Simple Advanced
Simple Advanced
in Quantum Mechanics
M. Fannes
Instituut voor Theoretische Fysica
K.U.Leuven
March 2013
Contents
1 Introduction
1.4 Teleportation . . . . . . . . . . . . . . . . . . . . . . . . . . .
9
9
2.2 States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.2.1
Density matrices . . . . . . . . . . . . . . . . . . . . . 12
2.2.2
Convex sets . . . . . . . . . . . . . . . . . . . . . . . . 14
2.2.3
2.2.4
25
Unitary gates . . . . . . . . . . . . . . . . . . . . . . . 29
3.3.2
Dissipative operations . . . . . . . . . . . . . . . . . . 31
3.4.1
Generators . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.4.2
Lindblads theorem . . . . . . . . . . . . . . . . . . . . 37
3.4.3
3.4.4
3.5.2
3.5.3
3.5.4
3.5.5
4 Entropy
. . . . . . . . 48
50
Introduction
1.1
General principles
f (j ) | ej , |2 = , f (A) .
=
j
(1)
1.2
The qubit
|0 =
and |1 =
0
.
1
(2)
In the context of quantum information theory this basis is often called the
computational basis.
An arbitrary state vector in
= |0 + |1
is of the form
with ,
(3)
(4)
correspond to antipodal
(5)
Clearly, observing only z does not provide enough information to reconstruct the state vector: there is no way to determine . The variance of
measurement outcomes is given by
2 (z ) = z2
= 1 cos2 = sin2 .
(6)
(8)
1.3
Two qubits
with
Let us consider a very entangled state vector 00 = 12 |00 + |11 and let
us computed the expectation of an observable of the rst subsystem
A
00
= 00 , A 00 =
1
2
0|A 0 + 1|A 1 .
(11)
1
2
0|A 0 + 1|A 1 .
(12)
1.4
Teleportation
Suppose that we want to transmit a qubit state from A(lice)s lab B(ob)s
lab using standard means of communication, that is to say, by transmitting
numerical data from A to B. The state is just any possible unknown
state of a qubit and we are allowed a single use of the system. Measuring
an observable in A will not be very helpful. Indeed, the outcome is random
and after the measurement the state has turned into one of the eigenstates
of the observable. Suppose, however, that beside transmitting numerical
information A and B share a (maximally) entangled state two-qubit state
like 00 above. Sharing a two-qubit state means that A can act on the
composite system consisting of the unknown qubit state and the rst 2
factor of the space 2 2 to which 00 belongs. B on the other hand can
act on the second factor in 2 2 .
10 =
1
2
1
|00 + |11
01 =
|00 |11
11 =
1
2
1
|01 + |10
|01 |10 .
(15)
1
2
1
(16)
0 1
.
1 0
This procedure is called teleportation. It is actually not transporting physical
objects from A to B but rather the structure of an arbitrary quantum state
at A to that of a state at B. Remark also that neither A nor B actually
measure the teleported state. The only eect of the whole procedure is that
an unknown (, ) superposition of two states at lab A has been exactly
reconstructed at lab B.
Problem 8. Work out the actions that have to be taken in Bobs lab to
restore the original qubit for other measurement outcomes of Alice.
7
2.1
Positive matrices
f (1)
0
0
0
f (2)
0
f ..
..
.. .
.
.
.
.
.
.
0
0
f (d)
9
P P = P and
(18)
P .
(19)
a b
c a
b c
c
b positive.
a
2.2
States
We now come to the notion of state. This will encode all the statistical
information that we can obtain about measurements of observables for a
system prepared with innite care following a given procedure. In this sense
a state is a preparation procedure. So a state is an expectation functional
on the observables:
A A .
(20)
Up to now the term state vector was used for a quantum system. A state
vector denes a state through
A
:= , A , A an observable.
It turns out that the following requirements on states t very well the experimental observations and allow, moreover, for a probabilistic interpretation
of the theory:
Definition 1. A state
on an algebra of observables A is a functional that
satises the following requirements
1. A A
ables
Density matrices
f =
pj f (j).
(21)
j=1
pj = 1.
(22)
= , A = Tr | |A , A Md .
(23)
Therefore the one-dimensional projector | | is the density matrix corresponding to the state .
Canonical Gibbs matrices are another important example: given an inverse
temperature > 0 and a Hermitian Hamiltonian H, the canonical equilibrium state has density matrix
eH
with Z = Tr eH .
(24)
Z
The normalization factor Z is called the partition function and it yields the
free energy log(Z)/ of the system.
=
13
2.2.2
Convex sets
There is a general theory for convex subsets of a real vector space that shows
how a compact convex set can be reconstructed in terms of its extreme points.
These results provide us with a number of useful notions about state spaces.
We consider for simplicity only subsets of nite dimensional real vector
spaces. Most results generalise to innite dimensions modulo some additional technical assumptions. A subset C d is convex if px + (1 p)y C
whenever x, y C and 0 p 1. If C d is convex, x1 , x2 , . . . , xk C,
pj 0 and j pj = 1 then also
k
j=1
pj xj C.
(25)
pj xj
j
xj X, pj 0,
pj = 1 .
(26)
The closed convex hull is the closure of Conv(X), it is the set obtained by
adding all limits points.
Let C be a closed convex subset of d . A point c C is called an extreme
point of X if it cannot be written as a non-trivial mixture of points of C. In
formulas: x is extreme if x = px1 + (1 p)x2 with 0 < p < 1 and x1 , x2 C
implies that x1 = x2 . The set of extreme points of a closed convex set is
called the extreme boundary of C and denoted by ext (C).
Problem 13. Find the necessary and sucient conditions on a, b, c to turn
the matrix of Problem 10 in a density matrix. Is the set of density matrices
that you obtain convex? If so, nd its extreme points.
Theorem 1 (Minkowski). Let C be a closed, bounded, convex subset of d ,
then
C = Conv ext (C) .
(27)
14
(28)
(29)
Using this notion one can show that a closed convex set C d dierent from
d is equal to the intersection of all the positive half-spaces that contain C.
I.e. such a set can be characterised by a set of linear inequalities. In general
there will be a lot of redundant inequalities and one can look for a minimal
set, this is called linear programming.
Problem 14. Describe the unit disk in 2 by a set of linear inequalities.
15
2.2.3
This contains all the information that can be gained by measurements about
the particles emitted by the source. The details of the source {(pj , j )} is
called a quantum ensemble. Because the state space of a quantum system
is very dierent from a simplex dierent quantum ensembles can return the
same density matrix and no experiment can discern between such sources.
16
This is very dierent from the classical case where the state space is a simplex
and where therefore a mixed state automatically denes a unique ensemble.
Let us count the number of real parameters needed to describe a d-dimensional density matrix. For a general Hermitian matrix we need d + d(d 1) =
d2 . The positivity condition doesnt change that number but normalisation
removes 1 degree of freedom, hence we need d2 1. The topological boundary
imposes one additional real condition and so we still need d2 2. The points
of the extreme boundary correspond to one-dimensional subspaces of d .
Normalising a vector in such a subspace and multiplying it by a phase we
remain with 2(d 1) real parameters. So we see that the extreme boundary
is in general a much smaller set than the boundary except for a qubit (d =
2). It turns out that in any dimension the pure state space is a very nice
Riemannian manifold. The boundary of the state space is, however, very
complicated and contains many at parts.
Suppose that x is a point on the topological boundary of a compact convex
subset C of d . Such a point need not be extreme but it denes a face of C
F (x) = {y C | z C such that x = y + (1 )z
with z C and 0 < < 1}.
(32)
. The set
{z | z
, |z| = 1}
(33)
vectors in a ray yield the same pure state on Md . It is easily seen that the
correspondence between rays and pure states is actually one to one. The
space of rays in d is sometimes called the complex projective Hilbert
space of dimension d: CP(d). We actually just said that
d
CP(d) = {
| = 1}/U(1).
(34)
Here U(1) are the one-dimensional unitary matrices: the complex numbers
of modulus one.
The usual norm distance in
denote by [] the ray of
d
d
dSF [1 ], [2 ] := min 1 z2 .
zU(1)
(35)
The distance dSF is called the Study-Fubini distance and it is easily computed
(36)
dSF [1 ], [2 ] = 2 2| 1 , 2 |.
Let us consider a small perturbation d of the vectors z belonging to a ray
that does not change the norm to rst order
z + d
= 2 + z , d + z d, + o( d )
= 1 + z , d + z d, + o( d ).
(37)
+ d
= 1 + d
+ d
= + d
+ d
1
2
d 2 + o( d 2 ).
(38)
(39)
ds2 = 2 2 ,
18
(40)
dt
0
d(t)
.
dt
(41)
For two points [1 ] and 2 we can always choose the phases so that 1 , 2
0. The geodesic connecting [1 ] and [2 ] is then the circular arc in the
(1 , 2 )-plane connecting 1 and 2 . Its length is
cos1 | 1 , 2 | .
(42)
2.3
0 1
,
1 0
2 =
0 i
,
i 0
and 3 =
1 0
.
0 1
(43)
1
2
( + x ), x 3 .
(44)
The boundary and extreme boundary coincide in this case. The centre of
the ball is the uniform state A 21 Tr A which is for instance obtained as
an equilibrium state at innite temperature.
Problem 19. Check that the parametrisation of qubit vector states used
in (4) corresponds to the usual parametrisation of the unit sphere in terms
of spherical angular coordinates.
19
To make a full tomography of a qubit state we can perform series of measurements to obtain reliable gures for the expectations of 1 , 2 , and 3 .
Using the Bloch parametrisation (44) we have
xj = Tr j = j .
(45)
From the positivity condition for a qubit density matrix we see that we must
have
1 2 + 2 2 + 3 2 1.
(46)
The closer this expression comes to 1 the purer the state is. A full tomography of more complicated systems, like two qubits, requires many more
measurements and is therefore a costly and lengthy operation.
2.4
Bj =
j
(Bj )
j
for all countable collections {Bj } of disjoint Borel sets. This last property is
called -additivity. The Borel sets are the subsets of that can be given a
probability, they represent the events that can occur.
For a quantum system the projectors play the role of events: the corresponding measurement can only have two outcomes 0 or 1, true or false. The
projectors (on closed subspaces of a Hilbert space) form a lattice. Recall
that a projector P1 is smaller than P2 if P2 P1 is positive denite and this
is equivalent to P1 = P2 P1 . Being a lattice means that for any two projectors
P1 and P2 there is a projector that dominates both, e.g. P1 P2 . There is
also a projector that is dominated by both such as P1 P2 . Actually, the
20
lattice is nite because there is a smallest element, the zero operator, and a
largest one, the identity. The lattice is also orthocomplemented: the join of
P and P is and their meet is 0. One can also show that the lattice is
closed under countable joins and meets. A quantum probability measure can
now be characterized as in the classical case: a function
from the lattice
of projectors to [0, 1] such that 0 = 0, = 1, and
Pj =
j
Pj
j
(47)
A basic dierence between classical and quantum probability is that conditioning doesnt work in the quantum case. Let X and Y be two Borel subsets
of and let Prob(Y ) > 0. The conditional probability of X given Y is
dened to be
Prob(X Y )
Prob(X|Y ) :=
.
(48)
Prob(Y )
Suppose that {Yj } is a partition of in at most a countably innite number
of Borel subsets such that for every j Prob(Yj ) > 0. We can then write for
an arbitrary Borel subset X of
Prob(X) =
Prob(Yj ) Prob(X|Yj ).
(49)
This is often called Bayess law. In the quantum context this doesnt work,
the reason being that for a countable partition {Pj } of the identity, i.e. for
a countably innite family of projectors Pj such that j Pj = and for a
general projector Q one usually has the strict inequality
Q
Q Pj ,
(50)
X Yj .
21
(51)
2.5
(52)
(53)
j|A |k =
or
j|AB |k .
(54)
Problem 20. Show that the partial trace does not depend on the choice of
basis in the space over which the partial trace is taken.
Problem 21. Verify (54).
Given local density matrices A and B then there is always an extension to
the composite system: A B . This corresponds to independence between
both parties. In most cases there will be more possibilities compatible with
the local restrictions. Indeed, for a general bipartite state we have d2A d2B 1
real freedoms while specifying the reduced density matrices consumes only
d2A + d2B 2 parameters.
How strong quantum systems are correlated is partly captured in the notion
of entanglement. A state AB of a bipartite system is called separable if
it is a convex combination of product states
AB =
p A B, .
22
(55)
A state that is not separable is called entangled. By construction, the separable states form a convex set. Deciding whether a given bipartite state is
separable or entangled turns out to be a very dicult problem. Moreover,
a more rened distinction between classes of states seems to be necessary
in order to understand their usefulness for various tasks. These questions a
very dicult and go far beyond the scope of this course. Restricting to pure
bipartite states is, however, quite simple: a pure state is separable if and
only if it is a product of pure states.
Problem 22. Show that the extreme points of the set of separable states
are the pure product states.
There is a simple quantitative characterisation of the degree of entanglement
of a pure state based on the following proposition:
Proposition 6. For any AB dA dB of a composite system there exist
orthonormal families {ej } in dA and {fj } in dB and positive numbers cj
such that
AB =
cj ej fj .
(56)
j
Problem 26. Could one extend the Schmidt decomposition to more than
two parties?
Closely related to the Schmidt decomposition is the purification of a general
state. Let be a density matrix on d , then we can nd an orthonormal
basis {ej } and a probability vector {rj } such that
=
j
rj |ej ej |.
(57)
In fact we can limit the sum to the j with rj > 0. Suppose there are d such
j, the number d is the rank of , it is the dimension of the range space of
=
rj ej fj
(58)
j
rj >0
24
Quantum dynamics
We suppose that we can start the dynamics of a system at some initial time
t0 with an arbitrary initial state 0 of the system. At some later time t the
state of the system is (t; t0 , 0 ). The following general requirements on an
evolution appear to agree well with observations
1. The map 0 (t; t0 , 0 ) is affine: it preserves convex mixtures. This
allows us to introduce a linear evolution map (t, t0 ) on Mk such that
(t; t0 , 0 ) = (t, t0 ) 0 , t t0 .
(59)
(60)
3.1
(61)
1 0 0 1
1 0
0 1
1 0 0 0 0
0 0
= 1 0 0
.
P =
0 0
2 0 0 0 0 2 0 0
1 0 0 1
1 0
0 1
1 0 0 0
1 0 0 1 0
1 0
0 1
1
(id2 T)(P ) =
0
1
0
0
0
1
1 0
2
2
0 0 0 1
which is not positive because of the last term. Hence, we already loose
positivity if we extend the transposition to a single additional qubit.
Let d 0 . A -linear map : Mk Mn can be extended to a -linear
map idd from Md Mk to Md Mn . Any element X Md Mk can
be written as a d d matrix with entries in Mk :
X11 X1d
..
.. , X M .
X = ...
(62)
ij
k
.
.
Xd1 Xdd
We then put
(X11 ) (X1d )
..
.. .
idd (X) = ...
.
.
(Xd1 ) (Xdd )
26
(63)
3.2
Sometimes the term super-operator is used to denote linear transformations of Mk , or linear maps from Mk to Mn . This is just to warn you that
we are not working on the level of the space of wave functions but rather
on that of states or observables considered as elements of the linear space of
transformations of vector states.
There is a standard way of encoding such a super-operator called the Choi
encoding. We start by introducing the standard matrix units eij := |i j|,
these are the matrices with all entries equal to 0 except for a 1 on row i and
column j. Obviously these matrix units form a basis of Mk considered as
vector space and we have
Mk A =
Aij eij .
(64)
ij
27
, then
: X V XV is completely positive.
(66)
Indeed:
idd (A B) = A (B) = A V BV = (d V )(A B)(d V ) .
Now, if C Md Mk is positive then it is of the form D D and
idd (C) = (d V )D D(d V )
= D(d V )
D(d V ) 0.
28
Vj Vj with
() =
j
(67)
Problem 30. Fill out the details of the proof of the theorem and the corollary.
Problem 31. Express the trace-preserving and unity-preserving conditions
on a quantum operation in terms of its Choi matrix.
3.3
Unitary gates
(68)
Unitary gates are single-shot unitaries that are applied to qubit systems
extending operations on bits to qubits or introducing operations that have
no classical counterpart. These gates are intended to represent basic logical
operations performed on systems of qubits.
29
(69)
Because the action of the gate can be expressed at the level of state vectors,
we also have
U(1 + 2 ) = U1 + U2 .
(70)
Hence, unitary gates not only preserve mixtures but also coherences.
Denoting by a classical bit {0, 1} we denote by 1 2 addition modulo
2. A few common classical gates are
NOT :
(71)
AND :
1
1 2
2
(72)
CNOT :
1
1
1 2
2
(73)
FAN OUT :
NAND :
(74)
1
1 1 2
2
(75)
(76)
and U : |1 |0 .
0 1
. A general
1 0
superposition of the computational basis states is then mapped into the same
superposition of the negated basis states. Such extensions of classical gates
to the quantum setting dont really full the expectations one could have.
For example the quantum NOT dened above does not map every qubit state
into its orthogonal complement. FAN OUT would map, modulo neglecting
some outputs, |0 to |00 and |1 to |11 and is therefore a classical copier.
A general qubit state would then be mapped as follows
Dissipative operations
pj pj .
(79)
(80)
32
(81)
Tr =
Tr j |i j| =
i
ij=1
ij=1 a=1
|i ia|
=
a=1
i=1
ia|ja |i j|
(83)
j=1
|ja j| =
Va Va
a=1
where
Va :
: |ib ab |i .
In this way we obtain a Kraus form for the partial trace which proves that it
is completely positive. Combining this example with the three previous ones
we can realise a generalised measurement set-up using a pointer system.
Example 7 (A generalised measurement). A more complicated and
more realistic set-up uses an auxiliary pointer system k : the incoming
state is composed with an initial state of the pointer system. This composite system then unitarily evolves by passing through a unitary gate. Then
a von Neumann measurement, without ltering any particular outcome, is
applied to the pointer part of the system and we are nally left with the
reduction of the resulting state to the system we observe. We compute the
transition from initial to nal state using some simplifying features but it
turns out that the overall result is still a completely general quantum operation.
Let {fj } denote the measurement basis in our pointer system k and suppose
that the initial state of the pointer system is f1 . It is useful to write the
unitary gate U in the given basis of the pointer system
U = Uij
(84)
ij
where the Uij are d d matrices that satisfy the unitarity relations
U
j
ij
Uji Uj = i d .
Uj =
j
33
(85)
Uj1 Uj1 =
Vj Vj .
(86)
j
(87)
11 12
21 22
with
(88)
1 0 0
0 0 0 0
(89)
C() =
0 0 0 0 .
0 0 1
() = + (1 ) ,
2
(90)
Tr .
0
0
0
.
0
(91)
(92)
1+
2
This matrix is positive i 13 1. The extreme value = 13 corresponds to /3. This is the most we can do reversing the sign of all
components of angular momentum without destroying complete positivity.
3.4
Generators
(93)
Even simpler are autonomous evolutions. Here (t, t0 ) only depends on tt0 .
Slightly abusing notation we write
(t, t0 ) = (t t0 ).
(94)
(95)
Even more restrictive are reversible autonomous dynamics where one has a
one-parameter group of trace-preserving completely positive maps
{(t) | t },
(96)
(97)
Markovian dynamics are characterised by their generator, which is in general time-dependent. Let us dene this time-dependent generator as
(t) := lim
0
(t, t ) id
(t + , t) id
= lim
.
0
(98)
(99)
(100)
(t, t0 ) = Texp
ds (s)
t0
= id +
ds1 (s1 ) +
t0 s1 t
t0 s2 s1 t
3.4.2
Lindblads theorem
(102)
(103)
with the explicit form (102) for the generator is called Lindblads equation.
It is the generalisation of Schrodingers equation to general non-reversible
Markovian dynamics. The dierential equation for a state with initial
condition 0 reads
d
= i[H(t), ]+
dt
To derive this result it suces to consider the autonomous case and we need
the general characterisation of positive semi-deniteness for a 2 2 block
matrix.
A C
acting on d1 d2 is positive
Proposition 8. The block matrix
C B
semi-definite if and only if
1. A and B are positive semi-definite and
2. there exists a U :
d2
d1
AU B.
(107)
(108)
C(id) =
i
jj| .
(109)
into
=
i
|ii
|ii
(110)
d + C()11 C()12
.
C()21
C()22
(111)
(112)
Now we can essentially repeat the proof of the Choi-Jamiolkowski-Kraus theorem. The main dierence is that C()22 acts on the orthogonal complement
of the maximally entangled vector i |ii . Combining this with preservation
of the trace leads to (102).
All these arguments can be reversed to show that a super-operator of the
form (102) is the generator of a semi-group of quantum operations.
Problem 35. Fill out the details of the proof of Theorem 4
Before turning to applications we mention the quite useful two-positivity
inequality:
38
(113)
(114)
X X X
.
X
X =
(115)
(X X) (X )
.
(X)
(116)
(117)
X X X
X
The inequality (114) follows from (113) by expanding exp(t ) for small
positive t.
3.4.3
a11 + b22
c12
11 12 =
(118)
21 22
c 21
a11 b22
39
a 0 0 c
0 a 0 0
C() =
(119)
0 0 b 0 0 on
c 0 0 b
where =
1
2
1 0 0 1 . This leads to
a 0, b 0, and a + b c c.
(120)
From (118) we see that the o-diagonal matrix elements are eigenvectors of
and that mixes the diagonal entries of a density matrix. This makes the
exponentiation of rather straightforward:
et
11 12
21 22
1
a+b
etc 12
a (a11 b22 )et(a+b)
(121)
The conditions (120) on the parameters of the generator are equivalent with
etc
et(a+b) .
(122)
Inspecting (121) we see that every initial density matrix evolves toward the
equilibrium state when t
=
1
a+b
b 0
.
0 a
(123)
(124)
The vector state with zero photons is called the vacuum, we represent it by
or |0 . The dening feature of is
a = 0.
(125)
(126)
(127)
|n is the n photon state and {|n } is called the particle number basis. It
is easy to express a and a in this basis
(128)
a |n = n + 1 |n + 1 and a|n = n |n 1 .
The number operator N := a a counts the number of photons in a state
N|n = n|n .
(129)
eA+B = e 2 [A,B] eA eB .
(130)
t F (t) := e
t2
[A,B] tA tB
2
e e
(131)
(132)
eza
(133)
Remark that the left hand side is the exponential of a skew Hermitian transformation and that it is therefore unitary.
It turns out that the radiation of a laser is well-described by coherent states
za
eza
(134)
eza
n=0
zn
|n .
n!
(135)
|z|2n
.
n!
(136)
|z|2 xes the average number of photons, i.e. the intensity of the beam.
A phenomenological description of radiation loss to the vacuum can be obtained through a weak-coupling type limit. It leads to a Lindblad generator
() = aa 21 {a a, }.
(137)
Computing the action on the matrix units in the number basis yields
(|n m|) = a |n m| a 21 a a |n m| 21 |n m| a a
= mn |n 1 m 1| 12 (n + m) |n m|.
(138)
(139)
k=1
ck (t) |n k m k|.
(140)
42
(141)
Hence
(142)
et (N) = et N.
(143)
(N 2 ) = a N 2 a 21 {a a, N 2 } = 2N 2 + N.
(144)
et (N 2 ) = e2t N 2 + et (1 et )N.
(145)
= et N 0 .
(146)
et/2
N 0.
(147)
et
(148)
(149)
3.5
Thermalising maps
We present here very schematically a simple black box dynamics that describes how an environment in thermal equilibrium that is weakly coupled
43
(151)
(152)
The energy gaps {k } between the levels of H S are called the Bohr frequencies of the evolution. Generically, each k is non-degenerate except for
0 = kk which has a degeneracy d. An arbitrary observable A can be expanded in the matrix units {|k |} and this yields the following expression
for the dynamics
d
itH S
itH S
Ae
=
k,=1
44
Ak eitk |k |.
(153)
It is clear from this formula that, for a generic Hamiltonian, the constants
of the motion are precisely the linear combinations of the spectral projectors
{|k k|} of H S . In other words, the constants of the motion are just the
algebra of diagonal matrices. In the sequel we use the notation dia() with
d to denote the diagonal matrix whose entries on the diagonal are the
components of . It is also clear from (153) that the evolution of any system
observable is essentially periodic in time. It is not truly periodic because
the Bohr frequencies are in general not integer multiples of a fundamental
frequency, nevertheless after a suciently long time the system repeats itself
up to an arbitrary small error. Such a behaviour is called quasi or almost
periodic. It is in particular impossible for systems with this type of energy
spectrum to converge in the long run to some equilibrium situation. Convergence or return to equilibrium can only happen in innite systems and such
a behaviour is needed for the weak-coupling limit.
3.5.2
eH
=
Tr eH N
(154)
= N , X N , X AN .
45
XY
Tr eH XY
Tr eH Y eH X eH
=
=
Z
Z
= Y, i (X) .
Here
(155)
(156)
, XY = , Y eH X eH ) , X, Y A.
(157)
Because of the innite dimensions, we can now have a very dierent behaviour
of the dynamics, such as return to equilibrium
B
Z = , XZ , Y .
(158)
Such a behaviour is excluded for nite systems as in that situation all timedependent expectations are quasi-periodic.
46
3.5.3
The weak-coupling limit describes the simplied situation that arises when
we couple a d-level system weakly with a thermal bath at inverse temperature
. Because of the weak interaction we need to wait a long time before the
coupling aects the expectation values of system observables. It turns out,
however, that after a proper rescaling of coupling strength and time, we end
up with a simple dissipative dynamics of the small system: all memory eects
disappear and we are left with a semi-group of completely positive maps of
the system that drives any initial state toward the equilibrium state of the
system at the inverse temperature imposed by the thermal bath.
So, we consider a total Hamiltonian of system + bath
H tot = H S + H B + H int = H free + H int.
(159)
The interaction part H int is assumed to be generic. The proper scaling for
the limit is to let 0 and t such that 2 t . This causes, however,
very fast oscillations that have to be corrected for using the free evolution
( = 0). Modulo technical assumptions on the bath one can show for 0,
for any A Md , and for any d-dimensional density matrix the existence
of the following limit
Tr (X)
=
3.5.4
lim
0, t
2 t
Tr | | eitH
free
eitH
tot
X eitH
tot
eitH
free
(160)
(161)
Physically, it means that all memory in the dynamics at nite coupling disappear in the limit. As a consequence there exists a time-independent generator
47
), t 0.
t = exp(t
(162)
is only for very local use and intended to make the letter
(The notation
available.)
The canonical equilibrium state of the system has density matrix
S
eH
=
.
Z
(163)
(164)
This state is not only invariant under each t but any initial state of the
system will eventually be driven to
t
(165)
and
Here is the dissipative part of
The super-operators and [H S , ] commute, i.e. the reversible dynamics of the system commutes with the dissipative part.
satises quantum detailed balance which is expressed as
being
Hermitian with respect to the scalar product induced by on the
observables of the system
(Y ) = Tr
(X)Y, X, Y Md .
Tr X
3.5.5
(166)
(167)
This means that sends constants of the motion into constants of the
motion. Because of linearity it implies that there is a d-dimensional matrix
L such that
(dia()) = dia(L ), d .
(168)
(169)
(170)
We can put the cs in a square table and ll it up with zeroes on the diagonal
to get a second d-dimensional matrix C. So, commutation of the dissipative
and reversible parts of the dynamics leads us to specify in terms of only
two d-dimensional matrices L and C.
is the generator of a semi-group of unity-preserving completely positive
maps if and only if () = 0 and the Choi encoding C( ) is positive on the
orthogonal complement of the maximally entangled state 1d k |kk . These
conditions are equivalent with
1. L is a generator of a semi-group of stochastic matrices preserving the
constant function 1 and
11 c12 c1d
c21 22 c2d
2. the matrix ..
.. is positive semi-denite, where L =
.. . .
.
. .
.
cd1 cd2 dd
[k ]k .
It remains to nd out the consequences of quantum detailed balance. A
straightforward application of (166) yields that L has to be the generator
of a classical detailed balance process with invariant measure given by the
diagonal elements of and moreover the matrix C has to be real symmetric.
Problem 40. Fill out the details in Section 3.5.5.
49
4
4.1
Entropy
Classical entropy
xj0 xj1 .
xj0 =
(173)
j1 =1
xj0 xj1 .
xj1 =
(174)
j0 =1
(175)
This means that the letters sent out by the source are completely uncorrelated. Correlations of time length 1 yield Markov measures:
xj0 xj1 xjk =
(176)
(xj xk )
.
(xj )
(177)
(xj xk ) =
k=1
(xk xj )
(178)
k=1
(xj )Nj .
j=1
51
(179)
Now there are N!/(N1 !N2 ! Nd !) such words. Therefore the probability
distribution over the possible words is multinomial. For large N such a
distribution is sharply peaked around its maximum that is attained at Nj =
(xj )N. This follows from Stirlings formula. It also follows that almost
the full weight of the probability distribution is concentrated on a subset of
words of size exp(N(h + )) where
d
h :=
(180)
j=1
4.2
(181)
((x)).
(182)
xX
The notation H(X) is traditional and not quite optimal, H() would t better.
H takes values in [0, log|X|]. The extreme value 0 is attained for degenerate
measures (giving weight 1 to one of the points in X) while log|X| is attained
for the uniform measure (x) = 1/|X|. Furthermore H is strictly concave:
the entropy of a non-trivial mixture of two measures is strictly larger than
the corresponding mixture of entropies.
Problem 42. Verify the statements just made.
52
(xy).
(183)
yY
(184)
(185)
(186)
(187)
H(X : Y ) = 0 i the joint measure is the product of its marginals, this means
that there is independence between the two subsystems and so certainly
no strong correlations that would allow to identify with high probability
elements of X with elements of Y . The maximal value that H(X : Y ) can
attain is H(X). This happens when there is a bijective map f connecting the
elements with positive probability in X with those of positive probability in
Y such that
(xf (y)) > 0 and (xy) = 0 for y = f (x).
(188)
(189)
(190)
(191)
1
H(n X).
n
(192)
(193)
h = lim Hn+1 Hn .
(194)
and
n
(xk xi x ).
(195)
i=1
i1 ,i2 ,...,in
(197)
4.3
Quantum Entropy
We now deal with a memoryless source emitting pure states {i } with probabilities {pi }. This denes a quantum ensemble {(pi , i )} of pure states. We
can in the same way consider a quantum ensemble {(pi , i )} of mixed states,
this can be thought to be a source emitting pure composite states states of
which only one party is accessible. Observing the source by repeated measurements, we see an average state
=
i
pi |i i | or =
pi i .
(198)
(199)
Note that, in contrast to classical systems, we cannot interpret as an ensemble. Dierently built sources may generate the same because a quantum
state space is very far from a simplex.
As before, we have an entropy S that measures the uncertainty in the state
spit out by the source, this is the von Neumann entropy
S() := Tr () = Tr log .
(200)
It is not hard to see that S() = 0 i all coincide (or all i are pure and
coincide).
For more complicated sources (not memoryless) one has a description as in
the classical case. The multi-time correlated expectations are given by re-
55
Tr
( An ) = Tr
(n)
An , shift-invariance.
(201)
(202)
(203)
(204)
The term S(1 ) in the right hand side of (204) is needed because monotonicity does not hold in the quantum case, e.g., 12 can be pure in
which case 1 and 2 have the same entropy, generally strictly positive. In
this case (204) is saturated. A more symmetric version of (204) reads
|S(1 ) S(2 )| S(12 ).
(205)
(206)
(207)
(208)
(209)
Relative entropy is an important quantity. In statistics it is related to hypothesis testing. In statistical mechanics it appears in the context of the
variational principle for the free energy. Suppose that is a canonical
equilibrium state: = exp(H) where the appropriate normalisation constant, the log of the partition function, has been absorbed in the Hamiltonian,
then
S(|) = H S()
(210)
is the expected internal energy of the system in the state minus the entropy
of . Thermodynamic equilibrium is attained when this quantity reaches its
minimal value.
Problem 49. Fill out the details of the proof.
Subadditivity now follows from
0 S(12 | 1 2 ) = S(1 ) + S(2 ) S(12 ).
(211)
A really profound result, the proof of which goes beyond the scope of these
lectures, is quantum strong subadditivity
S(123 ) + S(2 ) S(12 ) + S(23 ).
57
(212)
rj |ej ej |.
(213)
rj ej fj .
(214)
:=
j
Suppose now that we are given a two party state 12 . We can always purify
this to a pure state 123 . Because marginals of a bipartite pure state have the
same non-zero eigenvalues, and hence the same entropy, we have S(23 ) =
S(1 ). Strong subadditivity then yields
S(2 ) S(12 ) + S(1 )
(215)
s = lim
(216)
Using strong subadditivity one can show that for shift-invariant states the
entropy is monotonically increasing in the volume and that
0 S((n+2) ) S((n+1) ) S((n+1) ) S((n) ).
(217)
(218)
We conclude this short series of lectures with Schumachers noiseless compression theorem for sources that emit pure quantum states. After encoding and decoding a sequence i1 i2 in of states emitted by the
58
source we should end up with a state that is close to the original state for
a subset of states that appears with high probability. In order to compare
states, fidelity can be used. For two vector states and delity is the
ordinary transition probability:
Fid(, ) := | , |2 .
(219)
A delity close to 1 means that the vectors are also close. This notion can be
extended to arbitrary mixed states. The idea is to jointly purify both density
matrices and then to use the transition probability for these purication.
More explicitly: let 1 and 2 be joint purications of 1 and 2 :
j = Tr
|j j |, j = 1, 2
(220)
then
Fid(1 , 2 ) := max | 1 , 2 |2
(221)
where the maximum is taken over all joint purications of 1 and 2 . Uhlmann
obtained an explicit expression for the mixed state delity
Fid(1 , 2 ) = Tr
1 2 1
(222)
(223)
60