In the following, we shall mainly focus upon the time-evolution in the Schrödinger picture where the dynamics of the open system S is given by linear maps on the state space , that is on the convex set of density matrices, that are positive matrices of trace one, whose eigenvalues are interpreted as probabilities and fix the system S statistical properties.
The standard approach [
1,
2,
3,
13] to the description of the dynamics of
S in the presence of the environment
E starts from a factorised initial state of the systems
, namely
, where
is any possible system
S density matrix and
is an environment equilibrium reference state such that
. The compound initial state
evolves in time into
according to the Liouville-von Neumann equation
The state of
S at time
t is obtained by partial tracing over the environment degrees of freedom,
yielding a linear dynamical map
formally generated by an integro-differential master equation [
1,
2,
3,
13]
where the kernel
acts linearly on the state space
and depends on the full Hamiltonian dynamics
integrated over the past of
S.
Not much can in general be said about the linear map
solution to (
3), except that it must be Completely Positive [
1], namely that, besides preserving the positivity of any initial state
of
S, when extended to the dynamics
of the system
, that is to the system
S coupled to an identical inert ancilla
S, it also preserves the positivity of any density matrix in
. In the expression above, “
” denotes the identity operation on the algebra
of
complex matrices associated with the Hilbert space
of the system
S.
Dynamical maps
must be positivity preserving, in short Positive, as they must turn density matrices into density matrices leaving intact the positivity of their spectrum which must, at each positive time
, amount to a probability distribution. Complete Positivity instead refers to the fact that the extended dynamics
must preserve the positivity of all states of the compound systems
. The reason why Positivity of the map is not enough is that, without Complete Positivity, there might be entangled states of
that develop negative eigenvalues under the action of
[
14]. The fact that the solutions of (
3) are Completely Positive stems from the maps
being expressible in Kraus-Stinespring form [
1],
which identifies Completely Positive maps. In the above expression, the matrices
depend on the whole compound dynamics of open quantum system
S and environment
E between times 0 and
t, while the last equality guarantees the maps
to be trace-preserving:
for all
.
2.1. Quantum Dynamical Semigroups
Because of its dependence on the system
S past evolution, the linear map
in (
4) cannot in general fulfil the forward-in-time semigroup composition law
Such a property can be enforced enforce by a so-called Markovian approximation whereby in (
3)
is replaced by
and
by a time independent linear map
, so that the family of true, but analytically uncontrollable, Completely Positive maps
,
, solutions to (
3), is substituted by a one-parameter semigroup of linear maps
solutions to time-independent master equation
Only specific approximation procedures like the weak and strong coupling limits and the low density limit [
1], guarantee that the approximating maps
are Completely Positive: in general, one easily ends up with a not even Positive dynamics [
2,
3,
14,
15]. Instead, under the above mentioned mathematically rigorous approximations, the generator
has the so-called GKSL form in agreement with the renown theorems of Gorini-Kossakowski-Sudarshan [
16] and Lindblad [
17] about the generators of semigroups of Completely Positive, trace-preserving maps
.
Theorem 1. A semigroup , , on the state space of an open quantum system S consists of Completely Positive maps if and only if the generator readswhere the matrices are traceless and orthogonal with respect to the Hilbert-Schmidt scalar product , so that, together with , they constitute an orthonormal basis in , while the complex coefficients form a Positive semi-definite (Kossakowski) matrix . The following example discusses the physical consequences of Complete Positivity.
Example 1. Consider one qubit (2-level system) undergoing a purely dissipative time-evolution generated by the master equationwhere , with initial condition . By means of the Pauli matrices and the identity matrix , in the so-called Block representation, one writes the initial density matrix aswhere the condition on is necessary and sufficient for the positivity of the spectrum of ρ. Since , the Pauli matrices are eigenvectors of the generator:Then, the semigroup solution to (8) is such thatwhence the solution to (8) readswhence positivity of at all times is equivalent to . Whether the maps are also Completely Positive depends on whether the so-called Choi-matrixis also positive [18], whereis the projection onto the symmetric entangled two-qubit stateNotice that describes the time-evolution of the initial entangled two-qubit state under the dynamical maps . One computes that, at small times , the determinant of ,and can thus be positive if and only if . The condition with guarantees the positivity of , but it makes a negative eigenvalue appear in the spectrum of the time-evolving two-qubit state state which could not then be any longer considered a proper density matrix. The condition is indeed the only way to comply with the positivity of the Kossakowski matrix associated to the generator in (8) and thus, as demanded by Theorem 1, with Complete Positivity. Through the positivity of the Kossakowski matrix
(see (
7)), Complete Positivity puts constraints on the dynamics of the system described by
; in particular, in the above example the parameter
must be non-positive, while mere Positivity of
only asks
not to be larger than
. Complete Positivity is sometimes rejected as an unphysical mathematical simplification [
19] because the constraints it imposes upon the dynamics follow from a hypothetical initial statistical coupling of the open quantum system
S with an uncontrollable, dynamically inert copy of itself whose only role is to allow for initial entanglement between them. A more physically palatable ground for the necessity of Complete Positivity emerges when one considers the dynamical maps
on
, namely, when the physical context is one where both system and ancilla are under control and in independent weak interaction with a same environment so that they both evolve according to the same map
. Then, the following result holds [
20].
Theorem 2. is Positive on , if and only if is Completely Positive on .
The physical implications of the above theorem are the following: if
is to describe a physical time-evolution, it must be Positive for all
, otherwise negative probabilities may appear in the course of time when one starts with initial bipartite entangled states of
. Then, according to the above theorem, necessarily the single system dynamics
must be Completely Positive and not only Positive. Notice that if all initial states of
were separable,
, with
,
, and
, then the Positivity of
would suffice for physical consistency; indeed,
This shows that, far from just being a mathematical nicety, Complete Positivity is the dynamical
alter ego of quantum entanglement.
2.2. Non-Markovian Quantum Dynamics
The semigroup composition law (
5) follows from the time independence of the generator
in (
6) as results from suitable Markovian approximations operated on the integro-differential equation (
3). Let us now proceed without such approximations, taking the equation (
3) at its face value. Suppose the dynamics
generated by (
3) to be invertible as a linear map, then the time non-local equation can be recast into the time local form
with a time-dependent generator
. Explicitly time-dependent master equations of the form
are at the basis of the theory of non-Markovian open quantum dynamics [
5,
6,
21,
22,
23,
24]. Formal solutions to such equations, with initial condition
at
read
by means of the time-ordered exponentials
The maps
form a two parameter semigroup:
Setting
, the family of dynamical maps
,
, is automatically
divisible in the sense that for all
there exists an intertwining operator mapping
into
; indeed,
Remark 1. The inverse of a Completely Positive map Λ
is Completely Positive if and only if with U unitary [3]. In general, is not Completely Positive and neither is . However, if is Completely Positive for all such are also the dynamical maps . Nonetheless, as we shall see, the Complete Positivity of for all does not require the intertwining maps to be Completely Positive. Furthermore, unlike for one-parameter semigroups with time-independent generator, there have not yet been found necessary conditions on the time-dependent generator , as those provided by Theorem 1, ensuring the Complete Positivity of the generated dynamical maps . The following theorem fixes the form of the generators
when they provide Completely Positive intertwining maps
for all
[
21].
Theorem 3. The solutions to the time-dependent master equation (19), for all , are Completely Positive maps if and only if the generator readswhere the matrices are as in Theorem 1, while is a time-dependent Hamiltonian and a time-dependent positive semi-definite Kossakowski matrix. The previous result states that a positive semi-definite, time-dependent Kossakowski matrix is equivalent to the Complete Positivity of all intertwining maps , , associated with ; however, as commented in Remark 1, the latter property is not necessary for the Complete Positivity of , . Therefore, for all implies, but it is not implied by, being Completely Positive for all .
The following example provides a simple model of qubit dynamics that allows to discuss some of the most prominent aspects of non-Markovianity; in particular, the fact that another property of semigroups fails general one-parameter families of maps
,
. Namely, unlike in Theorem 2, there exist families of dynamical maps
solutions to (
19) such that
is Positive without
being Completely Positive.
Example 2. Consider the following 1-qubit master equation with time-dependent generator [7]:By comparison with (8), the generator is characterised by a time-dependent Kossakowski matrixwhich is Positive semi-definite only at , while for . Therefore, according to Theorem 3, the intertwining maps can never be Completely Positive; nevertheless, the generated maps are Completely Positive only for while they are Positive for . Indeed, as in Example 1,Such an action corresponds to being expressible very similarly to the Kraus-Stinespring decomposition in (4):with coefficientsThe maps are trace-preserving for all ; indeed, . Furthermore, by means of the convexity of the function for one checks that is negative for all and Positive if . Thus, the expression (28) reduces to (4) and is Completely Positive only for and . On the other hand, is Positive for all ; indeed, in the Block representation (9),whence the Block vector belongs to the unit sphere in and is Positive for all so that is a legitimate 1-qubit state for all . Let us now consider the tensor product dynamics which describes two qubits subjected to identical and independent interactions with their environment. By means of (28), its action on two qubit states can also be cast in a Kraus-Stinespring-like diagonal formHowever, because of , the coefficients , , are negative for . Thus, is Completely Positive when and only when or , namely when and only when is Completely Positive. Further, a result in [25] ensures that, for trace preserving qubit maps , the Positivity of the tensor product maps, , on two qubits is equivalent to the Complete Positivity of the squares, , of the 1-qubit maps. Since from (27) it follows that acts as by changing α into , is Positive for . Thus, for , and are both Positive but not Completely Positive. Finally, one easily computes the algebraic inverse of as a linear map on ,The intertwining maps thus read and are such thatwhere We have thus seen that, unlike for one-parameter semigroups,
can be Positive with
being not Completely Positive; however, similarly to Theorem 2, the following results holds for the intertwining maps
[
7].
Theorem 4. Given a one-parameter family of maps , , with intertwining maps , , then the tensor product maps on are Positive if and only if the maps are Completely Positive.
Theorem 4 has an important consequence in relation to Example 2: it excludes that the tensor products
of the maps in (
30) could be Positive for all
when
. Indeed, if
were Positive, then the single system maps
would be Completely Positive and thus such would be also the single system dynamical maps
, but this is impossible for the
in the interval considered.
The qubit time-evolution in the previous example is also an instance of a dynamics
which is fully legitimate, namely Completely Positive, for
, but non-Markovian according to the so-called
divisibility criterion that we are now going to discuss (se also Ref. [
3,
24,
26]).
Definition 1. A one-parameter family of Completely Positive dynamical maps on is called CP-divisible, respectively P-divisible, if for all , with Completely Positive, respectively Positive, intertwining maps . The one-parameter family , , is called Markovian by divisibility if and only if it is CP-divisible.
Because the generator
in Example 2 has a non-positive definite Kossakowski matrix, according to Theorem 3, the dynamical maps
studied there cannot be CP-divisible. However, they are always P-divisible. Indeed [
3,
27], trace-preserving maps
are Positive if and only if they are contractive on self-adjoint operators with respect to the trace norm
For self-adjoint operators
,
, where
are the positive orthogonal parts of
and for contractive
it holds that
Then, P-divisibility of invertible maps
becomes equivalent to [
28]
This condition is satisfied by the map
in the example: indeed, given
, using (27) with real
and
,
, the eigenvalues of
are
Since
,
is either constant,
, or it decreases in time,
for
.
On the other hand, in view of the last comment in Example 2, the double tensor products
are no longer P-divisible, whence P-divisibilty as much as Positivity is not a property which is in general stable under tensorisation. This is in contrast to Complete Positivity and CP-divisibility: if
is Completely Positive or CP-divisible, that is if
with
Completely Positive for all
, then both
and
are also Completely Positive. In fact, the Kraus-Sinespring form (
4) which identifies Completely Positive maps is robust against tensor products
.
The various properties of the maps and intertwining maps introduced in Example 2 are resumed in the following table (
Table 1).