You are on page 1of 21

Version 0.

2g Eduardo Martı́n-Martı́nez - AMATH 473

Block 3

Density operators, multipartite systems and


Quantum Entanglement

3.1 Density matrix, pure and non-pure states

3.1.1 Introduction: sources of uncertainty in quantum mechanics

Uncertainty in the outcome of measurements in quantum mechanics emerges from the superposition principle:
A state can be in a state which is a superposition of different eigenstates of a given observable. For example,
consider the following quantum superposition of eigenstates of σ̂z (for example, think of it as the z component
of the spin of an electron):
1  
|ψi = √ |0i + |1i . (3.1.1)
2

According to Born’s rule, the probability of finding the outcome 1 (say, spin up) is

2
2 1   1
P1 = |h1|ψi| = √ h1|0i + h1|1i = (3.1.2)
2 | {z } | {z } 2
0 1

However, it is very often the case that, on top of the ‘quantum uncertainty’, we may have certain degree
of classical uncertainty, and due to this classical uncertainty we do not know exactly the state of a system.
This ‘ignorance’ about the system can come from an undisclosed random state preparation (e.g. you put two
balls in a bag, one white and one black, and someone picks one and does not disclose to you the information
of which colour it is; you do not know with certainty the colour of the ball that is left). This ‘ignorance’ can
also come from the removal of information about the system (e.g. you have a deck, and someone shuffles it
and removes a number of cards from it, you do not know which cards you have, so to all effects you need to
consider that you have a probability distribution over all the possible cards that may have been taken out of
the deck). Atypical source of classical uncertainty is for instance thermal fluctuations.

In the next section we will introduce the density matrix, which is a linear operator in the Hilbert space
that describes a quantum system in a mixed state, i.e., a statistical ensemble of several quantum states. This
should be compared with a single state vector that describes a quantum system in a pure state, whose quantum
probability distribution we know precisely. The density matrix is the quantum-mechanical analogue to a
phase-space probability measure (probability distribution of position and momentum) in classical statistical
mechanics, and it is the formal way to introduce classical uncertainty in the formalism of quantum mechanics.

Quantum Dynamics 1 Eduardo Martı́n-Martı́nez


Version 0.2g Eduardo Martı́n-Martı́nez - AMATH 473

3.1.2 The Density operator

Postulate (alternate to postulate 1 of QM): To each physical state there corresponds a unique state
operator (which we call the density operator or, more commonly, the density matrix) which has to satisfy the
following properties:

1. It is positive definite : hψ|ρ̂|ψi ≥ 0, ∀|ψi ∈ H. Equivalently, it only has non-negative eigenvalues.


P
2. Its trace is normalized to unity: Tr ρ̂ = i hei |ρ̂|ei i = 1, {|ei i} being an orthonormal basis of H.

3. It is self-adjoint: ρ̂ = ρ̂† .

The second property comes from the fact that we would like such an operator to represent a classical
probability distribution over a set of possible quantum states |φi i. Then,
X
ρ̂ = pi |φi ihφi | , (3.1.3)
i

where the coefficients pi are the probabilities to find the system in the state |φi i, and hence must add up to
1. Then it is straightforward to see that X
Tr ρ̂ = pi = 1. (3.1.4)
i

Actually it can be shown [?, ?] that an operator ρ̂ is density matrix (i.e. satisfies the the three properties
above) ⇔ it can be written in the form (3.1.3).
Consider now that we have an observable Ô. The expectation value of such an observable on the state
represented by ρ̂ is defined as hÔi = Tr(ρ̂Ô). This definition comes from the fact that we expect this value
to be the statistical average of all the expectation values over all the quantum states that form the ensemble.
We can quickly check that this is the case:
X X X
hÔi = Tr(ρ̂Ô) = hej | ρ̂ Ô|ej i = pi hej |φi ihφi | Ô|ej i = pi hφi | Ô|φi i, (3.1.5)
j i,j i
P
where in the last equality we have used that {|ej i} is an orthonormal basis of H and as so j |ej ihej | = 11.

The probability of finding the value oi when measuring the observable Ô will be the average of the proba-
bilities of finding the outcome oi over all the states that form the ensemble, which is given by

Poi = Tr(ρ̂ |oi ihoi |), (3.1.6)

where |oi i is the eigenstate of Ô with eigenvalue oi .

3.1.3 Pure states and non-pure (mixed) states

If the state of a quantum system can be represented by a state vector |ψi, i.e., there is no classical uncertainty
in the system, the system’s density matrix will be given by

ρ̂ = |ψihψ| . (3.1.7)

Quantum Dynamics 2 Eduardo Martı́n-Martı́nez


Version 0.2g Eduardo Martı́n-Martı́nez - AMATH 473

In this case the density operator is a projector over the state |ψi. As such we know that

ρ̂2 = |ψihψ|ψi hψ| = |ψihψ| = ρ̂ (3.1.8)

and
Tr ρ̂2 = Tr ρ̂ = 1. (3.1.9)

The opposite is also true: if Tr ρ̂2 = 1, then ρ̂ is a projector onto a one-dimensional space
P and hence represents
1
a pure state. Indeed, we can write ρ̂ in terms of its spectral decomposition ρ̂ = i ρi P̂i , where P̂i is the
2 =
P 2
projector onto the eigenspace with eigenvalue ρi with multiplicity n i . Then ρ̂ i ρi P̂i . The conditions
2 2
P
Tr ρ̂ = Tr ρ̂ = 1 imply that i (ρi − ρi )ni = 0. Since the eigenvalues must be in the interval [0, 1] and ni ≥ 1,
the only possibility is that one of the ρi ’s with be equal to 1 and the rest equal to zero. The normalization of
the density matrix requires that the corresponding ni be equal to 1.

We can quantify how pure a state is by using the trace of the square of its density operator. We define
the purity of a state ρ as
P(ρ̂) = Tr ρ̂2 . (3.1.10)

The purity of a quantum state satisfies


1
≤ P(ρ̂) ≤ 1 (3.1.11)
d
where d is the dimension of the Hilbert space. P(ρ̂) = 1 for pure states. The states with maximal classical
uncertainty (maximally mixed states) have the minimum possible purity P(ρ̂) = 1/d.

The concept of purity is strongly related to quantum coherence. In this sense, we can call a state ‘more
quantum’ when its purity is high, as the system suffers decoherence, it loses purity, and therefore it would be
more difficult to observe any effect related with the quantum nature of the state. In the limit of minimum
purity P(ρ̂) = 1/d, we say that the state is ‘maximally mixed’. At this point the state is described by a density
matrix which is a multiple of the identity, which has the same form in all bases, namely

1
ρ= 11. (3.1.12)
d
Such a state does not display any quantum behaviour and all its uncertainty is of purely classical origin. In
fact, since the state is diagonal in all bases, the state does not contain any information at all.

The concept of entropy is strongly connected with the amount of classical uncertainty a state contains (it
gives a measure of the number of micro-states compatible with a given macro-state). We will not go much into
this topic here, but analogous to the Shannon entropy in information theory, or the thermodynamical Gibbs
entropy, we can define the Von Neumann entropy in quantum mechanics measuring the amount of classical
uncertainty in the system. This is defined as
X
S = − Tr(ρ̂ log ρ̂) = − ρi log ρi (3.1.13)
i

where ρi are the eigenvalues of ρ̂. The entropy of a pure state is always zero, whereas the entropy of a
maximally mixed state is the maximum possible. In particular its value is Smax = log d.

Exercise: Prove these two statements.


1
This can done because ρ̂ is a compact operator. For details see e.g. [?, ?].

Quantum Dynamics 3 Eduardo Martı́n-Martı́nez


Version 0.2g Eduardo Martı́n-Martı́nez - AMATH 473

3.1.4 Examples

In this section we will see with several example the operational difference between classical and quantum
uncertainty in the outcome of a measurement. to begin with we are going to analyze the state of a two-level
quantum system (a qubit) that we mentioned in the introduction representing the spin degree of freedom of
a spin 21 particle. Let |1i and |0i be eigenstates of σ̂z = |1ih1| − |0ih0|. We define the superposition state

1
|ψ1 i = √ (|0i + |1i) (3.1.14)
2
let us compare this state with the 50% classical ensemble of the eigenstates of σ̂z (also called an incoherent
mixture of eigenstates of σ̂z ):
1
ρ2 = (|0ih0| + |1ih1|) (3.1.15)
2
We ask the following questions:

1. What is the probability of finding the value 1 when measuring σ̂z ?

2. What is the expectation of the value of σ̂z ?

Question 1:
For the first case (pure state, superposition of |0i and |1i) it is very easy:

|ψ1 i 1
Pσz =1 = |h1|ψ1 i|2 = . (3.1.16)
2

For the second case (mixed state, incoherent mixture of |0i and |1i) it is also easy:
 
1 1
Pσρz2=1 = Tr(ρ̂2 |1ih1|) = h1|ρ̂2 |1i = h1|0ih0|1i + h1|1ih1|1i = . (3.1.17)
2 | {z } | {z } 2
=0 =1

Question 2:
For the first case it is, again, very easy:
 
1
hσz i|ψ1 i = hψ1 | σz |ψ1 i = h0|σz |0i + h0|σz |1i + h1|σz |0i + h0|σz |0i = 0. (3.1.18)
2 | {z } | {z } | {z } | {z }
=−1 =0 =0 =1

For the second case,

hσz iρ̂2 = Tr(ρ̂2 σ̂z ) = Tr [ρ̂2 (|1ih1| − |0ih0|)] = h1|ρ̂2 |1i − h0|ρ̂2 |0i = 0. (3.1.19)

So, both, the incoherent mixture and the quantum superposition give the same results. Are they then
phenomenologically indistinguishable? What is the difference? There are many differences between both
states. The states are actually radically different. One simple way of seeing it is for example computing what
would happen if we observe the σ̂x = |0ih1| + |1ih0|. Let us compute the expectation of σ̂x and the probability
of finding spin-up when measuring σ̂x in both cases.

Quantum Dynamics 4 Eduardo Martı́n-Martı́nez


Version 0.2g Eduardo Martı́n-Martı́nez - AMATH 473

3. What is the probability of finding the value 1 when measuring σ̂x ?

4. What is the expectation of the value of σ̂x ?

Let us first note that the eigenstates of σ̂x are


1 1
|+i = √ (|0i + |1i) , |−i = √ (|0i − |1i) . (3.1.20)
2 2

Question 3:
For |ψ1 i the probability of obtaining one is trivial to compute, since |ψ1 i = |+i. Therefore,
|ψ i
Pσ̂x 1=1 = |h+|ψ1 i|2 = 1. (3.1.21)

For the second case (mixed state, incoherent mixture of |0i and |1i) it is also easy:
 

h+|0ih0|+i + h+|1ih1|+i = 1 .
1
Pσρx2=1 = Tr(ρ̂2 |+ih+|) = h+|ρ̂2 |+i =

(3.1.22)
2 | {z } | {z } 2

= 12 = 12

Question 4:
For the first case it is, again, very easy just remembering that |ψ1 i = |+i is an eigenstate of σx with
eigenvalue 1:
hσx i|ψ1 i = hψ1 | σx |ψ1 i = hψ1 |ψ1 i = 1. (3.1.23)

For the second case it is also not very difficult looking at ρ̂2 in (3.1.15) and noticing that it is diagonal in
the σ̂z eigenbasis:

hσx iρ̂2 = Tr(ρ̂2 σ̂x ) = Tr [ρ̂2 (|0ih1| + |1ih0|)] = h1|ρ̂|0i + h0|ρ̂|1i = 0. (3.1.24)
| {z } | {z }
=0 =0

So, as we have seen, the answer to the questions about σ̂x are radically different in one case (coherent
superposition) and the other (incoherent mixture). In particular, the incoherent mixture in this case shows
identical statistics in all the spin observables, whereas the coherent superposition displays different behaviour
when considering observers that do not commute such as σ̂z and σ̂x . It is definitely not equivalent to have
classical or quantum uncertainty about the value of the z component of spin. Even if the uncertainty is similar
in both cases for the observable σ̂z , differences appear.

3.1.5 Thermal states

It is well known that a finite temperature temperature introduces decoherence and makes it more difficult to
observe quantum effects. We have already learned that quantum coherence is related with the degree of purity
of the state. Therefore, it seems reasonable that we should be able to describe the effect of temperature on a
quantum system using the density matrix formalism.

Quantum Dynamics 5 Eduardo Martı́n-Martı́nez


Version 0.2g Eduardo Martı́n-Martı́nez - AMATH 473

Gibbs thermality

Let us begin with the notion of thermality introduced by Gibbs. Gibbs introduced the notion of thermality as
the stationary states that maximize entropy at constant energy. We may wonder what would be the density
matrix that satisfies this constraint.
If we maximize the entropy under the constraint that the expectation of the Hamiltonian hĤi = Tr(ρ̂Ĥ)
is constant we get the following Langrange multipliers problem: We have to maximize S = − Tr[ρ̂ log(ρ̂)]
under the constraints that Tr(ρ̂Ĥ) = E and Tr(ρ̂) = 1. If we enforce these constraints by means of Lagrange
multipliers β and µ, we end with the equation

∂ h   i
S − β Tr(ρ̂Ĥ) − E − µ[Tr(ρ̂) − 1] = 0. (3.1.25)
∂ ρ̂
P
We can rewrite this equation in a more convenient form using the spectral theorem. Since ρ̂ = i ρi P̂i , where
ρi are the eigenvalues (without repetition) of ρ̂ each of them with multiplicity ni and P̂i the projectors onto
the corresponding eigenspaces (which can have rank up to ni ), the entropy and the constraints can be written
as X X X
S= ni ρi log ρi , ni Ei ρi = E, ni ρi = 1, (3.1.26)
i i i

where Ei = Tr(P̂i Ĥ). Then Eq. (3.1.25) is equivalent to


" ! !#
∂ X X X
ni ρi log ρi − β ni Ei ρi − E − µ ni ρi − 1 = 0, (3.1.27)
∂ρi
i i i

which implies
− log ρi − 1 − βEi − µ = 0, (3.1.28)
or, going back to operator language, multiplying by P̂i and summing to all i:

− log ρ̂ − 11 − β Ĥ − µ11 = 0, (3.1.29)

where in the last step we have used the spectral decomposition of the operators in the eigenbasis of ρ̂. Therefore
we can straightforwardly obtain the solution
h i
ρ̂ = exp −β Ĥ − (µ + 1)11 = e−(µ+1)11 e−β Ĥ . (3.1.30)

The Lagrange multiplier µ can be calculated using the constraint Tr(ρ̂) = 1:

(µ + 1) = log Z(β) Z(β) = Tr e−β Ĥ . (3.1.31)

The multiplier β is determined by the constraint Tr(ρ̂Ĥ) = E:

1 ∂
E= Tr(Ĥe−β Ĥ ) = log Z(β). (3.1.32)
Z(β) ∂β

Therefore, we can conclude that the state that maximizes the entropy for a given energy is of the form

e−β Ĥ
ρ̂β = , Z(β) = Tr e−β Ĥ . (3.1.33)
Z(β)

Quantum Dynamics 6 Eduardo Martı́n-Martı́nez


Version 0.2g Eduardo Martı́n-Martı́nez - AMATH 473

where β is the solution of Eq. (3.1.32). We call that state Gibbs (thermal) state.
The parameter β labels the Gibbs state, and can be quickly associated with temperature T . As we know
from classical statistical mechanics, the state of a mechanical system at thermal equilibrium with a heat
bath at a finite temperature is given by the canonical ensemble. The probability of finding the system in a
given energy state is proportional to the Boltzmann distribution e−E/kb T divided by the partition function.
Consequently, the quantum state representing a thermal system will be an incoherent mixture of eigenstates
of the Hamiltonian, each one contributing with a probability given by the Boltzmann factor. Translating this
into the density matrix formalism one gets that
1 X −Ei /(kb T )
ρ̂T = e P̂Ei , (3.1.34)
Z(T )
i

where P̂Ei is the projector onto the subspace with energy Ei and, by comparison, we can assign the value
β = 1/kb T , which relates the parameter β (that we call inverse temperature) with the temperature T . The
expression (3.1.34) corresponds to the expression (3.1.33) obtained imposing Gibb’s thermality criterion on a
density matrix, which represents the the state of a finite-dimensional quantum system at thermal equilibrium
with a reservoir at temperature T .
Notice that in the limit of infinite temperature, β → 0 and thus
1 X 1 1
ρ̂β→0 = P̂Ei = 11 = 11 (3.1.35)
Z(β → 0) Z(β → 0) d
i
P
where d = i ni is the dimension of the Hilbert space. (3.1.35) is the maximally mixed state. At high
temperature, quantum states lose all their quantum coherence in all bases.

Kubo-Martin-Schwinger thermality and the KMS condition

The definition of thermality a-la Gibbs is not satisfactory when we think of infinite-volume quantum systems.
Indeed, in those cases, the definition of the partition function can be problematic and may not even exist.
This is the case, for instance, of quantum fields. Sometimes this is addressed by considering formally that we
have a Gibbs state and regularizing the quantization so that one has for instance the system constrained in
a finite box, and later taking a limit that lifts the volume cutoff. However this approach is problematic since
such limits are not always guaranteed to converge uniformly or even exist.
A more general and practical definition of thermally was given by Kubo, Martin and Schwinger [] in terms
of the so-called KMS condition. ρ̂ is a KMS state (with inverse KMS temperature β) with respect to the
time parameter τ parameterizing translations generated by a Hamiltonian Ĥ if and only if it satisfies the
following two conditions for any pair of self-adjoint bounded Heisenberg picture operators Â(τ ) = Û Â(0)Û †
and B̂(τ ) = Û B̂(0)Û † (where Û = eiĤτ )

i. The expectation values hÂ(0)B̂(τ )iρ̂ and hB̂(τ )Â(0)iρ̂ are boundary values of some complex functions
hÂ(0)B̂(z)iρ̂ and hB̂(z)Â(0)iρ̂ holomorphic in the complex plane strips 0 < Imz < β and −β < Imz < 0,
respectively.

ii. The following complex anti-periodicity (of period β) is satisfied by the boundary values of these complex
functions:
hÂ(0)B̂(τ + iβ)iρ̂ = hB̂(τ )Â(0)iρ̂ . (3.1.36)

Quantum Dynamics 7 Eduardo Martı́n-Martı́nez


Version 0.2g Eduardo Martı́n-Martı́nez - AMATH 473

The first condition is often overlooked and the KMS condition is given as just as condition ii. We will see
in later chapters, when we study the thermalization of particle detectors with quantum fields, that condition
i is crucial in order to derive detailed balance conditions out of KMS.
The KMS condition implies stationarity: From condition ii for the particular choice  = 11 and for any
self-adjoint operator B̂, we obtain
hB̂(τ + iβ)i = hB̂(τ )i, (3.1.37)
which means that hB̂(z)i is periodic in the imaginary direction and in particular that hB̂(iβ)i = hB̂(0)i.
Schwarz’s inequality implies that for z = x + iy

|hB̂(z)i| ≤ ||B̂(z)|| = ||B̂(x + iy)|| = ||eiĤx B̂(iy)e−iĤx || = ||B̂(iy)||. (3.1.38)

Therefore, on the strip 0 < Imz < β, |hB̂(z)i| is bounded by M = sup{||B̂(iy)|| : y ∈ [0, β]}; since hB̂(iy)i
is periodic and holomorphic in that strip then M < ∞. Finally, because of the complex periodicity of
hB̂(τ + iβ)i for any value of τ we know that hB̂(z)i is bounded everywhere in the complex plane. Finally
from Liouville’s theorem we know that since hB̂(z)i is bounded and holomorphic, it has to be constant and
therefore hB̂(z)i = hB̂(0)i, and in particular hB̂(τ )i = hB̂(0)i and thus, since B̂ is an arbitrary observable, we
have shown that all observables have constant in time expectation values (which includes all the moments of
all the observables of the system). Therefore all KMS states are stationary.
The characterization of thermal states in terms of the KMS condition is motivated by the following theorem:
Whenever a canonical ensemble can be defined, a state ρ̂ is a Gibbs state if and only if it is a KMS state.
In view of this theorem, KMS thermality generalizes Gibbs thermality to the cases where the partition
function cannot be properly defined (for example in quantum field theory2 ). The KMS condition also highlights
an interesting relationship between thermality and the behaviour of correlations: thermality can be defined
through the properties of the complex extensions of the time correlations between all bounded operators.
Let us show that indeed all Gibbs states are KMS states, and conversely, that KMS states are Gibbs states
when the partition function can be defined.
First, let us prove that Gibbs ⇒ KMS: Let ρ̂ be a Gibbs state of inverse temperature β.

• Condition i: The operator

Â(0)B̂(τ + iσ)ρ̂ = Â(0)e−σĤ B̂(τ )e−(β−σ)Ĥ /Z(β) (3.1.39)

is of trace class3 for 0 ≤ σ ≤ β because all the operators on the right-hand side are bounded and e−σĤ
is a trace-class operator. The product of a bounded operator and a trace-class operator is trace class4 .
This means that in the complex strip 0 < Im z < β,

hÂ(0)B̂(z)iρ̂ = Tr[Â(0)B̂(z)ρ̂] (3.1.40)

is a well-defined function of z. Besides it satisfies the Cauchy-Riemann equations as can be checked


straightforwardly. Indeed, (3.1.40) defines the complex function f (z) = Tr(Â(0)eiz Ĥ B̂(0)e−iz Ĥ ρ̂). Writ-
ing z = τ + iσ, the derivative with respect to τ and the derivative with respect to σ are equal except
2
Notice that we have demanded that A and B be bounded operators. We have also claimed that the KMS condition is
convenient to define thermality in quantum field theory. A reader may be questioning the compatibility of these two statements
since in quantum field theory field operators are not generally bounded. Indeed the KMS condition in field theory is a bit trickier
than it is addressed in this section and we will revisit it when we review quantum field theory.
1
3
An operator Ô is of trace class if Tr[(Ô† Ô) 2 ] is finite.
4
See Reed and Simon, section VI: Bounded operators. Exercises 27 and 28 imply this statement.

Quantum Dynamics 8 Eduardo Martı́n-Martı́nez


Version 0.2g Eduardo Martı́n-Martı́nez - AMATH 473

for a factor i: ∂τ f = −i∂σ f . An entirely analogous reasoning can be applied to hB̂(z)Â(0)i. Therefore
a Gibbs state satisfies condition i.

• Condition ii: To prove this condition we take the boundary value at Im z = β, z = τ + iβ and use the
cyclic property of the trace in (3.1.40):
On the one hand we know that due to stationarity, the Hamiltonian cannot be time dependent. then

B̂(τ ) = eiĤτ B̂(0)e−iĤτ . (3.1.41)

Extending τ to the complex plane we get that

B̂(τ + iβ) = eiĤ(τ +iβ) B̂(0)e−iĤ(τ +iβ) = e−β Ĥ B̂(τ )eβ Ĥ . (3.1.42)

substituting this and the Gibbs state ρ̂ = [Z(β)]−1 e−βH in (3.1.36),

hÂ(0)B̂(τ + iβ)iρ̂ = Tr[Â(0)e−β Ĥ B̂(τ )eβ Ĥ e−β Ĥ ]/Z(β)


= Tr[B̂(τ )Â(0)e−β Ĥ ]/Z(β) = Tr[B̂(τ )Â(0)ρ̂]
= hB̂(τ )Â(0)iρ̂ . (3.1.43)

This proves that Gibbs ⇒ KMS.


Second, let us prove that KMS ⇒ Gibbs when Gibbs is well defined: Let us assume that ρ̂ is a KMS state.
If we use condition ii. at τ = 0, we see that Tr(Â(0)e−β Ĥ [B̂(0), eβ Ĥ ρ̂]) = 0 for any bounded operator Â.
This implies that [B̂(0), eβ Ĥ ρ̂] = 0 for any bounded operator  and hence eβ Ĥ ρ̂ must be proportional to the
identity, i.e. ρ̂ is a Gibbs state.
KMS states satisfy the following set of necessary conditions on the two-point correlator of an arbitrary
Heisenberg picture observable Â(τ ) defined as
D E
Wρ̂ (τ, τ 0 ) := Â(τ )Â(τ 0 ) , (3.1.44)
ρ̂

which will be useful when dealing with quantum fields:

• Stationarity: Wρ̂ (τ, τ 0 ) = Wρ̂ (0, ∆τ ), with ∆τ = τ 0 − τ .

• Holomorphicity of the two-point correlator Wρ̂ (0, z) in the upper complex strip 0 < Im z < β.

• Complex anti-periodicity (of period β) of the two point correlator:

Wρ̂ (0, ∆τ + iβ) = Wρ̂ (0, −∆τ ). (3.1.45)

3.2 Multipartite systems

When it comes to studying multipartite systems, a very old question comes to mind. Is the whole greater
than the sum of its parts? Or in other words, is it possible to fully describe a system by entirely describing
their parts? We will discuss that, in quantum mechanics, this is not the case.
We would like to describe the state of a multipartite quantum system. For simplicity we will focus on a
bipartite system consisting of two parts A and B. The Hilbert spaces where the individual systems’ states

Quantum Dynamics 9 Eduardo Martı́n-Martı́nez


Version 0.2g Eduardo Martı́n-Martı́nez - AMATH 473

live are respectively HA and HB . The joint quantum system state will live in the tensor product of the two
individual Hilbert spaces H = HA ⊗ HB . If {|ai i} and {|bj i} are respectively bases of HA and HB , a basis of
the joint system Hilbert space H will be given by

{|ai , bj i} = {|ai i ⊗ |bj i}. (3.2.1)

In this product Hilbert space we can define a density operator for the bipartite system ρ̂AB acting on
HA ⊗ HA . The expectation value of observables of the joint system is computed in exactly the same way as
we discussed above
X
hÔiρAB = Tr(ρ̂AB Ô) = hai , bj |ρ̂AB Ô|ai , bj i
i,j
XX
= hai , bj |ρ̂AB |an , bm ihan , bm |Ô|ai , bj i. (3.2.2)
i,j m,n

We can define the reduced state of each subsystem as follows:


X X
ρ̂A = TrB (ρ̂AB ) = hbi |ρ̂AB |bi i, ρ̂B = TrA (ρ̂AB ) = hai |ρ̂AB |ai i, (3.2.3)
i i

where TrB denotes the partial trace of ρAB with respect to the system B. The partial trace effectively
erases the information about the subsystem B and the reduced states describe the individual states of each
subsystem. Indeed it is easy to check that if we compute the expectation of an observable which belongs only
to the subsystem A, for example Ô = ÔA ⊗ 11B , the outcome cannot depend on the state of B and therefore
hÔiρAB = hÔA iρA , as we can easily see to be case (exercise). Therefore the expectation of local observables
(belonging to the subsystem A) only depends on the individual state of the system A, ρ̂A . The question that
we formulated in the beginning of this section can be reformulated as, can the joint state be fully described
just knowing the partial state of the subsystems A and B?

3.2.1 A qubit example

Let us consider a bipartite system formed by two qubits, A and B. The bases of the respective 2-dimensional
Hilbert spaces are n o n o
HA = span |0A i, |1A i , HB = span |0B i, |1B i . (3.2.4)

The state of the two qubits would live in the tensor product Hilbert space
n o
HAB = HA ⊗ HB = span |0A 0B i, |0A 1B i, |1A 0B i, |1A 1B i . (3.2.5)

Consider an arbitrary pure state for the two qubits

|ψA i = α|0A i + β|1A i, |ψB i = γ|0B i + δ|1B i, (3.2.6)

with |α|2 + |β|2 = |γ|2 + |δ|2 = 1. If the two qubits are in arbitrary pure states, the bipartite state fully
describing the two qubits will be the tensor product of the two states

|ψAB i = |ψA i ⊗ |ψB i = αγ|0A 0B i + αδ|0A 1B i + βγ|1A 0B i + βδ|1A 1B i. (3.2.7)

Quantum Dynamics 10 Eduardo Martı́n-Martı́nez


Version 0.2g Eduardo Martı́n-Martı́nez - AMATH 473

We can, for example compute the partial trace of the bipartite system with respect to B to obtain the state
of the subsystem A:
X
ρ̂A = TrB |ψAB ihψAB | = hnB |ψAB ihψAB |nB i (3.2.8)
n=0,1

= h0B |ψAB ihψAB |0B i + h1B |ψAB ihψAB |1B i. (3.2.9)

A quick calculation shows that

h0B |ψAB i = αγ|0A i + βγ|1A i, h1B |ψAB i = αδ|0A i + βδ|1A i, (3.2.10)

which substituted in (3.2.9) gives

ρ̂A = |α|2 |0A ih0B | + |β|2 |1A ih1A | + αβ ∗ |0A ih1A | + α∗ β |1A ih0A | = |ψA ihψA | , (3.2.11)

where we have made use of the normalization |γ|2 + |δ|2 = 1. Therefore, as we saw, the partial state ρ̂A
obtained as the partial trace with respect to B of ρ̂AB is just the pure state |ψA i.

We can also compute, for example, the expectation value of the operator σ̂zA ⊗ 11B on the state |ψAB i:

hσ̂zA ⊗ 11B i = hψAB |σ̂zA ⊗ 11B |ψAB i = hψA |σ̂zA |ψA ihψB |11B |ψB i
= hψA |σ̂zA |ψA i = |β|2 − |α|2 . (3.2.12)

where, in the last equality we have used that σz A |0A i = −|0A i, σz A |1A i = |1A i.

3.2.2 A matrix representation, the Kronecker product

Let us quickly see a way to represent tensor product given two matrix representation of states

If we have two m × n and p × q matrices  and B̂:


   
a11 a12 ... a1n b11 b12 . . . b1q
 a21 a22 ... a2n   b21 b22 . . . b2q 
 =  , B̂ =  (3.2.13)
   
.. .. .. .. .. .. .. .. 
 . . . .   . . . . 
am1 am2 . . . amn bp1 bp2 . . . bpq

the Kronecker product of the two matrices  ⊗ B̂ will be the mq × np matrix

 
a11 B̂ a12 B̂ ... a1n B̂
 a21 B̂ a22 B̂ ... a2n B̂ 
 ⊗ B̂ =  (3.2.14)
 
.. .. .. .. 
 . . . . 
am1 B̂ am2 B̂ . . . amn B̂

Quantum Dynamics 11 Eduardo Martı́n-Martı́nez


Version 0.2g Eduardo Martı́n-Martı́nez - AMATH 473

or explicitly
 
a11 b11 a11 b12 ··· a11 b1q ··· ··· a1n b11 a1n b12 ··· a1n b1q
 a11 b21 a11 b22
 ··· a11 b2q ··· ··· a1n b21 a1n b22 ··· a1n b2q 
 .. .. .. .. .. .. .. .. 
 .
 . . . . . . .  
 a11 bp1 a11 bp2
 ··· a11 bpq · · · ··· a1n bp1 a1n bp2 ··· a1n bpq 
 .. .. .. .. .. .. .. 
 . . . . . . . 
 ⊗ B̂ =  .

.. .. .. .. .. ..  .
 (3.2.15)
 .. . . . . . . 
 
am1 b11 am1 b12
 ··· am1 b1q · · · ··· amn b11 amn b12 ··· amn b1q 
am1 b21 am1 b22
 ··· am1 b2q · · · ··· amn b21 amn b22 ··· amn b2q 
 .. .. .. .. .. .. .. .. 
 . . . . . . . . 
am1 bp1 am1 bp2 ··· am1 bpq · · · ··· amn bp1 amn bp2 ··· amn bpq

In the particular case of the two-qubit example studied above, if we take the representation that we have
being using up to now    
0 1
|0i = , |1i = , (3.2.16)
1 0
then    
β δ
|ψA i = , |ψB i = , (3.2.17)
α γ
and therefore  
βδ
βγ 
|ψAB i = |ψA i ⊗ |ψB i = 
 αδ  .
 (3.2.18)
αγ
Also, for instance, the operators σ̂zA ⊗ 11B and σ̂zA ⊗ σ̂xB would be represented by
 
    1 0 0 0
1 0 1 0  0 1 0 0
σ̂zA ⊗ 11B = ⊗ = , (3.2.19)
0 −1 0 1 0 0 −1 0 
0 0 0 −1
 
    0 1 0 0
1 0 0 1 1 0 0 0
σ̂zA ⊗ σ̂xB = ⊗ = . (3.2.20)
0 −1 1 0 0 0 0 −1
0 0 −1 0

The expectation value computed above, can then be evaluated as

hσ̂zA ⊗ 11B i = hψAB |σz A ⊗ 11B |ψAB i


  
1 0 0 0 βδ
0 1 0 0 βγ 
= (βδ)∗ , (βγ)∗ , (βδ)∗ , (βγ)∗     = |β|2 − |α|2 .
 
(3.2.21)
0 0 −1 0   αδ 
0 0 0 −1 αγ

same as obtained before.

Quantum Dynamics 12 Eduardo Martı́n-Martı́nez


Version 0.2g Eduardo Martı́n-Martı́nez - AMATH 473

3.3 The Einstein-Podolsky-Rosen Argument, Bell Inequalities and non-


locality of QM

Einstein, Podolski and Rosen (EPR) wrote a paper in 1935 entitled “Can Quantum-Mechanical Description of
Physical Reality Be Considered Complete?” [?] challenging the usual probabilistic Copenhagen interpretation
of quantum mechanics. To understand the idea behind that paper let us consider the following four maximally
entangled bipartite states:
1 1
|Φ± i = √ (|0A 0B i ± |1A 1B i) , |Ψ± i = √ (|0A 1B i ± |1A 0B i) . (3.3.1)
2 2
They are called Bell States and form a complete orthonormal basis of the Hilbert space of two qubits.
Imagine that we take one of those states, for example |Φ+ i. In that state the value of Alice’s qubit (A)
and Bob’s quibit (B) is not determined, but, as in Schrödinger’s cat, they are in a superposition of 1 and
0. However, what happens if, say, Alice measures her qubit and obtains a 0? From a strictly Copenhaguian
interpretation, the state would collapse to the projection onto the |0A i space, but that projection also kills
the second summand in the state:
P̂|0A i |Φ± i
= |0A 0B i, (3.3.2)
P̂|0A i |Φ± i

and then, the result of any measurement that Bob carries will be predetermined to be always 0. If we considered
this as a dynamical process, this would have implied that there is ‘instantaneous action at distance’ that
propagated instantaneously the outcome of the measure from Bob. That is independent from the distance
that physically separates Alice and Bob. This is utterly in conflict with the theory of relativity, because even
who measures first is an observer dependent statement.
According to EPR there were two possible explanations. Either there was some interaction between the
particles, even though they were spacelike separated, or the information about the outcome of all possible
measurements was already present in both particles and it is no longer probabilistic. The EPR authors
preferred the second explanation according to which that information was encoded in some ‘hidden variables’.
The first explanation, that an effect propagated instantly, across a distance, is in conflict with the theory
of relativity as stated above. They then concluded that quantum mechanics was incomplete since, in its
formalism, there was no room for such hidden parametersleading to the so-called ‘hidden variable theories’.
This provided hope that a more-complete (and less-troubling) theory might one day be discovered.
The discussion remained theoretical (even philosophical) until the formalization of Bell’s theorem [?]. It
gave raise to an experiment that was able to falsify the hypothesis of local realism (the physical magnitudes
have a definite value and are elements of reality, while probability in quantum mechanics only comes from not
knowing their defined values). This experiment relied in the measurement of correlations between outcomes
of measurements in the two components of a bipartite system. To even state Bell’s theorem via a derivation
of the Bell-CHSH (Clauser, Horne, Shimony, Holt) inequality [?, ?], we first have to formalize the concept of
‘local realism’. Here we follow [?] for the presentation of this inequality.
Consider a two-qubit system and a pair of instruments that can measure on each of the components of
the system. The possible results of a measurement on either Alice’s or Bob’s side are obviously ±1. Each
instrument has a range of settings (corresponding, for example, to the possible orientations of a spin-measuring
apparatus such as a Stern–Gerlach device). These settings will be denoted as a for the first instrument and b for
the second instrument. Let us assume that the result of a measurement will be deterministic, and may depend
on the controllable parameters a and b, and on any number of uncontrolled parameters (hidden variables)

Quantum Dynamics 13 Eduardo Martı́n-Martı́nez


Version 0.2g Eduardo Martı́n-Martı́nez - AMATH 473

denoted collectively as λ. The result A = ±1 of the measurement on the first subsystem may depend on the
setting a of the first instrument and on the shared hidden variables λ. Therefore we assume that there is a
function A(a, λ) = ±1 which determines the result of the measurement on the first particle. From the point
of view of someone who does not have access to the hidden variables λ the outcome will look probabilistic.
Similarly we assume that there is a function B(b, λ) = ±1 which determines the result of the measurement
of the second qubit. But, in accordance with Einstein’s principle of locality, we assume that the result
of a measurement on the first qubit does not depend on the setting b of the second instrument, and that
the result of a measurement on the second qubit does not depend on the setting a of the first qubit. We
make no assumptions on the uncontrollable parameters λ. They may be associated with the qubits, with
the instruments, with the environment, or jointly with all of these. It makes no difference to the argument.
Consider also that we make the measurements on A and B in spacelike separated regions, so it is impossible
that any signal could make it from Alice to Bob.
Now we would like to study the correlations between the outcomes of measurements in Alice’s and Bob’s
subsystems, i.e., the product between the two outcomes of measurements on A and B, since the possible
outcomes are ±1 the correlators can be either 1 or −1. To make it even more general, the hidden parameters
λ are subject to a certain classical probability distribution indicating our ignorance of their value on every
single experiment. Therefore, for a given configuration of the controllable experimental settings a, b, the
correlator is of the form Z
C(a, b) = A(a, λ) B(b, λ) Λ(λ) dλ, (3.3.3)
R
where A(a, λ) = ±1, B(b, λ) = ±1, and Λ(λ) satisfies Λ(λ) dλ = 1. The hidden variables are integrated out
since we do not have access to their values.
Let us consider two different alternate configurations a y a0 for Alice’s experimental instrument and b y b0
for Bob’s. In this case,
Z
0
A(a, λ) B(b, λ) − A(a, λ) B(b0 , λ) Λ(λ) dλ
 
C(a, b) − C(a, b ) =
Z
= A(a, λ) B(b, λ) 1 ± A(a0 , λ) B(b0 , λ) Λ(λ) dλ
 

Z
− A(a, λ) B(b0 , λ) 1 ± A(a0 , λ) B(b, λ) Λ(λ) dλ,
 
(3.3.4)

where in the second equality we have added and subtracted the same term. Now, since |A(a, λ)| ≤ 1, |B(b, λ)| ≤
1, we get that
Z
0
1 ± A(a0 , λ) B(b0 , λ) Λ(λ) dλ
 
C(a, b) − C(a, b ) ≤
Z
1 ± A(a0 , λ) B(b, λ) Λ(λ) dλ
 
+

= 2 ± C(a0 , b0 ) + C(a0 , b) ,
 
(3.3.5)
R
where we have used that Λ(λ) dλ = 1. This implies the following weaker inequality

C(a, b) − C(a, b0 ) + C(a0 , b0 ) + C(a0 , b) ≤ 2. (3.3.6)

This result is known as Bell’s inequality. If the results of correlated outcomes in an experiment can be explained
as EPR proposed, then the correlators have to satisfy (3.3.6). It is not a sufficient condition, but it is indeed
necessary.

Quantum Dynamics 14 Eduardo Martı́n-Martı́nez


Version 0.2g Eduardo Martı́n-Martı́nez - AMATH 473

As it is expressed now, it is not obvious whether quantum mechanics would violate that inequality. To
check it, consider the system of two qubits (for example, electron spins) to be in one of the Bell states, for
example |Ψ− i, and consider that Alice can choose between two detector settings a and a0 corresponding to
measurement of spin along the z and the x axis, respectively. Bob has a different reference frame (his axes are
rotated with respect to Alice’s; think of a Stern-Gerlach again). He can choose between two detector settings
b and b0 corresponding to measurement of spin along his reference frame’s z 0 and x0 axis, respectively, where
the (x0 , z 0 ) coordinate system is rotated an angle 3π/4 relative to the (x, z) coordinate system. If we refer
all our measurements to Alice’s reference frame,√Alice is measuring σ̂z for the configuration √ a and σ̂x for the
0
configuration a , whereas Bob is measuring (1/ 2)(σ̂z + σ̂x ) for the configuration b and (1/ 2)(σ̂z − σ̂x ) for
the configuration b0 . Expressing the outcome of the measurements in the notation introduced above we get

A(a) = σ̂z ⊗ 11, A(a0 ) = σ̂x ⊗ 11, (3.3.7)


1 1
B(b) = − √ 11 ⊗ (σ̂z + σx ), B(b0 ) = √ 11 ⊗ (σ̂z − σ̂x ). (3.3.8)
2 2
Since the measurements on A all commute with measurements on B it is very easy to evaluate the expectation
of the correlators on the entangled state |Ψ− i, yielding

1
hA(a)B(b)i = hA(a0 )B(b0 )i = hA(a0 )B(b)i = −hA(a)B(b0 )i = √ . (3.3.9)
2
Substituting this result in the left-hand side of (3.3.6) we obtain

|hA(a)B(b)i − hA(a)B(b0 )i| + |hA(a0 )B(b0 )i + hA(a0 )B(b)i| = 2 2 > 2. (3.3.10)

So quantum mechanics violates Bell’s inequalities. Quantum mechanics is not compatible with the existence
of definite values for the observables that we do not know, or for that matter any other complicated set of
hidden parameters that we cannot control, either in the system, environment or experimental setting. The
EPR argument preferred option was not correct.
Does that mean that there is a spooky action that propagates faster than light? The answer is no. Quantum
entanglement cannot be used for signalling, since if Alice wants to inform of the outcome of a measurement
to Bob, she needs to send a signal through a classical channel that is inherently limited by the speed of
light. So, Alice and Bob can never use an entangled state to violate causality or signal faster than light. The
existence of quantum entanglement responds to the non-locality of quantum states and it was generated in
the state preparation. This is compatible with the theory of relativity and still neither Alice or Bob will know
beforehand the outcome of their measurements. The outcome is not predefined, but they know that no matter
who measures first and what the outcome is, their measurements will be correlated. These kind of correlations
do not exist in classical physics.
The ‘realism’ assumption is actually somewhat idealistic, and Bell’s theorem only proves non-locality with
respect to variables that only ‘exist’ for metaphysical reasons. However, before the discovery of quantum
mechanics, both realism and locality were completely uncontroversial features of physical theories.

3.4 State separability and quantum entanglement

For pure states, the question whether it is enough to describe the subsystems A and B to describe the composite
system AB can be very easily formulated as the following: for a pure bipartite state |ψAB i, is it enough to
fully know the partial states ρ̂A = TrB (|ψAB ihψAB |) and ρ̂B = TrA (|ψAB ihψAB |)?

Quantum Dynamics 15 Eduardo Martı́n-Martı́nez


Version 0.2g Eduardo Martı́n-Martı́nez - AMATH 473

It will be true if the system is in a product state, that is

|ψAB i = |ψA i ⊗ |ψB i. (3.4.1)

Hence, to know the whole state it would be sufficient to know the partial state of the subsystems A and B
(|ψA i and |ψB i) to explain the result of any experiment.
However, there are bipartite states that cannot be expressed as a tensor product of the partial states of
the subsystems A and B. Let us see an example. Consider the following two bipartite states

1
|ψs i = (|0A 0B i + |1A 1B i + |0A 1B i + |1A 0B i) , (3.4.2)
2
1
|ψe i = √ (|0A 0B i + |1A 1B i) . (3.4.3)
2
Can we express these two states as a tensor product of two independent states one for for the system A and
one for the system B (following convention, we will call them Alice and Bob)?
One can easily find that for the first state

1   
|ψs i = |0A i + |1A i ⊗ |0B i + |1B i . (3.4.4)
2 | {z } | {z }
|ψsA i |ψsB i

However, one can see that for the second, there is no αA,B and βA,B that factorizes the state

|ψe i =
6 (αA |0A i + βA |1A i) ⊗ (αB |0B i + βB |1B i) . (3.4.5)

We call this state non-separable.


Also, if we compute the partial states of Alice and Bob we obtain, for the first case,

ρA
s = TrB (|ψs ihψs |) = h0B |ψs ihψs |0B i + h1B |ψs ihψs |1B i
1
= (|0A ih0A | + |0A ih1A | + |1A ih0A | + |1A ih1A |) = ψsA ψsA , (3.4.6)
2

which is a pure state. Analogously, ρB B


s = ψs ψsB .
However, we can as easily see that the partial states of A and B in |ψe i will not be pure states. Indeed,

ρA
e = TrB (|ψe ihψe |) = h0B |ψe ihψe |0B i + h1B |ψe ihψe |1B i
1
= (|0A ih0A | + |1A ih1A |) 6= ψsA ψsA . (3.4.7)
2
Eliminating the information about B we have also eliminated information about A in the sense that A now
contains classical uncertainty. In fact, in this case, the resulting state is the maximally mixed state (it contains
no information at all) as can be seen from the fact that Tr ρ̂2A = 1/2 6= 1.
A pure state that is non-separable is an entangled state. The entangled state we played with is such
that the partial states are maximally mixed, whereas the separable state is such that the partial states are
pure states. This is suggesting that the amount of mixedness of the partial states may be related with the
amount of entanglement of the joint state (at least if the joint state is pure). In general, a bipartite state
represented by a density matrix ρ̂AB is non-separabe (thus, entangled) if and only if it cannot be expressed

Quantum Dynamics 16 Eduardo Martı́n-Martı́nez


Version 0.2g Eduardo Martı́n-Martı́nez - AMATH 473

as a probability distribution of the uncorrelated individual states i.e. if and only if it cannot
P be expressed as
a convex combination of product states. Given a set of positive numbers {pi } such that i pi = 1
X
ρ̂AB 6= pi ρ̂A ⊗ ρ̂B ⇐⇒ ρ̂AB is entangled, (3.4.8)
i
P
with pi ∈ [0, 1] and i pi = 1.
Although determining if a state is entangled or not is conceptually simple, computationally speaking is
a very hard problem for general states of arbitrary dimension. Actually there is no such thing as a unique
measure of entanglement. Instead, a measure of entanglement is any positive function of the state E(ρ) which
satisfy the following axioms

• Must be maximum for maximally entangled states (Bell states).

• Must be zero for separable states.

• Must be non-zero for all non-separable states.

• Must not grow under local operations plus classical communication (LOCC).

3.4.1 Entanglement entropy

For pure states we can use the entropy of the partial states as a measure of entanglement. We define the
Entanglement entropy as the Von Neumann Entropy of the partial state:

Sent (ρ̂AB ) = S(ρ̂A ) = S(ρ̂B ). (3.4.9)

Exercise: Show that S(ρ̂A ) = S(ρ̂B ).


The closest to the maximum value for the entanglement entropy, the closest the state will be to maximal
entanglement. The entanglement entropy is zero for separable states. In our example, for the separable state
we found that the partial states are pure, therefore the entanglement entropy would be zero, whereas for the
entangled state the partial states were maximally mixed, therefore the entanglement entropy is maximal (this
was a maximally entangled state).
This applies only to pure joint states. For entangled mixed states, the entanglement entropy no longer
serves as a measure of entanglement. For pure states the entanglement entropy (entropy of the reduced states
of A or B) is a natural measure of entanglement which have also a well understood physical interpretation,
but it does not fulfill the previous axioms for non-pure states.

3.4.2 Peres criterion

To account for the entanglement of general states let us introduce the partial transpose density matrix. For
a general density matrix of a bipartite system AB
X
ρAB = ρijkl |iiA |jiB hk|A hl|B , (3.4.10)
ijkl

Quantum Dynamics 17 Eduardo Martı́n-Martı́nez


Version 0.2g Eduardo Martı́n-Martı́nez - AMATH 473

the partial transpose is defined as


ρpT
X
B
AB = ρijkl |iiA |liB hk|A hj|B (3.4.11)
ijkl

or, equivalently for our purposes, as

ρpT
X
A
AB = ρijkl |kiA |jiB hi|A hl|B . (3.4.12)
ijkl

There is a theorem for the lower dimensional cases, for bipartite systems of dimension 2 × 2 (two-qubit
states) and 3×2 (qutrit-qubit states) the well-known Peres criterion [?] guarantees that a state is non-separable
(and therefore, entangled) if, and only if, the partial transposed density matrix has, at least, one negative
eigenvalue.
Unfortunately, for higher dimensions the condition is no longer necessary and sufficient, but only suffi-
cient due to the existence of bound entanglement: there are states which are entangled, but that cannot be
distilled, i.e. no pure entangled states can be obtained from them by means of local operations and classical
communication (LOCC). Such states are called bound entangled states [?] and its entanglement is of no utility
to quantum information tasks. Peres criterion only accounts for the existence of entanglement that can be
distilled and therefore useful to perform quantum information tasks. In this book we will only be interested
in distillable entanglement, so in principle we will not need to worry about the existence or not of bound
entanglement.

3.4.3 Negativity

Based on Peres criterion a number of entanglement measures have been introduced. In this book we will
use negativity (N ) to account for the quantum correlations between the different bipartitions of the system
[?]. It is an entanglement monotone sensitive to distillable entanglement defined as the sum of the negative
eigenvalues of the partial transpose density matrix, in other words, if σi are the eigenvalues of any ρpT
AB then

1X X
NAB = (|σi | − σi ) = − σi . (3.4.13)
2
i σi <0

The minimum value of negativity is zero (for states with no distillable entanglement) and its maximum (reached
for maximally entangled states) depends on the dimension of the maximally entangled state. Specifically, for
max = 1/2.
qubits NAB

3.4.4 Concurrence and entanglement of formation

Another relevant measure of entanglement that is commonly employed is the concurrence. The concurrence
has the feature that it allows us to quantify the amount of entanglement in terms of maximally entangled
pairs of qubits (ebits) [?, ?]. For an arbitrary two-qubit state ρ the concurrence is equal to

C (ρ) = max (0, λ1 − λ2 − λ3 − λ4 ) , (3.4.14)

where the λi ’s are the square roots of the eigenvalues of the matrix ρρ̃, where ρ̃ = (σy ⊗ σy )ρ∗ (σy ⊗ σy ) with
σy being the Pauli y matrix, ordered such that λ1 ≥ λ2 ≥ λ3 ≥ λ4 [?].

Quantum Dynamics 18 Eduardo Martı́n-Martı́nez


Version 0.2g Eduardo Martı́n-Martı́nez - AMATH 473

In a general two-qubit system concurrence and negativity are related by the following inequality [?]
p
C > 2N > (1 − C)2 + C 2 − (1 − C). (3.4.15)

The negativity is equal to (half) the concurrence if the eigenvector of the partially transposed state ρΓAB
A

corresponding to its negative eigenvalue is one of the Bell states (up to local unitary transformations).
For a system of two qubits, the concurrence can be used to calculate the entanglement of formation EF ,
defined as the number of the maximally entangled states needed to prepare ρAB [?, ?]
 q 
1 + 1 − C (ρAB )2
EF (ρAB ) = h   (3.4.16)
2

where h(x) = −x log2 x−(1−x) log2 (1−x). This gives the amount of standard units of a bipartite entanglement,
commonly known as ebits.

3.4.5 Mutual information

The mutual information of two random variables (X, Y ) is a function of these two variables that measures
how much uncertainty about one of the variables is reduced by our knowledge about the other. It accounts
for the correlations, both quantum and classical between the two variables.
Given two random variables (X, Y ) the mutual information IXY is defined as
I(X, Y ) = H(X) + H(Y ) − H(X, Y ), (3.4.17)
where H(X), H(Y ) are the so-called marginal Shanon entropies and H(X, Y ) the joint entropy defined as
X
H(X) = − P (x) log2 [P (x)] , (3.4.18)
x
X
H(Y ) = − P (y) log2 [P (y)] , (3.4.19)
y
X
H(X, Y ) = − P (x, y) log2 [P (x, y)] , (3.4.20)
x,y

where P (x, y) is the joint probability distribution of the random variables X, Y and
X X
P (x) = P (x, y), P (y) = P (x, y) (3.4.21)
y x

are the marginal probability distributions for X and Y .


For a quantum bipartite system of density matrix ρAB the quantum mutual information is expressed in
terms of the Von Neumann Entropy
IAB = SA + SB − SAB , (3.4.22)
where the Von Neumann entropies are5
SAB = − TrAB (ρAB log2 ρAB ) , (3.4.23)
SA = − TrA (ρA log2 ρA ) , (3.4.24)
SB = − TrB (ρB log2 ρB ) , (3.4.25)
5
The log2 is chosen to be base 2 because in quantum information it is common to work with qubits, but any other basis can
be chosen instead.

Quantum Dynamics 19 Eduardo Martı́n-Martı́nez


Version 0.2g Eduardo Martı́n-Martı́nez - AMATH 473

and the partial systems are ρA = TrB (ρAB ), ρB = TrA (ρAB ).


Mutual information accounts for both, classical and quantum correlations, so that it can be used together
with an entanglement measure to distinguish the behaviour of classical correlations: in a system which has no
quantum correlations, mutual information accounts exclusively for classical correlations.

3.5 An application of entanglement: Quantum Teleportation

Entanglement is known to be a resource for quantum computing, cryptography and communication. In


particular, we are going to see a protocol of quantum communication known as ‘quantum teleportation’.
Quantum teleportation is a protocol by which quantum information (e.g. the exact quantum state of an atom,
photon, etc) can be transmitted from one location to another, with the help of classical communication and
previously shared quantum entanglement between the sending and receiving location. It was first proposed
in 1993 by Bennet, Brassard, Crepeau, Jozsa, Peres and Wootters [?], and first experimentally tested in 1997
[?].
Let us analyze the protocol step by step for qubits:
Step 0 - A and B share a Bell pair, for example
1 h i
|Φ+ iA2 B3 = √ |00iA2 B3 + |11iA2 B3 . (3.5.1)
2
A and B prepare this bell pair and then B goes away. After that Alice is given a qubit |ϕiA1 that she wants
to send to Bob:
|ϕiA1 = α0 |0i + α1 |1i. (3.5.2)

Step 1- Alice will make a joint measurement of her two qubits (the one she wants to teleport, A1 , and her
own half of the entangled pair, A2 ) in a Bell basis. For that we first write the tripartite system of the three
qubits:

|ψiA1 A2 B3 = |ϕiA1 ⊗ |Φ+ iA2 B3


  1 h i
= α0 |0iA1 + α1 |1iA1 ⊗ √ |00iA2 B3 + |11iA2 B3
2
α0   α 
1

= √ |000iA1 A2 B3 + |011iA1 A2 B3 + √ |100iA1 A2 B3 + |111iA1 A2 B3 . (3.5.3)
2 2
If we factor out explicitly Alice’s two qubits we get
α0  
|ψiA1 A2 B3 = √ |00iA1 A2 ⊗ |0iB3 + |01iA1 A2 ⊗ |1iB3
2
α1  
+ √ |10iA1 A2 ⊗ |0iB3 + |11iA1 A2 ⊗ |1iB3 . (3.5.4)
2
As mentioned in the previous section, the four Bell pairs form an orthonormal basis of the space of two qubits.
We can rewrite Alice’s two qubits in the Bell basis, noticing that
1   1  
|00i = √ |Φ+ i + |Φ− i , |11i = √ |Φ+ i − |Φ− i , (3.5.5)
2 2
1   1  
|01i = √ |Ψ+ i + |Ψ− i , |10i = √ |Ψ+ i − |Ψ− i (3.5.6)
2 2

Quantum Dynamics 20 Eduardo Martı́n-Martı́nez


Version 0.2g Eduardo Martı́n-Martı́nez - AMATH 473

and use it to rewrite (3.5.4) as


1  
|ψiA1 A2 B3 = |Φ+ iA1 A2 ⊗ α0 |0iB3 + α1 |1iB3
2
1 −  
+ |Φ iA1 A2 ⊗ α0 |0iB3 − α1 |1iB3
2
1 +  
+ |Ψ iA1 A2 ⊗ α0 |1iB3 + α1 |0iB3
2
1 −  
+ |Ψ iA1 A2 ⊗ α0 |1iB3 − α1 |0iB3 . (3.5.7)
2
If Alice measures her states in the Bell basis, Bob’s states get projected with equal probability to one of the
following four states (given in matrix representation):
       
α0 α0 α1 −α1
|ϕiB3 = , , , . (3.5.8)
α1 −α1 α0 α0

They are related with the original qubit that Alice wanted to teleport |ϕiA1 by simple local operations.
Step 3- Alice announces the result of her measurement to Bob through a classical channel (2 classical bits).
With the obtained information, Bob can recover, through local unitary operations, the quantum state that
Alice wanted to teleport. In particular

Local
A measured B has B does operation
used
      
α0 1 0 α0 α0
|Φ+ i |ϕiB3 = |ϕ0 i B3 = = = |ϕiA1 11
α1 0 1 α1 α1
      
α0 1 0 α0 α0
|Φ− i |ϕiB3 = |ϕ0 i B3 = = = |ϕiA1 σz
−α1 0 −1 −α1 α1
      
α1 0 1 α1 α0
|Ψ+ i |ϕiB3 = |ϕ0 i B3 = = = |ϕiA1 σx
α0 1 0 α0 α1
      
−α1 0 1 −α1 α0
|Ψ− i |ϕiB3 = |ϕ0 i B3 = = = |ϕiA1 iσy
α0 −1 0 α0 α1

At the end of the protocol, Bob ends up with a state which is identical to the state of A1 that Alice initially
had. What happened is that the subsystem B3 has acquired the state A1 . The Bell state that Alice and Bob
shared is destroyed applying this protocol, as it is the state of A1 . Causality is preserved by the fact that Bob
needs the information input about the outcome of Alice, or otherwise he is unable to know which operation to
perform to recover the original qubit. Also, it is very simple to prove that it is impossible to clone quantum
states (for further reference, see the no-cloning theorem [?]), so to teleport one has first to destroy the original.

Quantum Dynamics 21 Eduardo Martı́n-Martı́nez

You might also like