1. Introduction
In this introduction we shall present a general discussion with detailed definitions and results given in later sections. A coarse-graining of an observable
A is an imprecise version of
A [
1,
2]. Generally speaking, coarse-graining means a reduction in the statistical description of a system. If
A and
B are observables, we say that
B is a coarse-graining of
A if the probability distribution of
B is an affine function of the probability disctribution of
A. A specific type of coarse-graining is when
B is an unsharp or fuzzy version of
A [
3,
4,
5].
In
Section 2, we discuss a coarse-graining of probability measures that involves stochastic kernels. Since probability measures are the classical counterparts of quantum states, we can consider this as pertaining to coarse-graining in classical physics. This work is applied in
Section 3 to studying coarse-graining of quantum observables and instruments.
Section 3 begins with an application for the dynamics of a quantum system described by a strongly continuous unitary group. We then discuss parts and discretizations of observables and show that any two discretizations of observables coexist.
Section 4 discusses finite observables and shows that in this case coarse-graining is the same as post-processing [
5,
6]. Rank 1, sharp and atomic observables are considered next. Sequential products and conditioned observables are treated. The concepts of this section are illustrated with the example of finite position and momentum observables where a crucial role is played by the finite Fourier transform.
Section 5 is more speculative than the previous sections and we do not arrive at many definite conclusions This section discusses symmetric informationally complete (SIC) observables. An important unsolved problem is whether SIC observables exist for every finite dimensional Hilbert space [
5]. It is not even known whether high dimensional SIC observables exist. Applying some of the work in
Section 4, we propose a possible method for attacking this problem. Unfortunately, we have not been able to complete the method and leave this for future work.
2. Coarse-Graining of Measures
We denote the set of probability measures on a measurable space
by
. Let
,
be measurable spaces. A map
that satisfies
is measurable for all
and
for all
is called a
stochastic kernel [
5]. If
v is a stochastic kernel, define
by
. We call
v the stochastic kernel for
V and we say that
is a
coarse-graining of
. We think of
as an imprecise version of
on
. Notice that
V is an affine map because if
,
, then
for all
. Moreover, if
is measurable, then
Example 1. The map is a stochastic kernel from to . The corresponding coarse-graining map satisfies for all . Hence, so V is the identity map.
Example 2. Let
and let be defined by for all , . Then v is a stochastic kernel and the corresponding coarse-graining map isHence, V is the constant map .
We define the Dirac measure at x on by where if and only if .
Lemma 1. (a) If is a stochastic kernel for then .(b) If has a stochastic kernel v, then v is unique.
Proof. (a) If
v is a stochastic kernel for
V, then
(b) follows from (a). □
It can be shown that an arbitrary affine map need not have a stochastic kernel and hence need not be a coarse-graining. One way to accomplish this is to construct such a map V where is not measurable for some . We leave the details of this to the reader. Then V does not have a stochastic kernel v because if it did, then by Lemma 1(a), so is not measurable for some which is a contradiction.
Let , , be measurable spaces and let , be stochastic kernels. Define by . Then is a stochastic kernel.
Lemma 2. Let and be coarse-grainings with corresponding stochastic kernels . Then their composition has stochastic kernel .
Proof. For all
,
we have that
Hence, the stochastic kernel for is . □
We say that
is
part of
if there exists a measurable function
such that
for all
. Define
by
. Thus,
is part of
, if and only if
for a measurable function
. Notice that
is affine because
and hence,
.
Lemma 3. A map is the stochastic kernel for if and only if for all , .
Proof. If
, then
v is a stochastic kernel and
for all
,
. Hence,
v is a stochastic kernel for
. Since stochastic kernels are unique, the converse holds. □
Lemma 4. If and are measurable, and the stochastic kernel for is .
Proof. For all
and
we have that
Hence,
. It follows from Lemma 3 that the stochastic kernel for
is
We say that two probability measures coexist, if they are both parts of another probability measure.
Lemma 5. If , then coexist.
Proof. Define
by
and define
by
,
by
. Then
f and
g are measurable and if
we obtain
Hence, is a part of . Similarly, if , then so is a part of . □
Let
be a measurable space and let
be a finite measurable space with
. Let
be a
measurable partition of
. That is
for
and
. Define
by
. Then
V is affine because
so that
. We call
V a
discretization map and
a
discretization of
. A stochastic kernel
is called a 0 –1
stochastic kernel if
for all
.
Theorem 1. An affine map is a discretization if and only if V has a 0 –1 stochastic kernel.
Proof. Suppose
V is a discretization and
V has stochastic kernel
. Then by Lemma 1 we obtain for
that
Hence,
for all
,
. To show that
is actually the stochastic kernel for
V we have that
Of course,
is a 0 –1 stochastic kernel. Conversely, suppose
is a 0 –1 stochastic kernel for
Then
for all
,
. Let
,
, be the measurable sets
If
for
, then
and
which is a contradiction. Hence,
for
. If
and
for all
, then
which is a contradiction. Hence, there exists an
i such that
so
. We conclude that
is a measurable partition of
. Since
we have for all
that
We conclude that V is a discretization map. □
When we consider a finite measurable space
we always assume that
so
need not be specified. For
we identify a
with the column vector with entries
where we write
,
. An
matrix
is a
stochastic matrix if
and
for all
. In this finite case, the stochastic kernels are replaced by stochastic matrices. This is because, in the finite case, if
is a stochastic kernel, then
is a stochastic matrix and conversely, if
is a stochastic matrix, then
is a stochastic kernel.
Theorem 2. Let , and let be affine. Then there exists a unique stochastic matrix such that for every we have . Conversely, if M is an stochastic matrix, then there exists an affine map such that .
Proof. Let
be affine. Since every element of
is a convex combination of
,
we have that
where
and
,
. We conclude that
is an
stochastic matrix and
. Letting
we obtain
where
,
. Since
,
and
V is affine, we conclude that
To show that
is unique, suppose
where
is an
matrix. We then obtain
Conversely, let be an stochastic matrix. Define by , and extend V affinely to all of . By our previous work, . □
We conclude that in the finite case, every affine map is a coarse-graining and is implemented by a unique stochastic matrix . We then identify V and .
3. Observables and Instruments
In this section, we employ our previous work to study coarse-graining of observables and instruments. Let
H be a complex Hilbert space that represents a quantum system
S. We denote the set of bounded linear operators on
H by
. For
, we write
if
for all
. An operator
is an
effect if
where
are the zero and identity operators, respectively. We denote the set of effects by
and interpret an
as a 1– 0 (true-false) measurement [
5,
7,
8]. If
is a measurable space, an
observable with
outcome space is an effect-valued measure
[
5,
7,
8]. That is,
when
,
, and
. We interpret
as the effect that occurs when a measurement of
A results in an outcome in
. A
state for
S is an effect
that satisfies
. We denote the set of states on
H by
. If
,
we interpret
as the probability that
E occurs (is true) when
S is in the state
. If
A is an observable, its statistics in the state
is given by the
distribution
for all
. Of course,
for all
[
5,
7,
8].
We now discuss a method for constructing stochastic kernels from observables. Let
,
be measurable spaces,
a collection of states and
A an observable with outcome space
. We say that
is
measurable if
is measurable for all
. If
is measurable, we define the stochastic kernel
with the corresponding coarse-graining
If
are pure states
,
, then (
1) and (
2) become
and
We interpret (
1) as the probability that a measurement of
A results in an outcome in
when
S is in the state
.
Example 3. Let be a finite measurable space. We show that any stochastic matrix , can be written in the form of the previous paragraph. Let H be a complex Hilbert space with dimension n and let be an orthonormal basis for H. Let A be the observable with outcome space Ω
satisfying Letting be the pure state , , we obtain We now give an application of the previous structure to the study of the dynamics of the system
S. Suppose the dynamics of
S is described by the strongly continuous unitary group
,
, where
K is the Hamiltonian for
S. If
is the initial state, then
is the state at time
. We can consider
as a collection of states indexed by the points of the measurable space,
. Let
A be an observable with outcome space
. Since
is continuous we have that
is continuous for all
. It follows that
is measurable. We conclude that the map
given by
is a stochastic kernel called the
dynamical kernel for
. We interpret
as the probability that a measurement of
A at time
t results in an outcome in
. In terms of the dynamical group we have
The observable
which gives the time evolution of
A is the
Heisenberg picture of quantum mechanics while (
5) gives the
Schrödinger picture. The corresponding coarse-graining map
satisfies
For a particular time
we have
Let
A be an observable with outcome space
and let
be a measurable space. If
is a stochastic kernel, we define the observable
with outcome space
by
We call
v the
stochastic kernel for
V and
is a
coarse-graining of
A [
5,
7]. We see that
is the unique effect satisfying
for all
. We now show that this idea extends to observables.
Lemma 6. is the unique observable with distribution Proof. For all
,
we obtain
The observable
is unique because two observables on
H with the same distributions for every
are identical [
5,
7,
8]. □
If
,
, are observables on
H with the same outcome set and
,
, it is clear that
is again an observable. Thus, such observables form a convex set. We conclude that
is an affine map because
Let
,
be measurable spaces and let
be measurable with corresponding stochastic kernel
and coarse-graining
given by (
1) and (
2). If
B is an observable with outcome space
we obtain the following result.
Lemma 7. (a)
For all we have that(b)
For all , , we have that Proof. (a) Since
for all
, we obtain
(b) For all
,
, applying (a) we obtain
□
An observable
B is
part of an observable
A if there exists a measurable surjection
such that
[
6,
9,
10].
Lemma 8. Let be observables on H with outcome spaces , , respectively. Then B is part of A if and only if there is a measurable surjection such that for all .
Proof. If
B is a part of
A, there exists a measurable surjection
such that
. If
is the corresponding stochastic kernel, then for
we obtain
Conversely, if for all , then letting we obtain by reversing the previous argument. Hence, so B is part of A. □
By Lemma 6 if
B is part of
A so that
, then
and hence
is part of
for all
. Two observables
coexist if there exists an observable
A such that
B and
C are part of
A [
5,
7,
11,
12]. It is well-known that unlike in Lemma 5, two observables need not coexist [
5,
7,
12]. Let
A be an observable with outcome space
. If
V is a discretization of
, we call
a
discretization of
A [
5]. If
is the corresponding stochastic kernel we obtain
Lemma 9. If is a discretization of A, then is a part of A.
Proof. Let
where
so the outcome space of
is
. Let
be the corresponding stochastic kernel. Define
by
if
. Then by (
7)
and it follows that for all
we obtain
Hence, is part of A. □
Corollary 1. Any two discretizations of an observable coexist.
Let
be the set of trace-class operators on
H. An
operation on
H is a trace non-increasing, completely positive linear map
[
5,
7,
8,
11]. If an operation
T preserves the trace, then
T is called a
channel on
H. An
instrument on
H with
outcome space is an operation-valued measure
on
such that
is a channel. The statistics of an instrument
for a state
is given by its
distribution
for all
. Of course,
is a probability measure on
. We say that an instrument
measures an observable
A if
and for all
and
we have
It can be shown that an instrument measures a unique observable, but an observable is measured by many instruments [
5]. If
measures
A we write
. We think of
as an apparatus that can be employed to measure the observable
and conclude that there are many such apparatuses. Although
reproduces the statistics of
,
gives more information than
. This is because when a measurement of
produces a result in
the instrument
updates the state of the system to the new state
when
[
5,
7,
8].
If
is an instrument on
and
is a stochastic kernel, then we shall show that
is an instrument with outcome space
called a
coarse-graining of
. To show this we have that
is countably additive on
and
so
is a channel. Moreover, if
we obtain
It follows that is an instrument. It is easy to check that instruments form a convex set and that is affine.
Theorem 3. (a). (b)For instruments we have that for all if and only if .(c)If , then for all .
Proof. (a) For all
we obtain
It follows that
. (b) If
, then for all
we have that
Therefore,
for all
so
. Conversely, if
, then for all
we obtain
Hence, . (c) If , then by (a) . Applying (b) gives for all . □
The converse of Theorem 3(c) does not hold. That is, if for all , we need not have . For example, let the identity map. Then for all . However, there exist with for all so . Applying Theorem 3, we can consider the various special types of coarse-graining for instruments.
4. Finite Observables
In this section, we restrict our attention to finite observables. If
A is an observable with
, then
A is completely determined by
We then define
and write
. It follows that for all
we have that
. Let
be another observable and let
be an affine map. We write
if
for all
. We then say that
B is a
post-processing of
A [
5,
6]. Thus, post-processing is the same as coarse-graining for finite observables.
Theorem 4. If is affine, then if and only if for all where is the stochastic matrix corresponding to V.
Proof. Suppose
is affine and
. By Theorem 2
is a stochastic matrix and for all
we obtain
It follows that
. Conversely, suppose
for all
. Then for all
and
we obtain
Hence, . □
We can identify an observable with a set
satisfying
. We say that
A is
rank 1,
sharp,
atomic, respectively, if
are rank 1, projections, 1-dimensional projections. If
A is sharp, it follows that
for
[
5,
6]. If
A is atomic, there exists an orthonormal basis
for
H such that
,
. Notice that
is rank 1 if and only if
where
and
P is a 1-dimensional projection.
If
,
are observables on
H, their
sequential product is the observable with outcome space
given by [
6,
13].
We also define the observable
Bconditioned by the observable
A as
It can be shown that
coexists with
A [
6]. If
is a stochastic matrix of the appropriate size, then
and if
is a stochastic matrix of the appropriate size, then
Notice that (
9) is much more complicated than (
8). If
A is sharp, then (
8) and (
9) become
and
If
A and
B are atomic with
and
then (
8), (
9) become
and
Notice from (
12) and (
13) that both
and
are rank 1 observables. The next lemma shows that post-processing and conditioning interact in a regular way.
Lemma 10.
Proof. The result follows because
Hence, . □
Example 4. This example illustrates the concepts of this section in terms of finite position and momentum observables. Let H be a finite-dimensional Hilbert space with dimension d and let be an orthonormal basis for H. The finite Fourier transform is the unitary operator on H given by
where
[
5]. Equivalently, F is the operator satisfying
for all
. We call
where
the
finite position observable and
where
with
the
finite momentum observable. Notice that
,
and
. We also see that Q and P are atomic observables. The observable
has effects
Thus, is a rank 1 observable and is the trivial observable for . In a similar way, and for . The distribution of Q in the state becomes for .
More interesting observables are obtained by post-processing. Let
be a stochastic matrix so that
and
, for all
. Then the post-processing observable
satisfies
We see that the eigenvalues of
are
,
with corresponding eigenvectors
. The distribution of
in the state
becomes
The observable
satisfies
Equation (
14) also follows from (
13).
5. SIC Observables
This section is more speculative than the previous ones and we do not come to many definite conclusions. A finite observable
A is
informationally complete (IC) if
for all
implies that
. Equivalently,
A is informationally complete if
implies that
. It can be shown that there exist IC observables for every finite dimensional Hilbert space
H [
5]. An Observable
A on a Hilbert space
H with
is
symmetric if [
5]:
- (S1)
,
- (S2)
A has rank 1,
- (S3)
for all ,
- (S4)
for all .
It can be shown that
is the smallest cardinality for the outcome space of an IC observable [
5]. Furthermore,
if
is constant for all
and
for
if
is constant for
[
5]. A symmetric IC observable is called a SIC
observable. An important unsolved problem is whether SIC observables exist for every finite dimensional Hilbert space [
5]. It is not even known whether high dimensional SIC observables exist. We would like to propose a possible method for attacking this problem. Unfortunately, we have not been able to complete this method and we leave this to future work.
Let
and let
,
be atomic observables. For a
stochastic matrix
we define the observable
. For example,
of Example 4 is such an observable. Letting
be the vector given by
We conclude from (
13) that for all
we have that
It immediately follows that
C satisfies (S1) and (S2). We say that a stochastic matrix
is
doubly stochastic if
for all
y [
5]. The bases
,
are
mutually unbiased bases (MUB) if
for all
[
6]. It is easy to show that there exist pairs of MUB for every finite dimension. In fact, the two bases in Example 4 are MUB.
Theorem 5. (a)If ν is doubly stochastic and , are MUB, then for all .(b) If for all then ν is doubly stochastic.
Proof. (a) Applying (
15), (
16) we have that
for all
. If
is doubly stochastic and
,
are MUB we conclude that
for all
. (b) If
for all
, then by (a) we obtain
for all
. Summing over
y gives
for all
x, so
is doubly stochastic. □
In Theorem 5b, if for all , then , need not be MUB so the converse of Theorem 5a does not hold. For example, suppose for all . Then for all but , can be arbitrary bases. We conclude from Theorem 5a that if is doubly stochastic and , are MUB, then Condition (S3) holds.
Lemma 11. (a)Condition (S4) holds if and only if for all .(b)The observable C is IC if and only if for we have that for all implies that .
Proof. (a) Applying (
16) we have that
and the result follows. (b) Applying (
16) we have that
and the result follows. □
Theorem 5 and Lemma 11 complete conditions under which C become a SIC observable.
We now illustrate our SIC method in the qubit case
. Let
,
be the standard basis for
H and let
be a basis for
H such that
,
are MUB. For example, we could use
of Example 4. Define the atomic observables
,
with
. Let
be the doubly stochastic matrix
. Define the observable
and the effects
Letting
,
, be the vectors defined by (
15), we have by (
16) that
,
,
,
.
We have that
C satisfies Conditions (S1), (S2) and (S3). According to Theorem 5(a),
C satisfies Condition (S4) if and only if
when
. Now
Hence, (S4) is not satisfied.
If , 1 or , it is easy to check that C is not IC. Unfortunately, even when , C need not be IC. For example, let , as before. We then have the following result.
Theorem 6. If and G is a self-adjoint matrix, for all , if and only if where .
Proof. By (
16),
if and only if
for all
. Letting
we conclude that (
17) holds if and only if
Adding (
18) and (
20) gives
and hence,
. If
, then
so
which is a contradiction. Hence,
and it follows that
. Hence,
,
and the result follows. The converse is clear. □
Corollary 2. If , , then C is not IC.
Proof. Define
,
where
. Then
and it is easy to check that
by showing that the eigenvalues of
are
. Since
, it follows from Theorem 6 that
for all
. However,
so
C is not IC. □
It is possible that for other
we obtain an IC observable
C. It is also possible that for higher dimensional spaces we obtain SIC observables using this method. Even though
C is not IC, it satisfies two necessary (but not sufficient) conditions for IC [
5] (Prop 3.35). If
C is IC these conditions are: (a)
does not have both eigenvalues 0,1 and (b) for all
there exists
such that
Indeed, (a) is clear and (b) follows from the fact that