: | |
Symmetry: | |
Triangle inequality: | |
One can show that the dynamic distance measure is also unitarily invariant:
The proof of Theorem 1 will be based on a fairly involved mathematical setup.
3. Standard Purification of Isospectral Mixed States
A state is called pure if its density operator has rank 1. In quantum mechanics, especially quantum information theory,
purification refers to the fact that every density operator can be thought of as representing the reduced state of a pure state. More precisely, if
ρ is a density operator acting on
, and
is a Hilbert space of large enough dimension, then there is a normalized ket
in
such that
ρ is the partial trace of
with respect to
. By the
standard purification of density operators on
of rank at most
k we will mean the surjective map
defined by
. Here,
is the space of linear functionals on
and
is the unit sphere in
. If we canonically identify
with the space
of linear maps from
to
, equipped with the Hilbert–Schmidt inner product, then
Write
for the diagonal
matrix that has
σ as its diagonal, and let
be the space of those Ψ in
that satisfies
when
is expressed as a matrix relative to the standard basis in
. Then
is a subspace of the unit sphere in
, and the standard purification in Equation (
5) restricted to
is a principal fiber bundle over
with right acting gauge group
,
consisting of those unitaries in
that commutes with
. The following two special cases are well known.
Example 1. If , then is the complex projective n-space, is the -dimensional unit sphere, and π is the generalized Hopf bundle.
Example 2. If , then is the Grassmann manifold of k-planes in , is the Stiefel manifold of k-frames in , and π is the Stiefel bundle.
The vertical and horizontal bundles over
are the subbundles
and
of the tangent bundle of
. Here
denotes orthogonal complement with respect to the Hilbert–Schmidt product. Vectors in
and
are called vertical and horizontal, respectively, and a curve in
is called horizontal if its velocity vectors are horizontal. Recall that for every curve
ρ in
and every
in the fiber over the initial operator
, there is a unique horizontal lift of
ρ to
that extends from
[
7] (page 69, Prop 3.1). For convenience, we assume that all curves in this paper are defined on a common unspecified interval
. Moreover, we assume that they are piecewise smooth.
The infinitesimal generators of the gauge group action in Equation (
6) yield canonical isomorphisms between the Lie algebra
of
and the fibers in
. The Lie algebra consists of all anti-Hermitian
matrices that commutes with
, and the isomorphisms are
Furthermore,
is the kernel bundle of the gauge invariant mechanical connection form
, where
and
are the locked inertia tensor and moment map, respectively,
Using Equation (
8) we can derive an explicit formula for the connection form. Indeed, if
are the multiplicities of the different eigenvalues in
σ, with
being the multiplicity of the greatest eigenvalue,
the multiplicity of the second greatest eigenvalue,
etc., and if for
,
then
for every
X in
and every
ξ in
. Hence
Observe that the orthogonal projection of
onto
is given by the connection form followed by the identification in Equation (
7). Therefore, the vertical and horizontal projections of
X in
are
and
, respectively. We finish this section with a discussion on the distance between the fibers of
π over two given density operators
and
in
.
Consider the space
of piecewise smooth curves that start in the fiber
and end in the fiber
. This space can be given a natural smooth structure such that the tangent space at a curve Ψ consists of all smooth vector fields
χ along Ψ that are vertical at the end points of Ψ. Let
E be the energy functional on
,
The differential of
E at Ψ is given by
where
denotes the covariant derivative of
along Ψ. We call Ψ an extremal for
E if
. Clearly, extremals for
E are geodesics:
.
The length of a curve Ψ in
is
Moreover, the distance between
and
is defined as the infimum of the lengths of all curves in
. There is at least one curve in
whose length equals the distance between the two fibers. This is because the fibers are compact. In addition, each such curve is an extremal for
E. Therefore they are horizontal:
Proposition 2. If Ψ is a geodesic in , then is constant.
Proof. Choose any
η in
and consider the variation
. We have that
since
acts through isometries. Hence
for any
. We conclude that
is constant. ☐
4. Properties of the Dynamic Distance Measure
In this section we prove that equals the distance between the fibers and . Theorem 1 follows easily from this observation.
Proposition 3. Suppose ρ solves Equation (1). Let Φ be a horizontal lift of ρ. Then . Moreover, if . Lemma 4. We have that for every ξ in , and if and only if ξ is a constant multiple of the identity.
Proof. Write , where δ is a real diagonal matrix and U belongs to . We have that and since commutes with U. Moreover, , and if and only if δ is a constant multiple of the identity. This is because is strongly convex. ☐
Proof of Proposition 3. Let Φ be a horizontal lift of
ρ, and let Ψ be any lift of
ρ such that
and
. Then
where
is the positive time-ordered exponential. Now,
and
Hence
by Lemma 4. We conclude that
. Moreover, if Ψ is horizontal, and thus
, then
according to Equation (
16). In this case
. ☐
Corollary 5. The dynamic distance between two density operators and in equals the distance between the fibers and .
Proof. Immediate from Proposition 3 and the fact that every curve in is the solution to some Schrödinger equation. This is because the unitary group of acts transitively on .
Proposition 3 and Corollary 5 shows that for any pair of density operators and in , there is a curve Ψ in that extends from the fiber over , ends in the fiber over , and for which . This observation makes the proof of Theorem 1 fairly straightforward:
Proof of Theorem 1. The function
D is positive because
is always a non-negative number by Equation (
2). Moreover,
D is non-degenerate. Indeed, let Ψ be a curve in
such that
. If
, then Ψ is stationary, and hence
. The opposite implication is obvious.
To see that D is symmetric let Ψ be a curve like the one in the proof of non-degeneracy. Define Φ by . Then Φ is a horizontal curve that projects onto a curve in from to . Consequently, . An identical argument shows that . Thus, D is symmetric.
Finally, to see that
D satisfies the triangle inequality let
be a horizontal curve in
covering a curve in
from
to
,
. Also assume that
. Then
is a horizontal curve connecting the fibers over
and
. Therefore,
Thus,
D satisfies the triangle inequality. ☐
6. Relation Between the Dynamic Distance Measure and Bures Distance
Suppose
is
n-dimensional. Let
be the space of all invertible maps in
with unit norm, and
be the space of all invertible density operators acting on
. Then
defined by
is a
-bundle, which we call
Uhlmann’s bundle since it first appeared in [
6]. The geometry of Uhlmann’s bundle has been thoroughly investigated, and it is an important tool in quantum information theory, mainly due to its close relationship with the Bures distance measure [
8,
9].
Uhlmann’s bundle is equipped with the mechanical connection, which means that the horizontal bundle is the orthogonal complement of the vertical bundle with respect to the Hilbert–Schmidt inner product. Moreover, the Bures distance between two density operators in
equals the distance between the corresponding fibers of Π, see [
9]. We denote the Bures distance measure by
.
Suppose
σ has length
n. Then
is a submanifold of
. Moreover, the vertical bundle of
is subbundle of the restriction of the vertical bundle of
to
. However, no nonzero horizontal vector in Uhlmann’s bundle is tangential to
. To see this, let Ψ be any element in
. Then
X in
is horizontal,
i.e., is annihilated by the mechanical connection of the Uhlmann bundle, if and only if
see [
6]. On the other hand, every
X in
satisfies
since
. Clearly, only the zero vector satisfies both Equation (
26) and Equation (
27).
The distance between
and
in
is never smaller than the Bures distance between them. Indeed, every curve between
and
in
is a curve between
and
in
, and since the metrics on the total spaces of the two bundles are induced from a common ambient metric we can conclude that
Uhlmann [
9] and Dittmann [
10,
11] have derived explicit formulas for the Bures distance for density operators on finite dimensional Hilbert spaces. For density operators on
the formula reads
We use this formula to show that there are density operators
and
acting on
for which the inequality in Equation (
28) is strict.
Suppose
, let
, and define a curve Ψ in
by
Set
and
. Then, for
ε small enough, the length of Ψ equals
. In this case,
. However, Equation (
29) yields