You are on page 1of 31

Composite quantum systems and

9 entanglement

And yet you have no moral pangs in asking..., how can the first half of the
pair (here) be influenced by the apparatus which measures the second half of
the pair (there). After you know that each particle is stripped by quantum
theory of all its classical attributes (precise position, momentum, ...), you still
believe that it retains a well-defined ”existence,” as a separate entity.
Peres (1984)

9.1 Tensor product

9.1.1 Definition

We The composition of subsystems in quantum theory is described in terms of a math-


temporarily ematical operation known as tensor product. The notion of the tensor product orig-
abandon the
Dirac inates from the following property of matrices.
notation for Consider two vectors
the    
definitions ϕ1 ψ1
of Sec. 9.1.1.  ϕ2   ψ2 
ϕ=  ∈ Cn ,
 ...  ψ =  
 ...  ∈ C .
m
(9.1)
ϕn ψm

Using these vectors we can construct a n × m matrix, denoted by ϕ ⊗ ψ, with


elements (ϕ ⊗ ψ)ij := ϕi ψj , where i = 1, 2, . . . , n in j = 1, 2, . . . , m, that is,
 
ϕ 1 ψ1 ϕ 2 ψ1 ... ϕ n ψ1
 ϕ 1 ψ2 ϕ 2 ψ2 ... ϕ n ψ2 
ϕ⊗ψ =
 ...
 (9.2)
... ... ... 
ϕ 1 ψm ϕ 2 ψm ... ϕ n ψm

We define the tensor product Hilbert space Cn ⊗ Cm as the Hilbert space with
vectors all linear combinations of the form
X
r
ϕ(k) ⊗ ψ (k) ,
k=1

for any r-tuples of vectors ϕ(k) ∈ Cn and ψ (k) ∈ Cm , where k = 1, 2, . . . , r. Note


that the index k is inside a parenthesis because it labels the selected vectors and
235
A
236 Composite quantum systems and entanglement

Box 9.1 Definition of the tensor product


First, we define the free vector space F (S) associated to a set S: F (S) is a complex vector
space, characterized by a basis |s⟩ where s runs over all elements of S. Obviously, if S
has N elements then F (S) = Cn .
Consider two Hilbert spaces H1 and H2 , with vectors expressed as |ϕ⟩1 and |ψ⟩2 ,
respectively. The Cartesian product H1 × H2 consists of pairs (|ϕ⟩, |ψ⟩). The free vector
space F (H1 × H2 ) consists of elements of the form

α = ri=1 ci (|ϕi ⟩, |ψi ⟩) = c1 (|ϕ1 ⟩, |ψ1 ⟩)+c2 (|ϕ2 ⟩, |ψ2 ⟩)+. . .+cr (|ϕr ⟩, |ψr ⟩), for ci ∈ C.
We define a norm on F (H1 × H2 ) by
∑ ∑ ∑
∥ ri=1 ci (|ϕi ⟩, |ψi ⟩)∥2 = ri=1 rj=1 ci c∗j ⟨ϕj |ϕi ⟩⟨ψj |ψi ⟩.
This definition is improper because there are vectors α in α ∈ F (H1 × H2 ), such that
∥α∥ = 0. We obtain a proper norm by identifying any two vectors α, β ∈ F (H1 × H2 )
if∥α − β∥ = 0. To be precise, we define the equivalence class [α] as the set of all vectors
β ∈ F (V1 × V2 ) that satisfy ∥α − β∥ = 0. By the triangle inequality, |∥α − β∥ − ∥α∥| ≤
∥β∥ ≤ ∥α − β∥ + ∥α∥. It follows that ∥β∥ = ∥α∥, i.e., all vectors in the equivalence
class [α] have the same norm. Hence, the norm ∥[α]∥ of the equivalence class [α] is well
defined.
The tensor product H1 ⊗ H2 is the vector space of all equivalence classes [α]. The
norm ∥[α]∥ defines an inner product through the polarization identity (Prop. 4.7). We
denote the equivalence class [(|ϕ⟩, |ψ⟩)] as |ψ⟩ ⊗ |ϕ⟩ ∈ H1 ⊗ H2 .

not the components of a vectors, like the indices of Eq. (9.1). The inner product on
Cn ⊗ Cm is first defined for vectors ϕ ⊗ ψ as

(ϕ ⊗ ψ, ϕ′ ⊗ ψ ′ )Cn ⊗Cm = (ϕ, ϕ′ )Cn × (ψ, ψ ′ )Cn , (9.3)

and then extended to any linear combination of the form (9.3) by linearity. For
example,

(ϕ(1) ⊗ ψ (1) + ϕ(2) ⊗ ψ (2) , ϕ′ ⊗ ψ ′ ) = (ϕ(1) , ϕ)(ψ (1) , ψ ′ ) + (ϕ(2) , ϕ′ )(ψ (2) , ψ ′ ).

Any n×m matrix can be written as a sum (9.3), hence, the Hilbert space Cn ⊗Cm
is isometric to the Hilbert space Cnm of n × m matrices,

Cn ⊗ Cm = Cnm . (9.4)

The preceding description suffices for the definition of the tensor product H1 ⊗H2 ,
for finite-dimensional Hilbert space H1 and H2 . If at least one of the Hilbert spaces
is infinite dimensional, then a more intricate definition is required. It is described in
Box 9.1. The definition of a tensor product of n Hilbert spaces H1 ⊗ H2 ⊗ . . . ⊗ Hn
is a straightforward generalization.

9.1.2 Dirac notation

Dirac notation is very convenient for describing the tensor product. The Hilbert We switch
Pr back to
space H1 ⊗H2 consists of vectors k=1 |ϕk ⟩⊗|ψk ⟩, where |ϕk ⟩ ∈ H1 and |ψk ⟩ ∈ H2 ,
Dirac
notation.
A
237 Tensor product

while the inner product is expressed as


 ′  !
Xr Xr X r′
r X
 ⟨ϕl | ⊗ ⟨ψl | |ϕk ⟩ ⊗ |ψk ⟩ = ⟨ϕl |ϕk ⟩⟨ψl |ψk ⟩. (9.5)
l=1 k=1 k=1 l=1

Vectors of the form |ϕ⟩ ⊗ |ψ⟩ ∈ H1 ⊗ H2 are called separable. We will often write
the vectors |ϕ⟩ ⊗ |ψ⟩ as |ϕ, ψ⟩, but this notation requires prior explanation in order
to avoid ambiguities.
If the vectors |n⟩ define an orthonormal basis on H1 and the vectors |i⟩ define
an orthonormal basis on H2 , then the vectors |n, i⟩ := |n⟩ ⊗ |i⟩ define the induced
orthonormal basis on H1 ⊗ H2 . Any vector |Ψ⟩ ∈ H1 ⊗ H2 can be expressed as
P
|Ψ⟩ = n,i an,i |n, i⟩, where

an,i = ⟨n, i|Ψ⟩ (9.6)

The resolution of the unity in H1 ⊗ H2 is


X
Iˆ = |n, i⟩⟨n, i|. (9.7)
n,i

The tensor product and the direct sum satisfy the distributive property,

H1 ⊗ (H2 ⊕ H3 ) = (H1 ⊗ H2 ) ⊕ (H1 ⊗ H2 ). (9.8)

To show this, one selects orthonormal basis in each Hilbert space H1 , H2 and H3 .
Then, it is straightforward to show that the induced basis on H1 ⊗ (H2 ⊕ H3 )
coincides with the induced basis on (H1 ⊗ H2 ) ⊕ (H1 ⊗ H2 ).

9.1.3 Tensor product of operators

Let  be an operator on the Hilbert space H1 and B̂ an operator on the Hilbert


space H2 . We define the operator  ⊗ B̂ on the Hilbert space H1 ⊗ H2 by
X X
(Â ⊗ B̂) |ϕi ⟩ ⊗ |ψi ⟩ = Â|ϕi ⟩ ⊗ B̂|ψi ⟩. (9.9)
i i

The definition implies that

(Â ⊗ B̂)(Ĉ ⊗ D̂) = ÂĈ ⊗ B̂ D̂. (9.10)

Note: we use the same symbol (⊗) for the tensor product of Hilbert spaces, for
the factorized vectors and for the tensor product of operators. Strictly speaking,
these are distinct mathematical operations, and the use of the same symbol for all
three of them in an abuse of notation.
We will also use the same symbol (I)ˆ for the unit operator in H1 , in H2 and in
H1 ⊗ H2 . With this convention, equations of the form Iˆ = Iˆ ⊗ Iˆ are valid.
An operator on H1 ⊗ H2 that can be written as  ⊗ B̂ is called factorizable.
Any operator X̂ on H1 ⊗ H2 can be expressed as a sum of factorizable operators
P
X̂ = i Âi ⊗ B̂i , where Âi are operators on H1 and B̂i are operators on H2 .
A
238 Composite quantum systems and entanglement

Let |n⟩ and |i⟩ be orthonormal bases in the Hilbert spaces H1 and H2 , respec-
tively. Writing |n, i⟩ = |n⟩ ⊗ |i⟩, we find that
X X X
T r(Â ⊗ B̂) = ⟨n, i|Â ⊗ B̂|n, i⟩ = ⟨n|Â|n⟩ · ⟨i|B̂|i⟩ = T r(Â)T r(B̂)
n,i n i

9.2 Quantum combination of subsystems

9.2.1 Fundamental rule

First, we examine how the combination of subsystems is effected in classical physics.


Consider two systems with state spaces Γ1 and Γ2 . For example, if both systems
are particles in one dimension, the elements of Γ1 are pairs (q1 , p1 ) of the particle’s
position and momentum. The elements of Γ2 are similar. We have full information
about the composite system of the two particles if we know the positions and
momenta of both particles. Hence, the state space of the two particle system consists
of elements (q1 , q2 , p1 , p2 ), i.e., it coincides with Γ1 × Γ2 .
In classical physics, the state space of a composite system is the Cartesian product
Γ1 × Γ2 × . . . × Γn of the state spaces Γi of its components.
To find the corresponding law in quantum theory, we consider two quantum sys-
tems described by the Hilbert spaces H1 and H2 , and with associated Hamiltonians
Ĥ1 and Ĥ2 . Let En1 and |n⟩1 be the eigenvalues and eigenvectors of Ĥ1 , and En2
and |n⟩2 the eigenvalues and eigenvectors of Ĥ2 . If the two systems do not interact,
we expect that the composite system will be described by a Hilbert space H, and
a Hamiltonian Ĥ. The eigenvalues Ĥ must be of the form En1 + En2 ′ , for different
values of n and n′ , because energy is an additive quantity.
The only way to satisfy this condition is by choosing the Hamiltonian

Ĥ = Ĥ1 ⊗ Iˆ + Iˆ ⊗ Ĥ2 (9.11)

on the Hilbert space H1 ⊗ H2 . The eigenvectors of Ĥ are |n, n′ ⟩ = |n⟩ ⊗ |n′ ⟩, since

Ĥ|n, n′ ⟩ = Ĥ1 ⊗ I|n,


ˆ n′ ⟩ + Iˆ ⊗ Ĥ2 |n, n′ ⟩ = Ĥ1 |n⟩ ⊗ I|n
ˆ ′ ⟩ + I|n⟩
ˆ ⊗ Ĥ2 |n′ ⟩
= En1 |n, n′ ⟩ + En2 ′ |n, n′ ⟩ = (En1 + En2 ′ )|n, n′ ⟩. (9.12)

The Hamiltonian (9.11) can also be obtained by assuming a factorizable evolution


operator Ût = e−itĤ1 ⊗ e−itĤ2 . By Leibniz’s rule,
d
i Ût = Ĥ1 e−itĤ1 ⊗ e−itĤ2 + e−itĤ1 ⊗ Ĥ2 e−itĤ2
dt
= (Ĥ1 ⊗ Iˆ + Iˆ ⊗ Ĥ2 )Ût . (9.13)
d
Since time evolution operators satisfy i dt Ût = Ĥ Ût , we confirm Eq. (9.11) for the
Hamiltonian of the composite system.
We conclude that subsystems in quantum theory must be composed through the
A
239 Quantum combination of subsystems

tensor product. We will refer to composite systems with n components as n-partite


system. Then, the following principle applies.

FP6. A n-partite system is described by the tensor product H1 ⊗H2 ⊗. . .⊗Hn


of the Hilbert spaces Hi , i = 1, 2, . . . , n of its subsystems.

The tensor products dramatically increases the number of available states in a


composite system. Consider two subsystems described by the Hilbert space Cn . The
space of pure states for each system has real dimension 2(n−1). In classical physics,
the state space for the composite system would have dimension 4(n−1). The Hilbert
2
space of the composite system Cn ⊗ Cn = Cn corresponds to a space of pure states
of dimension 2(n2 − 1), i.e., of 2n2 − 2 − 4(n − 1) = 2(n − 1)2 more dimensions than
the space obtained through the classical rule for combining subsystems!
If a particle is described by wave functions ϕ1 (x), (x ∈ R), and another particle
by wave functions ϕ2 (y) (y ∈ R), then the composite system of the two particles
is described by a wave function ψ(x, y), i.e, a function that takes pairs (x, y) ∈ R2
to the complex numbers. This implies a natural identification of Hilbert spaces
L2 (R, dx) ⊗ L2 (R, dy) = L2 (R2 , dxdy). In general,

L2 (Rn , dn x) ⊗ L2 (Rm , dm y) = L2 (Rn+m , dn xdn y). (9.14)

In a bipartite system with two identical subsystems, each described by the Hilbert
space H, it is convenient to define the operators Â1 := Â ⊗ Iˆ and Â2 := Iˆ ⊗ Â on
H ⊗ H, for any operator  in H. Hence, for two particles, we write the position
operator of the first particle as x̂1 = x̂ ⊗ Iˆ and of the second particle as x̂2 = Iˆ ⊗ x̂.

9.2.2 Entanglement

Consider a bipartite system with Hilbert space H = H1 ⊗ H2 . Pure states on H of


the form |ϕ⟩ ⊗ |ψ⟩, where |ϕ⟩ ∈ H1 and |ψ⟩ ∈ H2 , are called separable. Any pure
state that is not separable is called entangled. Entanglement of states is a unique
characteristic of quantum systems, with no analogue in classical physics.
We select an orthonormal basis |i⟩ on H1 , an orthonormal basis |n⟩ on H2 and
the orthonormal basis |i, n⟩ = |i⟩ ⊗ |n⟩ on H1 ⊗ H2 . For any density matrix ρ̂ on
H1 ⊗ H2 , measurements on the first subsystem H1 are described by projectors of
the form P̂ ⊗ Iˆ and probabilities
h i X XX
T r ρ̂(P̂ ⊗ I)
ˆ = ⟨n, i|ρ̂P̂ ⊗ I|n,
ˆ i⟩ = ⟨n, i|ρ̂|m, j⟩⟨m, j|P̂ ⊗ I|n,
ˆ i⟩
i,n i,n j,m
XX X
= ⟨n, i|ρ̂|m, j⟩⟨m|P̂ |n⟩⟨i|j⟩ = ⟨n|ρ̂1 |m⟩⟨m|P̂ |n⟩ = T r(ρ̂1 P̂ ). (9.15)
i,n j,m n,m

In Eq. (9.15) we defined the first partial trace of ρ̂ as a density matrix ρ1 on H1


A
240 Composite quantum systems and entanglement

with elements
X
⟨n|ρ̂1 |m⟩ = ⟨n, i|ρ̂|m, i⟩. (9.16)
i

ρ̂1 contains all information of the composite system that is accessible by measure-
ments on the first subsystem.
Similarly, we define the second partial trace of ρ̂ as the density matrix ρ̂2 on H2 ,
with elements
X
⟨i|ρ̂2 |j⟩ = ⟨n, i|ρ̂|n, j⟩. (9.17)
n

We will often write ρ̂1 = T rH2 ρ̂ and ρ̂2 = T rH1 ρ̂. The states ρ̂1 and ρ̂2 are also
called reduced density matrices.
For pure density matrices ρ̂ = |Ψ⟩⟨Ψ| on H, the following holds. If |Ψ⟩ is separable
(i.e., of the form |ϕ⟩ ⊗ |ψ⟩), then the two partial traces of ρ̂ are pure states: ρ̂1 =
|ϕ⟩⟨ϕ| and ρ̂2 = |ψ⟩⟨ψ|. However, if |Ψ⟩ is entangled, then the partial traces are
mixed states.
This remark is made precise by the following theorem—see Box 9.2 for proof.

Theorem 9.1 (Schmidt decomposition.) All vectors |Ψ⟩ ∈ H1 ⊗ H2 can be


expressed as
X
N
|Ψ⟩ = ci |ϕi ⟩ ⊗ |ψi ⟩, (9.18)
i=1

where ci > 0, the vectors |ϕi ⟩, i = 1, 2, . . . , N define an orthonormal set on H1 , and


the vectors |ψi ⟩, i = 1, 2, . . . , N define an orthonormal set on H2 . The integer N
is smaller than or equal to the smallest among the dimensions of H1 and H2 ; it is
called Schmidt order of the vector |Ψ⟩. The numbers ci are unique, the orthonormal
sets |ϕi ⟩ and |ψi ⟩ are not.

Obviously, if N = 1, then |Ψ⟩ is separable, and if N > 1, |Ψ⟩ is entangled. The


partial traces of the density matrix |Ψ⟩⟨Ψ| are .
X
N X
N
ρ̂1 = c2i |ϕi ⟩⟨ϕi |, ρ̂2 = c2i |ψi ⟩⟨ψi |. (9.19)
i=1 i=1
PN 4
The two partial traces have the same purity γ = i=1 ci . The quantity
E := 1 − γ (9.20)
takes values in the interval (0, 1], and it provides a measure of entanglement, as it
vanishes for all separable states1 . We will refer to E as quadratic entanglement, as
it is quadratic to the reduced state.
1 A better measure of∑ entanglement for pure states is the von Neumann entropy of ρ̂1 : S[ρ̂1 ] =
−T r(ρ̂1 log ρ̂1 ) = − N 4 2
i=1 ci log ci . For properties of the von Neumann entropy, see Box 6.3.
A
241 Quantum combination of subsystems

If N = dim H1 = dim H2 , and a state |Ψ⟩ is characterized by c1 = c2 = . . . = cn =


1
N, the partial traces are states of maximal ignorance, i.e., the most mixed states.
Then, the vector |Ψ⟩ is a maximal-entanglement state. For such states, E = 1−N −1 .
Note that any unitary transformation of the form Û1 ⊗ Û2 only changes the
orthonormal sets |ϕi ⟩ and |ψi ⟩ in the Schmidt decomposition. The values of the
ci remain unchanged. Hence, any entanglement measure defined solely from the
ci , like the quadratic entanglement E of Eq. (9.20) remains invariant under the
transformations |Ψ⟩ → Û1 ⊗ Û2 |Ψ⟩.
If a composite system is entangled, then its state cannot be reconstructed solely
from the knowledge of the states of its constituents. Since the reduced density ma-
trices of the latter are mixed, they have less information than maximal, even if the
state vector of the combined system is pure, hence, containing maximal information.
For this reason,

...the phenomenon of quantum entanglement suggests a holistic structure for the physi-
cal world that contrasts strongly with the predominantly reductionist views of Western
philosophy whereby composite systems may be analysed in terms of their constituent sub-
systems. It also seems at variance with our actual experience of how large objects behave:
a behaviour that is arguably one of the main reasons why Western philosophy has evolved
the way it has. (Isham, 1995)

9.2.3 Entanglement of mixed states

The notion of entanglement also applies to mixed states. Factorized density matrices
obviously correspond to separable states. Since mixing is a classical operation, we
expect that mixtures of factorized states will also behave classically. With this
reasoning, we call a density matrix ρ̂ on the Hilbert space H1 ⊗ H2 separable, if it
can be expressed as a mixture of factorized density matrices, i.e., if
X
r
(k) (k)
ρ̂ = λk ρ̂1 ⊗ ρ̂2 , (9.21)
k=1
Pr
where λk ≥ 0 and k=1 λk = 1. The maximal-ignorance density matrix on H1 ⊗H2
is obviously separable. Any non-separable density matrix is entangled.
P (k)
The first partial trace of the separable state (9.21) is ρ̂1 = k λk ρ̂1 , hence,
X
r
(k) (k)
ρ̂1 ⊗ Iˆ − ρ̂ = λk λk ρ̂1 ⊗ (Iˆ − ρ̂2 ) ≥ 0, (9.22)
k=1

(k) (k)
where the inequality in the last steps follows because ρ̂1 ≥ 0 and Iˆ − ρ̂2 ≥ 0. It
follows that a density matrix ρ̂ that fails the inequality (9.22) is entangled.
The violation of inequality (9.22) is a sufficient but not necessary condition for
entanglement, i.e., any state that violates it is entangled, but not all entangled
states violate it.
Another widely used entanglement criterion is due to Peres (1996) and Horodecki
A
242 Composite quantum systems and entanglement

Box 9.2 Proof of the Schmidt decomposition


The Schmidt decomposition is a special case of a key result of matrix theory, the singular
value expansion.
Singular value expansion. Any n × m complex matrix A can be expressed as
A = U ΣV † , where (i) U is a unitary n × n matrix, (ii) Σ is a diagonal n × m matrix,
whose elements are either positive numbers or zero, and (iii) V is a unitary m × m
matrix.
The unitary matrices U and V are not unique, the diagonal elements of Σ are unique
up to rearrangement. The latter are called singular values of A.
Consider two orthonormal bases |i⟩ and∑ |k⟩ in∑Cn and in Cn , respectively. A vector
|Ψ⟩ ∈ Cn ⊗ Cm can be expressed as |Ψ⟩ = n i=1 k=1 Cik |i⟩ ⊗ |k⟩, in terms of a n × m
m

matrix Cik . We use the singular value expansion for C, to obtain


∑ ∑m ∗
Cik = n j=1 l=1 Uij Σjl Vkl . (A)
Without loss of generality, we assume n ≤ m, hence, the matrix Σ is of the form
 
c1 0 0 ... 0 0 0 ... 0
 0 c2 0 ... 0 0 0 ... 0 
Σ=
... ... ... ... ... ... ... ... ... 
,
0 0 0 . . . cn 0 0 ... 0
where c1 , c2 , . . . , cn are the singular values of C. We order ci so that they appear in
decreasing order in the diagonal. Summation in Eq. (A) runs for j, l = 1, 2, . . . N , where
∑ ∗
N is the number of non-zero singular values ci . Hence, Cik = N j=1 cj Uij Vkj .
We conclude that
∑ ∑n ∑m ∗ ∑N
|Ψ⟩ = N j=1 i=1 k=1 cj Uij Vkj |i⟩ ⊗ |k⟩ = i=1 ci |ϕi ⟩ ⊗ |ψi ⟩,
∑ ∑m ∗
where |ϕj ⟩ = n U
i=1 ij |i⟩ and |ψ j ⟩ = V
k=1 kj |k⟩ define orthonormal sets on Cn and
on Cm respectively, since the matrices U and V are unitary.

et al. (1996). It is usually called PPT criterion where PPT stands for Positive Partial
Transpose.
Let |n⟩ be an orthonormal basis on the Hilbert space H1 and |i⟩ an orthonormal
basis on the Hilbert space H2 . The partial transpose transformation T2 swaps the
matrix elements of a density matrix ρ̂ on H1 ⊗ H2 with respect to the basis |i⟩,
⟨m, j|T2 [ρ̂]|n, i⟩ = ⟨m, i|ρ̂|n, j⟩. (9.23)
For a separable density matrix (9.21),
X
r
(k) (k)
T2 [ρ̂] = λk ρ̂1 ⊗ [ρ̂2 ]T , (9.24)
k=1

where by upper-case T we denote the transpose of an operator. The transpose of a


density matrix is still a density matrix. Hence, the operation T2 upon a separable
density matrix ρ̂ gives another (separable) density matrix T2 [ρ̂].
In general, entangled density matrices are not mapped to density matrices by
T2 . To see this, consider a pure entangled state ρ̂ = |Ψ⟩⟨Ψ|. We choose |Ψ to be
of Schmidt order 2, so that |Ψ⟩ = cos θ|1⟩ ⊗ |a⟩ + sin θ|2⟩ ⊗ |b⟩, where |1⟩ and |2⟩
form an orthonormal set in H1 , |a⟩ and |b⟩ form an orthonormal set in H2 , and
A
243 The two-qubit system

θ ∈ [0, π2 ]. Then,

ρ̂ = cos2 θ|1⟩⟨1| ⊗ |a⟩⟨a| + sin2 θ|2⟩⟨2| ⊗ |b⟩⟨b|


+ sin θ cos θ (|1⟩⟨2| ⊗ |a⟩⟨b| + |2⟩⟨1| ⊗ |b⟩⟨a|) . (9.25)

Hence,

T2 [ρ̂] = cos2 θ|1⟩⟨1| ⊗ |a⟩⟨a| + sin2 θ|2⟩⟨2| ⊗ |b⟩⟨b|


+ sin θ cos θ (|1⟩⟨2| ⊗ |b⟩⟨a| + |2⟩⟨1| ⊗ |a⟩⟨b|) . (9.26)

It is straightforward to verify that the vector |Φ⟩ = √1


2
(|1⟩ ⊗ |b⟩ − |2⟩ ⊗ |a⟩) is an
eigenvector of T2 [ρ̂] with negative eigenvalue,

T2 [ρ̂]|Φ⟩ = − sin θ cos θ|Φ⟩. (9.27)

Hence, T2 [ρ̂] is not positive.

The PPT criterion asserts that any density matrix ρ̂ on H1 ⊗ H2 , with a negative
eigenvalue for T2 [ρ̂] is entangled. In general, the PPT criterion is sufficient but not
necessary for entanglement. It can be proven that for bipartite systems with Hilbert
spaces C2 ⊗ C2 and C2 ⊗ C3 , PPT is both sufficient and necessary.
The absolute value of the smallest negative eigenvalue λmin of T2 [ρ̂] can be used to
quantify entanglement. For example, the reduced density matrix associated to ρ̂ of
Eq. (9.25) has matrix elements ⟨1|ρ̂1 |1⟩ = cos2 θ, ⟨2|ρ̂1 |2⟩ = sin2 θ, and ⟨1|ρ̂1 |2⟩ = 0.
Hence, the quadratic entanglement (9.20) is E = 1 − cos4 θ − sin4 θ = sin2 θ cos2 θ =
|λmin |2 .
There is no standard mathematical measure of entanglement for mixed states.
There are several different proposals for such measures, but they are all hard to
compute. All such measures must satisfy two crucial properties. First, they must be
invariant under factorized unitary transformations ρ̂ → (Û1 ⊗ Û2 )ρ̂(Û1† ⊗ Û2† ), simi-
larly to the measure (9.20) for pure states. Second, entanglement must not increase
when we extract local information, i.e., information from only one of the subsys-
tems. This means that the state ρ̂′ of the statistical ensemble after the measurement
of an operator  ⊗ Iˆ must have entanglement equal or less than the entanglement
of the state ρ̂ prior to measurement.

9.3 The two-qubit system

The system of two qubits is described by the Hilbert space H = C2 ⊗ C2 .


A
244 Composite quantum systems and entanglement

9.3.1 Maximal-entanglement states

The following vectors define an orthonormal basis on H, known as the Bell basis.
1
|Ψ± ⟩ = √ (|1, 0⟩ ± |0, 1⟩) (9.28)
2
1
|Φ± ⟩ = √ (|1, 1⟩ ± |0, 0⟩) . (9.29)
2
We evaluate the associated partial traces. Eq. (9.16) for the two-qubit system
becomes

⟨1|ρ̂1 |1⟩ = 1 − ⟨0|ρ̂1 |0⟩ = ⟨1, 1|Ψ⟩⟨Ψ|1, 1⟩ + ⟨1, 0|Ψ⟩⟨Ψ|1, 0⟩


⟨1|ρ̂1 |0⟩ = ⟨1, 1|Ψ⟩⟨Ψ|0, 1⟩ + ⟨1, 0|Ψ⟩⟨Ψ|0, 0⟩, (9.30)

and similarly for ρ̂2 .


Using Eq. (9.30) to the four states |Ψ± ⟩ and |Φ± ⟩, we find that they all satisfy
1
⟨1|ρ̂1 |1⟩ = ⟨0|ρ̂1 |0⟩ = ⟨1|ρ̂1 |0⟩ = 0. (9.31)
2
1
⟨1|ρ̂2 |1⟩ = ⟨0|ρ̂2 |0⟩ = ⟨1|ρ̂2 |0⟩ = 0. (9.32)
2
ˆ The partial traces for all four states |Ψ± ⟩ and |Φ± ⟩ is the
Hence, ρ̂1 = ρ̂2 = 12 I.
state of maximum ignorance for the qubit. Hence, the vectors |Ψ± ⟩ and |Φ± ⟩ define
maximum-entanglement states.

9.3.2 Probabilities and correlations

Next, we construct the probabilities for the measurement of an operator  = m · σ


on the first qubit and of an operator B̂ = n · σ on the second qubit, where m and
n are unit vectors. The associated operators on H are  ⊗ Iˆ and Iˆ ⊗ B̂.
The spectral projectors of  are P̂± = 12 (Iˆ ± m · σ) and the spectral projectors
of B̂ are Q̂± = 12 (Iˆ ± n · σ). The spectral projectors of  ⊗ Iˆ are P̂a ⊗ I,
ˆ where
a = ±1; the spectral projectors of Iˆ ⊗ B̂ are Iˆ ⊗ Q̂b , where b = ±1.
The operators  ⊗ Iˆ and Iˆ ⊗ B̂ commute. Hence, the joint probability of finding
a in the first qubit and b in the second qubit does not depend on the order of the
measurements, and it is given by Eq. (8.15),

Prob(a, b) = T r[ρ̂(P̂a ⊗ I)(


ˆ Iˆ ⊗ Q̂b )] = T r(ρ̂P̂a ⊗ Q̂b ), (9.33)

for all initial states ρ̂.


Example 9.1. We evaluate the probabilities Prob(a, b) for the initial state |Ψ− ⟩ of
Eq. (9.28),
1
Prob(a, b) = ⟨Ψ− |P̂a ⊗ Q̂b |Ψ− ⟩ = ⟨1|P̂a |1⟩⟨0|Q̂b |0⟩ + ⟨0|P̂a |0⟩⟨1|Q̂b |1⟩
2 
−⟨1|P̂a |0⟩⟨0|Q̂b |1⟩ − ⟨0|P̂a |1⟩⟨1|Q̂b |0⟩ . (9.34)
A
245 The two-qubit system

By Eq. (5.18),
1 1 1
⟨0|P̂± |0⟩ = (1 ∓ m3 ) ⟨1|P̂± |1⟩ = (1 ± m3 ) ⟨0|P̂± |1⟩ = ± (m1 + im2 )
2 2 2
1 1 1
⟨0|Q̂± |0⟩ = (1 ∓ n3 ) ⟨1|Q̂± |1⟩ = (1 ± n3 ) ⟨0|Q̂± |1⟩ = ± (n1 + in2 ).
2 2 2
We substitute into Eq. (9.34), to obtain
1
Prob(+, +) = Prob(−, −) = (1 − m · n) (9.35)
4
1
Prob(+, −) = Prob(−, +) = (1 + m · n) (9.36)
4
The correlation Cor(m, n) between a measurement of  = m·σ on the first qubit
and a measurements o B̂ = n · σ on the second qubit,

Cor(m, n) := Cor(Â ⊗ I,
ˆ Iˆ ⊗ B̂). (9.37)

For maximum entanglement states T r(ρ̂Â ⊗ I) ˆ = T r(ρ̂1 Â) = 1


2 T r  = 0, and,
similarly, T r(ρ̂I ⊗ B̂) = 0. Hence, by Eq. (6.33),
ˆ

Cor(m, n) = T r(ρ̂Â ⊗ B̂). (9.38)

Example 9.2. For ρ̂ = |Ψ− ⟩⟨Ψ− |,


1 
Cor(m, n) = ⟨1|Â|1⟩⟨0|B̂|0⟩ + ⟨0|Â|0⟩⟨1|B̂|1⟩ − ⟨1|Â|0⟩⟨0|B̂|1⟩ − ⟨0|Â|1⟩⟨1|B̂|0⟩ .
2
We insert the values of the matrix elements Aij and Bij , to obtain

Cor(m, n) = −m · n. (9.39)

We can also derive Eq. (9.39) as follows. By the spectral theorem  = P̂+ − P̂−
and B̂ = Q̂+ − Q̂− . Hence,

 ⊗ B̂ = P̂+ ⊗ Q̂+ + P̂− ⊗ Q̂− − P̂+ ⊗ Q̂− − P̂− ⊗ Q̂+ . (9.40)

It follows that

Cor(m, n) = Prob(+, +) + Prob(−, −) − Prob(+, −) − Prob(−, +). (9.41)

Then, Eq. (9.36) leads to Eq. (9.39).

9.3.3 Mixed states

An interesting family of entangled mixed states are the Werner states (Werner,
1989)
1−αˆ
ρ̂W = α|Ψ− ⟩⟨Ψ− | + I. (9.42)
4
Ostensibly, the are obtained as a mixture of the maximal entanglement state |Ψ− ⟩
ˆ which is separable.
and the maximal ignorance state 41 I,
A
246 Composite quantum systems and entanglement

We express the Werner state as a matrix in the basis defined by the vectors
|0, 0⟩, |0, 1⟩, |1, 0⟩, |1, 1⟩,
 
1−α 0 0 0
1 0 1 + α −2α 0 
ρ̂W =  
 (9.43)
4 0 −2α 1 + α 0 
0 0 0 1−α

The eigenvalues of the matrix (9.42) are 14 (1 + 3α) and 41 (1 − α) (triple). Hence,
the Werner states are well defined (the eigenvalues are positive) for − 13 ≤ α ≤ 1.
For − 31 ≤ α ≤ 0, the Werner states cannot be interpreted as a mixture of |Ψ− ⟩ and
the maximal ignorance state.
ˆ Hence,
The first partial trace is ρ̂1 = 21 I.
 
1+α 0 0 0
1 0 1 − α +2α 0 
ρ̂1 ⊗ Iˆ − ρ̂W =  . (9.44)
4 0 +2α 1 + α 0 
0 0 0 1−α

The matrix ρ̂1 ⊗ Iˆ − ρ̂W is identical to the matrix (9.43) for α → −α. Hence, its
eigenvalues are 14 (1 − 3α) and 14 (1 + α) (triple). The eigenvalues are both non-
negative only for α ≤ 13 . By the separability criterion (9.22), Werner states with
α > 13 are entangled.

9.4 Bell’s theorem

9.4.1 The EPR paradox

Bell’s theorem (Bell, 1964) is one of the most important results of quantum theory.
It shows that the assumption that measurements reveal pre-existent properties of
quantum systems leads to predictions that are incompatible with quantum theory.
Bell’s theorem originates from a paper by Einstein, Podolski and Rosen (EPR),
who used quantum entanglement in order to argue that quantum theory is incom-
plete (Einstein et al., 1935). The term ”entanglement” did not exist then, in some
sense, the EPR paper is the first study on the implications of entanglement.
In modern terminology, the argument is the following. Consider a system of two
qubits2 prepared in the state |Ψ− ⟩. By Eq. (9.36), the joint probabilities for the
measurement of the same observable (m = n) are
1
Prob(+, +) = Prob(−, −) = 0, Prob(+, −) = Prob(−, +) = . (9.45)
2
The associated conditional probabilities are Prob(+|−) = Prob(−|+) = 1.
2 The original EPR paper involved measurements of position and momentum in a pair of particles.
The simpler argument with qubits first appeared in Secs. 22.15-22.18 of Bohm (1951).
A
247 Bell’s theorem

Suppose then that we measure σ̂3 on the first qubit. If we find ±1, we can infer
that a measurement of σ̂3 on the second qubit will give with certainty ∓1. EPR
then invoked their ”criterion of reality”:
If,without in any way disturbing a system,we can predict with certainty (i.e.,with
probability equal to unity) the value of a physical quantity, then there exists an
element of physical reality corresponding to this physical quantity.
According to this criterion, the value for σ̂3 on the second qubit, is an element of
reality after the measurement of σ̂3 on the first qubit has been carried out, even
if we do not carry out a measurement on the second qubit3 . On the other hand,
had we chosen to measure σ̂1 on the first qubit, the element of reality of the second
qubit would correspond to its value for σ̂1 .
However, if the qubits are far enough, the choice of measurement on the first
qubit, should not affect what is the element of reality in the second qubit. Hence,
the qubit should have a definite value of both σ̂3 and σ̂1 , which is impossible in
quantum theory.
At this point in the book, the reader can easily identify the problem with the
EPR argument. The quantum conditional probabilities refer solely to measurement
outcomes, and not to any properties possessed by the system. Quantum theory
admits no elements of reality in the EPR sense. If quantum elements of reality
exist, they refer to the complex that consists of the microscopic system and the
measuring apparatus, and not solely to the microscopic system (Bohr, 1935).

9.4.2 Bell’s inequality

Bell realized that the EPR’s view of ”elements of reality” leads to concrete exper-
imental predictions that differs from those of quantum theory. This statement is
known as Bell’s theorem.
Again, we consider a two qubit system, subject to measurements of observables
m · σ for each qubit. The unit vectors m correspond to macroscopic parameters
that are specified in advance. The possible measurement outcomes are +1 and −1.
Suppose that the measurements reveal pre-existing properties of the quantum
system, ‘elements of reality’ at a level of description that is deeper, and more
fundamental than quantum theory. These properties must be expressed in terms of
some fundamental variables ξ that define a state space Γ for the two qubit system.
As shown in Sec. 1.5.3, the measurement of any observable corresponds to a function
f : Γ → R. For a dichotomic observable with values ±1, we can restrict to functions
f : Γ → {+1, −1}.
Hence, the measurement of m·σ̂⊗Iˆ corresponds to a function fm 1
: Γ → {+1, −1},
and the measurement of I ⊗ m · σ̂ corresponds to a function fm : Γ → {+1, −1}.
ˆ 2

3 The inference of the value for σ̂3 on the second qubit does not in any sense involve any transfer
of information. Since we know the initial state, we know that complete correlation of the
outcomes of the two measurements. The same applies to classical systems. For example, when
two particles scatter, the measurement of the momentum of one particle in the center-of-mass
frame fully specifies the momentum of the other particle.
A
248 Composite quantum systems and entanglement

Then, we consider the following experiment. A source emits entangled qubit pairs,
and each qubit is driven towards a different apparatus. In qubit 1, we measure m1 ·σ̂,
and in qubit 2, we measure m2 · σ̂. We carry out the measurement in a statistical
ensemble of N >> 1 qubit pairs. For the k-th pair, we denote by ak ∈ {−1, 1} the
outcome of the measurement of the first qubit, and by bk ∈ {−1, 1} the outcome
of the measurement of the second qubit. From these measurement outcomes, we
construct the correlation

1 X
N
Cor(m1 , m2 ) = a k bk (9.46)
N
k=1

which depends on the type of measurement carried out on the two qubits, as de-
scribed by the unit vectors m1 and m2 .
If Eq. Cor(m1 , m2 ) converges at the limit N → ∞, then we can attempt to
reproduce it from a classical probability distribution ρ on the state Γ,
Z
1 2
Cor(m1 , m2 ) = dξρ(ξ)fm 1
(ξ)fm 2
(ξ). (9.47)

Using the same source of qubit pairs, we perform measurements in four different
experiments, each characterized by a different pair of unit vectors for the two qubit
measurements: (i) (m1 , m2 ), (ii) (m1 , n2 ), (iii) (n1 , m2 ), and (iv) (n1 , n2 ).
We evaluate the measured correlation function in each experiment, and then, we
construct the quantity

∆ = Cor(m1 , m2 ) + Cor(m1 , n2 ) + Cor(n1 , m2 ) − Cor(n1 , n2 ). (9.48)

Since the four experiments were carried out with the same preparation of the
system, they must be described by the same probability distribution ρ(ξ). Then,
Z
∆ = dξρ(ξ)Φ(ξ), (9.49)

where
1
Φ := fm f 2 + fm
1 m2
1
f 2 + fn11 fm
1 n2
2
2
− fn11 fn22 (9.50)

is a function of Γ. We write Φ as
1
Φ = fm 1
2
(fm 2
+ fn22 ) + fn11 (fm
2
2
− fn22 ). (9.51)

The possible values of the functions fm 2


2
± fn22 are −2, 0 and +2. We note that if
fm2 + fn2 = ±2, then fm2 − fn2 = 0, and if fm
2 2 2 2 2
2
2
+ fn22 = 0, then fm 2
− fn22 = ±2.
Hence, if |fm
2
2
± fn22 | = 2, then |fm
2
2
∓ fn22 | = 0.
Therefore,

|Φ| ≤ |fm
1
1
||fm
2
2
+ fn22 | + |fn11 ||fm
2
2
− fn22 | = |fm
2
2
+ fn22 | + |fm
2
2
− fn22 | = 2.

We conclude that the measurable quantity ∆, Eq. (9.48) always satisfies

|∆| ≤ 2, (9.52)
A
249 Bell’s theorem

Eq. (9.52) is Bell’s inequality4 .

Bell’s inequality is a consequence of describing measurements through classical


probability theory. Quantum theory violates it. If the system has been prepared in
a pure state |Ψ⟩ ∈ H, ∆ = ⟨Ψ|D̂|Ψ⟩, where
D̂ = m1 · σ̂ ⊗ m2 · σ̂ + m1 · σ̂ ⊗ n2 · σ̂ + n1 · σ̂ ⊗ m2 · σ̂ − n1 · σ̂ ⊗ n2 · σ̂.
(9.53)
For the initial state |Ψ− ⟩ of Eq. (9.28), Eq. (9.39) implies that
∆ = −m1 · m2 − m1 · n2 − n1 · m2 + n1 · n2 . (9.54)
We select the following unit vectors
m1 = (1, 0, 0) m2 = (cos θ, 0, sin θ)
n1 = (0, 0, 1) n2 = (cos θ, 0, − sin θ),
for some angle θ. Eq. (9.54) yields |∆| = 2| sin θ + cos θ|. The maximum value

|∆|max = 2 2 > 2, (9.55)
is achieved for θ = π4 . Hence, some quantum states violate Bell’s inequality.
The problem is that Bell’s inequality follow from elementary, ostensibly obvious,
assumptions. The Bell-inequality assumptions (BIA) are the following.
BIA1. Measurements reveal pre-existing properties of quantum systems. Finding
+1 in a qubit measurement asserts that the qubit has a specific property, say a+ ,
while finding −1 implies that the qubit has a specific property a− that is not
compatible with a+ . This is simply described by the introduction of a state space
−1
Γ for the qubit. The property a+ corresponds to the subset C+ = fm (+1), and a−
−1
corresponds to the subset C− = fm (−1). Obviously C+ and C− are disjoint.
This assumption is called microscopic-property realism, microrealism for short,
or simply realism. The word ”realism” is used as a technical philosophical term,
meaning that something is objectively real and not dependent on human activities,
beliefs or narratives. For example, ”mathematical realism” stands for the statement
that the objects of mathematics exist independently of our mathematical activity,
and that a mathematical proposition is true or false, irrespective of whether we can
prove it or not. In the context of physics, microscopic-property realism means that
a microscopic system has objective properties, i.e., values for physical variables,
that do not depend on how we choose to measure them.
Traditionally, experimental sciences have been treating property realism as a
truism, i.e., they affirm that the measurements reveal properties of the system,
and that the aim of the experimentalist is to come as closer to the ”true value” of
physical variables as possible.
BIA2. We assumed that the probability density ρ on Γ is the same in all four
4 In fact, Eq. (9.52) is not due to Bell, but to Clauser, Horne, Shimony and Holt (Clauser et al.,
1969), and it is often called CHSH inequality. This inequality both simplifies and generalizes
the original inequality by John Bell.
A
250 Composite quantum systems and entanglement

experimental configurations. Again, this is an obvious assumption, given that the


procedure of preparing the qubit system and the procedure of configuring the appa-
ratuses (i.e., choosing the observables to be measured) are completely independent
from each other. BIA2 is often referred to as measurement independence.
Nonetheless, it is conceivable that the apparatuses exert some influence on the
qubits, by a process yet unknown to us, so that the choice of observables in the
apparatus affects the probability density ρ. Of course, we can always place the two
measurement apparatuses very far from each other and from the source. We can
also make the choice about what observable to measure in each apparatus so late,
that no signal traveling slower-than-light can reach qubit 2 from the configuration
of the apparatus for qubit 1.
Hence, BIA2 follows from the principle of locality, that there is no instantaneous
transmission of information. The locality principle is a consequence of relativity,
and, to our present state of knowledge, it holds with no exceptions. Note that
the locality principle is a more stringent condition than BIA2. BIA2 is eminently
reasonable even in a world with instantaneous transmission of information.

The violation of Bell’s inequality implies that either BIA1 or BIA2 is false. If
BIA2 is false then the principle of locality is also false. Hence the violation of Bell’s
inequality contradicts the conjunction of property realism (BIA1) and the locality
principle. This conjunction is known as local realism.

9.4.3 Experimental tests

Since both locality and realism are foundational to our understanding of the phys-
ical world, experiments that test Bell’s inequality are of the highest importance.
The first series of experiments to confirm the violation of Bell’s inequality was un-
dertaken by Alain Aspect in the early 1980s (Aspect et al., 1981, 1982). Aspect
used photonic qubits, in which the vectors m correspond to the direction of the
polarization. One of the experimental configurations is described in Fig. 10.1. Each
photon in a pair passes through an analyser of polarizations in the mi direction
(i = 1, 2), and it is recorded by detector Ai+ if its polarization is positive and by
detector Ai− if its polarization is negative.
For each choice of unit vectors m and n, we repeat the experiment with N qubit
pairs, and we record the following numbers of coincidences.

1. N++ : coincidences in A1+ and A2+ .


2. N+− , coincidences in A1+ and A2− .
3. N−+ , coincidences in A1− and A2+ .
4. N−− , coincidences in A1− and A2− .

The correlation for this experiment


N++ + N−− − N+− − N−−+
Cor(m1 , m2 ) = . (9.56)
N++ + N−− + N+− + N−−+
A
251 Bell’s theorem

Fig. 9.1: Experiment for measuring Bell-type correlations in photonic qubits. S: source
of entangled photon pairs. m1 , m2 : analysers that split a photon beam according with
respect to the polarization in the chosen direction. A1± , A2± : photodetectors. SP : signal
processor for counting coincidences.

In an ideal measurement, N = N++ +N−− +N+− +N−−+ . In an actual experiment,


the coincidences of many qubit pairs are not recorded, and these are not taken into
account in determining the correlation.
Repeating this procedure for four different pairs of polarization directions (m1 , m2 ),
one can evaluate the quantity ∆ of Eq. (9.48). The results of Aspect’s experiments
were unambiguous. Bell’s inequality is violated. This result has been confirmed in
a large number of subsequent experiments—see Box 9.3.

9.4.4 Meaning of the violation of Bell’s inequality

Since experiments confirm the violation of Bell’s inequality unambiguously, local


realism has been proven false. Some loopholes are still logically possible, but they
strain credulity. For example, BIA2 presupposes that the experimentalist is free
to choose which observable to measure for one qubit; hence, if the detectors are
sufficiently far separated this choice cannot affect the other qubit. What if the
experimentalist does not have this freedom of choice? What if his or her choice in
each particular instance has been predetermined long ago, and the qubits have also
been predetermined to behave in such a way as to violate Bell’s inequality. This
type of predetermination can very well be fully local. But then it invokes a cosmic
conspiracy, apparently aiming to make humanity misunderstand physics. We would
have to postulate an entity like Descartes’s malicious demon, i.e., ”a being of utmost
power and cunning that employs all his energies in order to deceive us” (Descartes,
1641/2017).
Evil demons aside, the violation of Bell’s inequality implies that either BIA1 or
BIA2 is false. Which one? BIA2 appears stronger. First, if BIA2 is false, then so is
the locality principle. We have to abandon the highly successful theory of relativity,
without any inkling of what to substitute it with. More importantly, the failure of
BIA2 implies the existence of strange, faster-than-light interactions between appa-
ratuses and microscopic systems. What is strange about such interactions is that
A
252 Composite quantum systems and entanglement

Box 9.3 Bell-inequality experiments


The greatest challenge in Bell-inequality tests is to exclude any reason of doubt about
the result, no matter how implausible. All experiments involve practical assumptions
in the preparation of the system or in the statistics of the outcomes, in addition to the
hypothesis that is being tested. These assumptions are usually thought to be benign and
to not affect the conclusions. However, the stakes of Bell-type experiments are enormous,
and these additional assumptions may work as loopholes. It is therefore necessary to
devise experiments that close such loopholes.
The most important loopholes are the following.
• Locality loophole. We have to account for the possibility that the apparatuses somehow
communicates the choice of measurement to the qubits, so that BIA2 is violated. This
can be avoided in set-ups (i) where the vectors m and n are selected randomly in
each run of the experiment when the qubits are already far from each other (ii) the
duration of the measurement is small, so that any such influence can only occur with
faster-than-light propagation.
• Detection loophole. As explained in Sec. 9.4.3, in Bell-test experiments with photons,
one measures coincidences of photodetection events. Since there are no perfect photode-
tectors, there will be undetected photons. If one photon of a pair is undetected, then
the pair does not count towards the evaluation of Cor(m, n). The obvious assumption
is that detected photon pairs are a representative sample of all emitted photon pairs.
It is highly implausible that this assumption is false, but it is logically possible, so one
has to find ways to carry out the experiment without this assumption.
Hence, to remove all doubts about the failure of local realism, Bell-type experiments
must involve qubits that are well separated before measurement, and detectors that are
both fast and highly efficient. These conditions are difficult to achieve. It took 35 years
from the first experiment that demonstrated the violation of Bell’s inequality to the
first experiment that confirmed the violation without the two loopholes above (Hensen
et al., 2013). Of course, a huge number of experiments on entangled systems had been
giving excellent agreement with the predictions of quantum theory in the intervening
years.
Some important Bell-test experiments are presented in the following table.

Year Location qubits loophole closed Distance Reference

1998 Geneva photonic - 10km Tittel et al. (1998)


1998 Vienna photonic locality 400m Weihs et al. (1998)
2001 Boulder atomic detection 3µm Rowe et al. (2001)
2009 Sta. Barbara solid-state detection 3mm Ansmann et al. (2009)
2013 Vienna photonic detection - Giustina et al. (2013)
2015 Delft electron spin both 1.3km Hensen et al. (2013)
2015 Vienna photonic both 58m Giustina et al. (2015)
2015 Boulder photonic both 180m Shalm et al. (2015)
2017 Munich atomic both 400m Rosenfeld et al. (2017)

they do not appear in the dynamics of the theory (i.e., the Hamiltonian), and they
can never be used in order to send faster-than-light signals (this is not allowed in
quantum theory, see Sec. 9.5.1).
Dropping BIA2 creates more problems than it solves. In contrast, dropping real-
ism is consistent with many other properties of quantum theory, as seen in Chapters
A
253 Bell’s theorem

7 and 9. Still, it is not entirely trouble-free. If microscopic systems have no proper-


ties, and macroscopic systems consist of microscopic ones, then how do macroscopic
systems have properties? This is still an important foundational problem, but most
researchers think it is under control—see Sec. 10.5 for a discussion. Compared to
the chaos that is unleashed by dropping BIA2, the denial of BIA1 appears a saner
choice at the moment.
Nonetheless, there is at least one theory that keeps microscopic realism and drops
BIA2, namely, Bohmian mechanics. This is a reformulation of Schrödinger wave
mechanics by Bohm, which assumes that physical systems are described by both
particles and waves. Each particle is described by its position and momentum,
but also by a physical wave, subject to Schrödinger’s equation that guides the
particle’s motion—see Sec. 10.3. Bohmian mechanics is built so as to reproduce the
predictions of non-relativistic quantum mechanics. However, quantum theory goes
well beyond that, and Bohmian mechanics has not yet been sufficiently generalized.
It is not an alternative theory to quantum mechanics, rather an instructive example
of how a theory that violates BIA2 could be.
The rejection of realism means that the notion of a ”property” has no meaning
beyond the classical macroscopic level that corresponds to measurements. A micro-
scopic system has not properties in itself, properties arise after its interaction with
an apparatus, and they refer to the joint system of microscopic system and appa-
ratus, not to the microscopic system itself. Hence, words like ”property” and ”fact”
cannot be used for microscopic entities the same way they are used for objects of
our everyday experience.

9.4.5 Cirel’son’s inequality

The quantity |∆| of Eq. (9.48) has not only an upper bound from classical prob-
ability, but also one from quantum theory. By definition the maximum value of
|∆| is 4. We showed that systems described by classical probability theory satisfy
|∆| ≤ 2. Cirel’son
√ (1980) proved that systems described by quantum theory sat-
isfy |∆| ≤ 2 2. Hence, experiments that test Bell’s inequality can also check for
violations of quantum theory. So far, none has bee observed.

Proposition 9.2 (Cirel’son’s inequality) |∆| ≤ 2 2.
Proof. The operator D̂ of Eq. (9.53) satisfies

D̂2 = 4Iˆ + [m1 · σ̂, n1 · σ̂] ⊗ [m2 · σ̂, n2 · σ̂].

The proof is straightforward, only recall that (m · σ̂)2 = |m|2 I.


ˆ
By Eq. (5.12),

D̂2 = 4Iˆ − 4(m1 × n1 ) · σ ⊗ (m2 × n2 ) · σ.

Then,

|∆|2 = (⟨ψ|D̂|ψ⟩)2 ≤ ⟨ψ|D̂2 |ψ⟩ = 4 − 4⟨ψ|(m1 × n1 ) · σ ⊗ (m2 × n2 ) · σ|ψ⟩. (9.57)


A
254 Composite quantum systems and entanglement

For any self-adjoint operators  and |B̂, the Cauchy-Schwarz inequality gives

|⟨ψ|Â ⊗ B̂|ψ⟩| ≤ ⟨ψ|Â2 ⊗ I|ψ⟩⟨ψ|
ˆ Iˆ ⊗ B̂ 2 |ψ⟩,

hence,
|⟨ψ|(m1 × n1 ) · σ ⊗ (m2 × n2 ) · σ|ψ⟩| ≤ |m1 × n1 ||m2 × n2 |. (9.58)
Eq. (9.57) becomes

|∆|2 ≤ 4 + 4|m1 × n1 ||m2 × n2 |. (9.59)

The maximum of |m1 × n1 ||m2 × n2 | is 1 (for m1 normal to n1 and m2 normal to n2 ).



Hence, |∆|2 ≤ 8, or |∆| ≤ 2 2.

9.4.6 Bell’s theorem without inequalities

Bell’s inequality involve correlations defined from measurement in a statistical en-


semble of quantum systems. Nonetheless, Bell’s theorem can be proved without
inequalities, and without any reference to a statistical ensemble, ideally solely in
terms of measurements on a individual quantum system. This proof is due to Green-
berger, Horne and Zeilinger, and it is known as the GHZ theorem (Greenberger
et al., 1989, 1990). Here, we present a simpler proof from Mermin (1990).
Consider a three-qubit system, described by the Hilbert space H = C2 ⊗ C2 ⊗ C2 .
We define four operators

 = σ̂1 ⊗ σ̂2 ⊗ σ̂2 , B̂ = σ̂2 ⊗ σ̂1 ⊗ σ̂2 , Ĉ = σ̂2 ⊗ σ̂2 ⊗ σ̂1 , D̂ = σ̂1 ⊗ σ̂1 ⊗ σ̂1 .

These operators commute with each other. To see this,

ÂB̂ = σ̂1 σ̂2 ⊗ σ̂2 σ̂1 ⊗ σ̂22 = σ̂3 ⊗ σ̂3 ⊗ I,


ˆ B̂ Â = σ̂2 σ̂1 ⊗ σ̂1 σ̂2 ⊗ σ̂22 = σ̂3 ⊗ σ̂3 ⊗ I,
ˆ

hence, [Â, B̂] = 0. Similarly, we show that [Â, Ĉ] = [B̂, Ĉ] = 0. Furthermore,

ÂD̂ = σ̂12 ⊗ σ̂2 σ̂1 ⊗ σ̂2 σ̂1 = −Iˆ ⊗ σ̂3 ⊗ σ̂3 , D̂Â = σ̂12 ⊗ σ̂1 σ̂2 ⊗ σ̂1 σ̂2 = −Iˆ ⊗ σ̂3 ⊗ σ̂3 ,

hence, [Â, D̂] = 0. Similarly, we show that [B̂, D̂] = [Ĉ, D̂] = 0.
The key point is that

ÂB̂ Ĉ D̂ = σ̂1 σ̂22 σ̂1 ⊗ σ̂2 σ̂1 σ̂2 σ̂1 ⊗ σ̂22 σ̂12 = −Iˆ (9.60)
(i) (i)
The GHZ theorem is essentially Eq. (9.60). Let us denote by f1 and f2 the
functions that correspond to the measurements of σ̂1 and σ̂2 at the i-th qubit (i =
1, 2, 3), as in Sec. 9.4.1; their values are ±1. Then, the product ÂB̂ Ĉ D̂ corresponds
to the function
 2
(1) (2) (3) (1) (2) (3) (1) (2) (3) (1) (2) (3) (1) (2) (3) (1) (2) (3)
[f1 f2 f2 ][f2 f1 f2 ][f2 f2 f1 ][f1 f1 f1 ] = f1 f1 f1 f2 f2 f2 = 1,

obviously contradicting Eq.(9.60).


Ideally, the predictions of quantum theory can be distinguished from those of a
local realist theory by a single measurement of the observables Â, B̂, Ĉ and D̂. We
B
255 Applications

only need to prepare the system in a common eigenstate of Â, B̂, and Ĉ—by Eq.
(9.60) it will also be an eigenstate of D̂.
Let us denote by a = ±1 the eigenvalues of Â, by b = ±1 the eigenvalues of B̂
and by c = ±1 the eigenvalues of Ĉ. We denote by |a, b, c⟩ the common eigenvectors
of Â, B̂, Ĉ and D̂. There are 23 such eigenvectors, known as the GHZ states, they
define an orthonormal basis on the Hilbert space H = C8 . They are the following

1 1
| − −−⟩ = √ (|1, 1, 1⟩ + |0, 0, 0⟩) | + ++⟩ = √ (|1, 1, 1⟩ − |0, 0, 0⟩)
2 2
1 1
| + +−⟩ = √ (|1, 1, 0⟩ + |0, 0, 1⟩) | − −+⟩ = √ (|1, 1, 0⟩ − |0, 0, 1⟩)
2 2
1 1
| + −+⟩ = √ (|1, 0, 1⟩ + |0, 1, 0⟩) | − −+⟩ = √ (|1, 0, 1⟩ − |0, 1, 0⟩)
2 2
1 1
| − ++⟩ = √ (|0, 1, 1⟩ + |1, 0, 0⟩) | − −+⟩ = √ (|0, 1, 1⟩ − |1, 0, 0⟩) . (9.61)
2 2

Suppose we measure Â, B̂ and Ĉ successively on one of these states, and we find
values a, b, and c, respectively. We can predict that a subsequent measurement of
D̂ will give value −1/(abc), while local realism predicts 1/(abc). Hence, we have
a sharp differentiation of predictions even in a single quantum system. We do not
need to perform measurements on a statistical ensemble.
Experiments, carried out with three entangled photons (Pan et al., 2000), do not
involve four successive measurements on the same system, but Â, B̂, Ĉ and D̂ are
measured on different systems, prepared in the same state. The low efficiency of the
detectors makes it necessary to take a statistical ensemble of such measurements. So
far, experiments of this type lead to results that are fully compatible with quantum
theory.

9.5 Applications

Next, we present four crucial results that elaborate on the quantum rule for the
composition of subsystems and the notion of entanglement.

9.5.1 No-communication theorem

The non-communication theorem asserts that it is impossible to use entangled states


to transmit information instantaneously.
Let us assume two observers, Alice and Bob. Alice can act on quantum systems
described by the Hilbert space HA , and B can act on quantum systems described by
the Hilbert space HB . Suppose that a source produces entangled pairs of systems:
one element of the pair is accessed by Alice and the other by Bob. The state ρ of
B
256 Composite quantum systems and entanglement

these pairs can always be expressed as


X
ρ̂ = Ŝi ⊗ Ti , (9.62)
i

for some operators Ŝi and T̂i on HA and HB , respectively.


Bob picks up his system and moves far away. Then Alice acts on her own system.
The most general action on a state ρ̂ on a Hilbert space H is of the form
X
ρ̂ → K̂a ρ̂K̂a† , (9.63)
a
P
where K̂a are operators on H that satisfy a K̂a† K̂a = I; ˆ they are called Kraus
operators. This result is proven in Sec. 20.1.2. However, the form (9.63) should not
be a surprise: unitary evolution corresponds to a single unitary Kraus operator,
and the rule (8.19) for measurements corresponds to has spectral projectors as
P
Kraus operators. If we combine a measurement of a operator F̂ = a fa P̂a and
time evolution through a unitary operator Û , the Kraus operators for the change
of state are K̂a = Û P̂a .
Suppose that Alice’s action on her system is described by the Kraus operators
K̂a on HA . When describing the entangled pair on the Hilbert space HA ⊗ HB ,
Alice’s Kraus operators are expressed as K̂a ⊗ I. ˆ After Alice’s action, the state of
the composite system is
X XX
ρ̂′ = ˆ K̂a† ⊗ I)
(K̂a ⊗ I)ρ̂( ˆ = (K̂a ⊗ 1)(Ŝi ⊗ T̂i )(K̂a† ⊗ I)
ˆ
a a i
XX
= K̂a Ŝi K̂a† ⊗ T̂i . (9.64)
a i
P
Then, Bob measures the operator Ŷ = yn P̂n on HB. This is represented by
P
Iˆ ⊗ Ŷ = n yn (Iˆ ⊗ P̂n ) on HA ⊗ HB . The associated probabilities are
!
X X
′ ˆ †
Prob(n) = T r[ρ̂ (I ⊗ P̂n )] = T r[K̂a Ŝi K̂a ] · T r(T̂i P̂n )
i a
" ! #
X X Xh i
= T r Ŝi K̂a† K̂a · T r(T̂i P̂n ) = T r(Ŝi ) · T r(T̂i P̂n ) . (9.65)
i a i

We conclude that the probabilities determined by Bob do not depend on the Kraus
operators K̂a , i.e., on the way Alice chose to act on her system. Hence, entanglement
cannot be used for instantaneous transfer of information from Alice to Bob.

Quantum theory does not allow entanglement to be used for instantaneous


transfer of information.

9.5.2 No-cloning theorem

Copying machines are ubiquitous in the modern world. Their inputs are physical
systems in a given state, and their outputs are exact copies of the input, while the
B
257 Applications

input is preserved. For example, a photocopier records all information on a piece


of paper and prints it on a blank page.
In a classical world, there is no limit to the accuracy of copiers. The quantum
world is different (Park, 1970; Wooters and Zurek, 1982; Dieks, 1982). Suppose we
want to copy a system that is found on the state |ψ⟩ of a Hilbert space H. To
this end, we need the quantum equivalent of a blank page of a photocopier. This is
another system, prepared in a specific state |e⟩ of the same Hilbert space H. The
quantum copier is an apparatus that acts with a unitary operator Û on the Hilbert
space H ⊗ H, so that
Û (|ψ⟩ ⊗ |e⟩) = |ψ⟩ ⊗ |ψ⟩, (9.66)
for all |ψ⟩ ∈ H. Eq. (9.66) means that the initial state is preserved, and its content
is fully copied into the ‘blank’.
Consider two vectors |ϕ1 ⟩ and |ϕ2 ⟩ that both satisfy Eq. (9.66): Û (|ϕ1 ⟩ ⊗ |e⟩) =
|ϕ1 ⟩⊗|ϕ1 ⟩ and Û (|ϕ2 ⟩⊗|e⟩) = |ϕ2 ⟩⊗|ϕ2 ⟩. We take the inner product of the left-hand
sides of these equations to obtain,
2
⟨ϕ1 |ϕ2 ⟩ = (⟨ϕ1 |ϕ2 ⟩) . (9.67)
Hence, either ⟨ϕ1 |ϕ2 ⟩ = 0 or ⟨ϕ1 |ϕ2 ⟩ = 1. These conditions do not hold for generic
vectors |ϕ1 ⟩ and |ϕ2 ⟩, but only for ones in an orthonormal set. Hence, there can
be no copying machine (i.e., unitary operator Û ) that works for an arbitrary set
of initial states. This statement is the no cloning theorem. We can only construct
copiers of limited purposes, i.e., apparatuses that can copy a specific orthonormal
set of states.
The non-cloning theorem is a foundational result for quantum cryptography.
Suppose that the state |ψ⟩ contains a message between Alice and Bob and that
Eve attempts to eavesdrop the conservation by coping the quantum state. It is
impossible to do so without affecting the message, i.e., by changing the transmitted
state |ψ⟩, hence, alerting Alice and Bob.

9.5.3 Quantum teleportation

Quantum teleportation is the transfer of the quantum state from one point in space
to another, using the properties of entanglement.
Let Alice and Bob be two observers that share an entangled pair of qubits in the
state |Φ+ ⟩, of Eq. (9.29). Alice’s qubit is described by the Hilbert space HA , and
Bob’s qubit is described by the Hilbert space H. We will keep track of the Hilbert
spaces by writing the state of the shared qubit as |Φ+ ⟩AB .
Alice has another qubit C, described by the Hilbert space HC . C has been pre-
pared in a state |ψ⟩C = α|0⟩C + β|1⟩C , where α, β ∈ C. The state of the system of
three qubits is |Φ+ ⟩AB ⊗ |ψ⟩C . This can be written as
1
|Φ+ ⟩AB ⊗ |ψ⟩C = [|Φ+ ⟩AC ⊗ (α|0⟩B + β|1⟩B ) + |Φ− ⟩AC ⊗ (α|0⟩B − β|1⟩B )
2
+ |Ψ+ ⟩AC ⊗ (β|0⟩B + α|1⟩B ) + |Ψ− ⟩AC ⊗ (β|0⟩B − α|1⟩B )] . (9.68)
B
258 Composite quantum systems and entanglement

To prove Eq. (9.68), start from the right-hand side, 12 out of the 16 terms cancel,
and the remaining ones give the left-hand side.
Next, Alice performs a measurement on her two qubits (A and C) that corre-
sponds to the orthonormal basis defined by the four vectors |Φ± ⟩AC , |Ψ± ⟩AC . Eq.
(9.68) implies that the probability of each of the four outcomes equals 41 . After the
measurement, the three-qubit system is found in one of the following four states.
|Φ+ ⟩AC ⊗ (α|0⟩B + β|1⟩B ), |Φ− ⟩AC ⊗ (α|0⟩B − β|1⟩B ),
|Ψ+ ⟩AC ⊗ (β|0⟩B + α|1⟩B ), |Ψ− ⟩AC ⊗ (β|0⟩B − α|1⟩B ). (9.69)
Then, Alice informs Bob about the outcome of her measurement. Bob follows that
following protocol.

1. If the measurement outcome corresponds to |Φ+ ⟩AC , Bob does nothing.


2. If the measurement outcome corresponds to |Φ− ⟩AC , Bob acts on his qubit with
the operator Û = σ̂3 .
3. If the measurement outcome corresponds to |Ψ+ ⟩AC , Bob acts on his qubit with
the operator Û = σ̂1 .
4. If the measurement outcome corresponds to |Ψ+ ⟩AC , Bob acts on his qubit with
the operator Û = σ̂2 .
In all cases, the state of the three qubit system is separable between the Hilbert
space HA ⊗ HC that describes Alice’s two qubit and the Hilbert space HB that
describes Bob’s qubit. The state of Bob’s qubit is identical with the initial state
of qubit C. Hence, the information contained in C was transferred to Bob with-
out any propagation of matter. For this reason, we call this phenomenon quantum
teleportation.
We note the following.

1. Teleportation is possible for any initial state |ψ⟩C , even if it is unknown.


2. Teleportation requires Alice communicating the outcome of her measurement to
Bob. This communication takes place at slower-than-light speeds. Hence, quan-
tum teleportation does not allow faster-than-light transmission of information.
3. The state of qubit C is fully transferred to B, but C does not remain in its initial
state. The no-cloning theorem is not violated.
4. Teleportation requires an entangled pair of qubits. For this reason, entanglement
constitutes a resource for quantum teleportation.
5. Alice sends to Bob only two bits of information (specifying one out of four out-
comes) and manages to teleport a qubit state. general, the exact information
contained in the qubit state depends on our ability to prepare any state with
desired accuracy. For example, if we can prepare a qubit with accuracy of one
degree with respect to the Bloch sphere angles, we would be able to distinguish
between 180 × 360 = 64, 800 quantum states. The specification of one of those
quantum states requires log2 (64, 800) ≃ 16 bits. Hence, Alice transmitted in-
formation for 16 bits using only 2 bits. Hence, teleportation is a protocol for
super-dense encoding of information.
B
259 Applications

Quantum teleportation was first proposed in Bennett et al. (1993). The first
successful teleportation experiment with photonic qubits was reported in Boschi
et al. (1998); teleportation experiments with atomic qubits were first reported by
Riebe et al. (2004) and Barrett et al. (2004). The current record on teleportation
distance is 1400 km and it was achieved with photonic qubits (Yin et al., 2017).

9.5.4 Entanglement distillation

We saw that quantum teleportation requires the preparation of a pair of qubits in


a maximal entanglement state. How can we have a steady supply of such states, for
use in teleportation or any other application? This is made possible by the process
known as entanglement distillation, the use of many qubit pairs with less-than-
maximum entanglement, in order to construct maximum entanglement pairs.
Here, we present a simple example of entanglement distillation (Bennett et al.,
1996). Consider two pairs of qubits, each prepared in the state |Ψβ ⟩ = cos β2 |0, 0⟩ +
sin β2 |1, 1⟩, where 0 < β < π. The first qubit in each pair belongs to Alice, and the
second one to Bob.
By the analysis of Sec. 9.2.4, the entanglement E of |Ψβ ⟩ is proportional to | sin β|,
and it is maximized for β = π2 , whence |Ψβ ⟩ becomes a maximal entanglement state.
Consider now the joint system of the two qubit pairs, described by the state
|Ψβ ⟩ ⊗ |Ψβ ⟩. We undertake a measurement of the operator
1 
 = σ̂3 ⊗ Iˆ ⊗ Iˆ ⊗ Iˆ + Iˆ ⊗ Iˆ ⊗ σ̂3 ⊗ Iˆ . (9.70)
2
This is a measurement on Alice’s two qubits. We substitute σ̂3 = |1⟩⟨1| − |0⟩⟨0|,
and we straightforwardly obtain the spectral analysis of Â,

 = (1)P̂1 + (−1)P̂−1 (9.71)

where

P̂1 = |1⟩⟨1| ⊗ Iˆ ⊗ |1⟩⟨1| ⊗ Iˆ + |0⟩⟨0| ⊗ Iˆ ⊗ |0⟩⟨0| ⊗ IˆP̂−1


P̂−1 = |0⟩⟨0| ⊗ Iˆ ⊗ |1⟩⟨1| ⊗ Iˆ + |1⟩⟨1| ⊗ Iˆ ⊗ |0⟩⟨0| ⊗ I.ˆ (9.72)

The probability of outcome −1 is

Prob(−1) = ⟨Ψβ , Ψβ |Pˆ0 |Ψβ , Ψβ ⟩ = 2⟨Ψβ ||0⟩⟨0| ⊗ I|Ψ


ˆ β ⟩⟨Ψβ ||1⟩⟨1| ⊗ I|Ψ
ˆ β⟩
β β 1
= 2 cos2 sin2 = sin2 β, (9.73)
2 2 2
while the state of the four-qubit system after a measurement of −1 is

Pˆ−1 |Ψβ , Ψβ ⟩ 1
|Ψ; −1⟩ = p = √ (|0⟩ ⊗ |0⟩ ⊗ |1⟩ ⊗ |1⟩ + |1⟩ ⊗ |1⟩ ⊗ |0⟩ ⊗ |0⟩)(9.74)
.
Prob(−1) 2

In Eq. (9.74), Alice has the first and third qubit, while Bob has the second and
B
260 Composite quantum systems and entanglement

fourth qubit. We rewrite this state, by placing Alice’s and Bob’s qubits in separate
kets, labelled by A and B respectively,
1
|Ψ; −1⟩ = √ (|1, 0⟩A ⊗ |1, 0⟩B + |0, 1⟩A ⊗ |0, 1⟩B ) . (9.75)
2
Alice’s two qubits involve only the two-dimensional subspace VA generated by
|+⟩A := |0, 1⟩A and |−⟩A := |1, 0⟩A . Similarly, Bob’s qubits involves only the two-
dimensional subspace VB generated by |+⟩B := |0, 1⟩B and |−⟩B := |1, 0⟩B . Hence,
Eq. (9.75) is the maximum entanglement state

1
|Ψ; −1⟩ = √ (|+⟩A ⊗ |−⟩B + |−⟩A ⊗ |+⟩B ) (9.76)
2
in the Hilbert space VA ⊗ VB .
With this protocol, two qubit pairs in the state |Ψβ ⟩ are used in order to obtain
a single maximum-entanglement qubit pair with probability 21 sin2 β. On average
we distill one maximum-entanglement qubit pair from sin42 β qubit pairs in the state
|Ψβ ⟩. This protocol can be made more efficient by performing measurements on n
qubit pairs.

9.6 Entanglement in continuous variable systems

Finally, we consider entanglement in continuous-variable systems. A particle on the


line is described by a wave function ψ(x) that is an element of the Hilbert space
L2 (R). N such particles are described by wave functions ψ(x1 , x2 , . . . , xn ) in the
Hilbert space the Hilbert space L2 (R) ⊗ L2 (R) . . . L2 (R) = L2 (Rn ), or by density
matrices ρ(x1 , x2 , . . . , xn ; x′1 , x′2 , . . . , x′n ) := ⟨x1 , x2 , . . . , xn |ρ̂|x′1 , x′2 , . . . , x′n ⟩.
Interestingly, in these systems, entanglement can be quantified in terms of a
variation of the Kennard-Robertson uncertainty relation (Simon, 2000; Duan et al.,
2000). We will show this for N = 2, the generalization to arbitrary N is straight-
forward.
For N = 2, we define position operators x̂1 := x̂ ⊗ I, ˆ x̂2 = Iˆ ⊗ x̂, and momentum
operators p̂1 := p̂ ⊗ I, p̂2 = I ⊗ p̂. The only non-zero commutators between these
ˆ ˆ
operators are [x̂1 , p̂1 ] = [x̂2 , p̂2 ] = iI. ˆ
We define the operators x̂± := 21 (x̂1 ± x̂2 ) and p̂± := p̂1 ± p̂2 . They satisfy the
commutation relations [x̂+ , p̂+ ] = [x̂− , p̂− ] = iIˆ and [x̂+ , p̂− ] = [x̂− , p̂+ ] = 0. The
Kennard-Robertson uncertainty relation implies that
1 1
(∆x+ )(∆p+ ) ≥ , (∆x− )(∆p− ) ≥ . (9.77)
2 2
Next, we apply the PPT criterion to a general density matrix ρ(x1 , x2 ; x′1 , x′2 ).
The PPT transformation exchanges x2 with x′2 . It is convenient to work with the
261 Exercises

Wigner function. The obvious generalization of Eq. (6.52) is


Z
1 y1 y2 y1 y2
W (x1 , x2 , p1 , p2 ) = 2
dy1 dy2 ρ(x1 − , x2 − ; x1 + , x2 + )eip1 y1 +ip2 y2 .
(2π) 2 2 2 2
The PPT transformation takes y2 to −y2 , and by a change of variables y2 → −y2
in the integration, we obtain
P P T : W (x1 , x2 , p1 , p2 ) → W (x1 , x2 , p1 , −p2 ). (9.78)
Hence, the PPT transformation transforms (x1 , x2 , p1 , p2 ) to (x1 , x2 , p1 , p2 ). Equiv-
alently, under PPT (x+ , x− , p+ , p− ) → (x+ , x− , p+ , p− ).
By the PPT criterion, if ρ̂ is separable, then Eqs. (9.77) apply also for the PPT-
transformed state. Hence, for separable states
1 1
(∆x+ )(∆p− ) ≥ (∆x− )(∆p+ ) ≥ . (9.79)
2 2
Therefore, we obtain a simple entanglement criterion: if either of Eqs. (9.79) is vio-
lated, then ρ̂ is entangled. However, the converse does not hold: there exist entan-
gled states that satisfy Eq. (9.79). The uncertainties (∆x+ )(∆p− ) and (∆x− )(∆p+ )
define entanglement witnesses.
Example 9.3. Consider a Gaussian state with Wigner function
W (x1 , x2 , p1 , p2 ) = Ce−a(x1 +x2 −2rx1 x2 )−b(p1 +p2 −2sp1 p2 ) ,
2 2 2 2
(9.80)
√ √
where a, b > 0 and |r|, |s| < 1, and C = ab 1 − r2 1 − s2 /π 2 . We change variables
to x± , p± ,
W (x+ , x− , p+ , p− ) = Ce−2a(1−r)x+ −2a(1+r)x− − 2 b(1−s)p+ − 2 b(1+s)p− ,
2 2 1 2 1 2
(9.81)
We straightforwardly obtain
1 1
(∆x± )2 = , (∆p± )2 = . (9.82)
4a(1 ∓ r) b(1 ∓ s)
The uncertainty relations (∆x+ )(∆p+ ) ≥ 21 and (∆x) − (∆p− ) ≥ 21 imply that
ab(1 − r)(1 − s) ≤ 1 and that ab(1 + r)(1 + s) ≤ 1 for all states.
The conditions (∆x+ )(∆p− ) ≤ 12 and (∆x− )(∆p+ ) ≤ 12 , satisfied by all separable
states imply that ab(1 − r)(1 + s) ≤ 1 and ab(1 + r)(1 − s) ≤ 1. Hence, a state with
either ab(1 − r)(1 + s) > 1 or ab(1 + r)(1 − s) > 1 is entangled. This is possible
only if r and s have opposite signs. Then, (1 + r)(1 + s) and (1 − r)(1 − s) are both
smaller than either (1 − r)(1 + s) (if r < 0, s > 0) or (1 + r)(1 − s) (if r > 0, s < 0).

Questions

9.1 State the difference between H1 ⊗ H2 and H1 ⊕ H2 using logical connectives.


9.2 In what sense do we treat entanglement as a quantity?
262 Composite quantum systems and entanglement

9.3 Is the successive measurement of  ⊗ Iˆ and Iˆ ⊗ B̂ equivalent to the measure-


ment of  ⊗ B̂?
9.4 Explain why all self-adjoint operators of the form  ⊗ Iˆ have degenerate
spectrum. What is their smallest degeneracy?
9.5 You are told that ”probabilities in quantum theory imply ignorance about the
true values of physical quantities, as these are determined by measurement.”
Explain how Bell’s theorem contradicts this claim. Does it provide a definite
disproof?
9.6 Explain why the violation of Bell’s inequality does not imply ”action from a
distance” or faster-than-light transmission of information.
9.7 Identify the errors in the following narrative (Chopra, 2015).
...the commonsense idea of local reality is true only at a certain level. The
whole of reality, as explained by quantum physics, lies deeper. A famous math-
ematical formula, known as Bell’s theorem (after its author, the Irish physicist
John Bell), holds that the reality of the universe must be nonlocal; in other
words, all objects and events in the cosmos are inter-connected with one an-
other and respond to one another’s changes of state.[...]
What kind of explanation would satisfy Bell’s requirement for a totally inter-
connected, nonlocal reality? It would have to be a quantum explanation, because
if [....] a change of spin in one particle causes an equal but opposite change
instantly in its partner somewhere in outer space, it is obvious that the infor-
mation going from one place to the other is traveling faster than the speed of
light. This is not permitted in ordinary reality, either by Newton or Einstein.

9.8 How would you go on explaining Bell’s theorem and its implications to an
audience of non-physicists?
9.9 What would it mean if a Bell-type experiment found ∆ = 3.1 ± 0.1?
9.10 (B) Suppose Alice and Bob are in different planetary systems. Why does Bob
have to wait years for Alice’s message in the teleportation protocol? Why
can’t he just measure his qubit?
9.11 (B) Consider the following comment on quantum teleportation:
”The transmission of observable effects is subject to the same restrictions
as classical communication, and there is no action at a distance—Einstein,
Podolski and Rosen’s bugbear—unless one believes that the state vector has real
physical existence. It is not universally accepted that it does....Nevertheless,
the teleportation device might be taken as a reason to believe in the reality
of the quantum state vector; if something can be transmitted from one object
to another, it might be argued, it must be real. The debate will continue.”
(Sudbery, 1993)
State explicitly the arguments in support of and against the reality of the
quantum state on the basis of the teleportation experiments.
263 Problems

Problems

9.1 A two-qubit system is prepared in the state |ψ⟩ = cos θ|1, 1⟩ + sin θeiϕ |0, 0⟩.
Evaluate the average and the variance of the observable  = 2i (σ̂1 ⊗ σ̂2 − σ̂2 ⊗
σ̂1 )
9.2 The swap operator for a bipartite system is defined by Ŝ(|ϕ⟩⊗|ψ⟩) := |ψ⟩⊗|ϕ⟩.
ˆ P σ̂i ⊗σ̂i ). (b) Find the associated
(a) Prove that for a pair of qubits Ŝ = 12 (I− i
eigenvalues and eigenvectors.
9.3 Find the eigenvalues and the spectral projectors of the operators Kˆ± = σ̂1 ⊗
σ̂2 ± σ̂2 ⊗ σ̂1 .
9.4 A two qubit system is characterized by a Hamiltonian Ĥ = 21 ω1 σ̂3 ⊗ Iˆ +
2 ω2 I ⊗ σ̂3 , and it is prepared in a state |θ⟩ =
1 ˆ √1 (|1, 0⟩ + eiθ |0, 1⟩). At time t1
2
we measure σ̂1 in the first qubit, and at time t2 , we measure σ̂1 in the second
qubit. Evaluate the probabilities for all possible measurement outcomes.
9.5 (i) Show that we can always choose a basis on C2 so that a pure two-qubit
state can be expressed as cos β2 |0, 0⟩ + sin β2 |1, 1⟩, for 0 < β < π. (ii) Show
p
that for this state, the maximum value of |∆| is 2 1 + sin2 β. It follows that
any entangled pure state violates Bell’s inequality.
9.6 Evaluate ∆ of Eq.(9.48) for the Werner state (9.42). Which values of α violate
Bell’s inequality?
9.7 (i) Prove Bell’s original inequality for the two-qubit system,

|Cor(m, n1 ) − Cor(m, n2 )| ≤ 1 + Cor(n1 , n2 ),

for all unit vectors m, n1 and n2 . (ii) Show that the inequality is violated for
the state |Ψ− ⟩.
9.8 The most general pure state for a two-qubit system is |ψ⟩ = a|0, 0⟩ + b|0, 1⟩ +
c|1, 0⟩ + d|1, 1⟩. (i) Show that the E of Eq. (9.20) is a increasing function of
|ad − bc|. This result suggest the use of the concurrence: C(ψ) = 2|ad − bc|
as entanglement measure. (ii) Show that C(ψ) = |⟨ψ|ψ̃⟩|2 , where |ψ̃⟩ = σ̂2 ⊗
σ̂2 |ψ ∗ ⟩ and |ψ ∗ ⟩ is the complex conjugate of |ψ⟩ in the standard basis.
9.9 The concurrence defined in the previous problem generalizes to an entangle-
P
ment measure for mixed states. For any density matrix ρ̂ = i pw |ψi ⟩⟨ψi |,
P
C(ρ̂) := min i pi C(ψi ), where the minimum is over all orthonormal sets that
diagonalize ρ̂. Evaluate the concurrence for the Werner state.
9.10 A state ρ̂ that is diagonal in the Bell basis is called Bell-diagonal. Show that
a Bell-diagonal state is separable only if its largest eigenvalue pmax ≤ 21 .
9.11 Consider the GHZ state |Ψ⟩ = √12 (|0, 0, 0⟩ + |1, 1, 1⟩) of a three-qubit sys-
tem. (i) Show that the reduced density matrix for the first two qubits is an
unentangled mixed state. (ii) Show that after the measurement of σ̂1 on the
third qubit, the reduced density matrix for the first two qubits is maximally
entangled for both outcomes.
264 Composite quantum systems and entanglement

9.12 For any operator  on a Hilbert space H, we can define a vector |Â⟩ =
P
ij ⟨i|Â|j⟩|i, j⟩ ∈ H ⊗ H, where |i⟩ is a basis on H. (i) Show that ⟨Â|B̂⟩ =
T r(† B̂), and, hence that |Â⟩ is well defined only if T r(† Â) < ∞. (ii) Show
that |Â⟩ does not depend on the choice of basis. (iii) Show that for any unitary
Û , |Û ⟩ is a maximum-entanglement vector. (iv) Show that in a normalized
state |Ĉ⟩, ⟨Ĉ|Â ⊗ B̂)|Ĉ⟩ = T r(Ĉ † ÂĈ B̂ T ).
9.13 A three-qubit system is prepared in the W-state, |W ⟩ = √13 (|1, 0, 0⟩+|0, 1, 0⟩+
|0, 0, 1⟩). (i) Show that the reduced density matrix for the first two qubits is
an entangled mixed state. (ii) Calculate the quadratic entanglement of the
first two qubits after a measurement of σ̂1 in the third qubit with outcome
+1.
9.14 In the Hilbert space Cn ⊗ Cn , we define the completely symmetric vector
Pn
|Ω⟩ = √1n i=1 |i⟩ ⊗ |i⟩, and the isotropic states

1−λˆ
ρ̂λ = I + λ|Ω⟩⟨Ω|. (9.83)
n2
(i) Show that the reduced density matrix is ρ̂1 = n1 I. ˆ (ii) Show that ρ̂λ is
1
entangled for λ > n+1 .
9.15 Consider the Hilbert space Cn ⊗ Cn , with the totally symmetric vector |Ω⟩
as in the previous exercise. We define Ŵ = Iˆ − n|Ω⟩⟨Ω|. (i) Show that all
separable density matrices ρ̂ satisfy T r(ρ̂Ŵ ) ≥ 0. This implies, that any state
ρ̂ that satisfies T r(ρ̂Ŵ ) < 0 is entangled. (ii) Find the values of λ, for which
the states (9.83) satisfy T r(ρ̂λ Ŵ ) < 0. (iii) Take n = 2. Can this criterion
identify |Ψ− ⟩ as an entangled state?
9.16 The Jaynes-Cummings model (Jaynes and Cummings, 1963) describes the
interaction of an atom (modeled by a two-level system of frequency Ω) with a
mode of the quantum electromagnetic field of frequency ω. The Hilbert space
is C2 ⊗ L2 (R), and the Hamiltonian is
1 g
Ĥ = Ωσ̂3 ⊗ Iˆ + ω Iˆ ⊗ ↠â + (σ̂− ⊗ ↠+ σ̂+ ⊗ â)
2 2
(i) Express the Hamiltonian as a sum Ĥ1 +Ĥ2 , where Ĥ1 = ω( 21 σ̂3 ⊗I+ ˆ † â),
ˆ I⊗â
and show that [Ĥ1 , Ĥ] = 0. (ii) Show that Ĥ1 has eigenvalues n − 21 , for
n = 0, 1, 2, . . ., for n = 0 there is no degeneracy, for n > 0 the eigenvalues
have double degeneracy, and that the corresponding eigenspace Vn spanned
by |0⟩ ⊗ |n⟩ and |1⟩ ⊗ |n − 1⟩. (iii) Since Ĥ commutes with Ĥ1 , it suffices to
diagonalize Ĥ in each Vn . Find the associated eigenvalues and eigenvectors,
and evaluate e−iĤt on each subspace. (iv) Evaluate the probability that the
atom is found in an excited state at time t, for an initial state |0⟩ ⊗ |n0 ⟩. What
is the associated Rabi-frequency? (v) Evaluate the reduced density matrix for
the first qubit as a function of time, and the associated value of the quadratic
entanglement.
9.17 (B) An improved quantum copier uses two copies of the ”blank” state |e⟩, in
order to achieve better results. Let a general input state |ψ⟩ be copied through
265 Problems

the unitary transformation |ψ⟩ ⊗ |e⟩ ⊗ |e⟩ → |ψ⟩ ⊗ |ψ⟩ ⊗ |f ⟩, where |f ⟩ is a


trash-state that occurs at the end of the procedure and depends on |ψ⟩. Show
that the no-cloning theorem still holds.
9.18 (B) Consider two qubit pairs prepared in the state |Ψ⟩ = cos θ|0, 0⟩+sin θ|1, 1⟩.
Alice and Bob share the first pair (qubits 1 and 2), while Bob owns the second
pair (qubits 3 and 4). Bob performs a measurement on qubits 2 and 3 that
corresponds to the Bell basis. (a) Show that the resulting state for the qubits
1 and 4 is a maximally entangled state with probability sin2 θ cos2 θ. (b) Show
that the other produced state for qubits 1 and 4 has less entanglement than
the initial one.
9.19 (B) In a N -particle system, define ξˆa , a = 1, . . . , 2N by (ξˆ1 , ξˆ2 , . . . , ξˆ2N −1 , ξˆ2N ) :=
(x̂1 , p̂1 , . . . , x̂N , p̂N ). Then the commutation relations become [ξa , ξb ] = iΩab Iˆ
for some antisymmetric matrix Ω. (i) Write the matrix Ω explicitly. (ii) Show
that the correlation matrix Cab := Cor(ξˆa , ξˆb ) satisfies C + 2i Ω ≥ 0. (iii) Show
that for N = 1, this inequality reduces to the the uncertainty relation (6.44),
and, hence, to the Kennard-Robertson uncertainty relation.
9.20 (B) (i) Show that the correlation matrix C (defined in the previous problem)
for a two-particle factorized state satisfies C + 2i Ω̃ ≥ 0, where Ω̃ = ΛΩΛ,
Λ = diag{1, 1, 1, −1}. (ii) Use this matrix inequality in order to rederive Eq.
(9.79).

Bibliography

• For quantum entanglement, see Chap. 6 of Jaeger (2007) and for quantum com-
munication, including teleportation, see Chap. 8 of Jaeger (2007).
• For Bell’s theorem, its background, related experiments and their implications,
see Chap. 6 of Peres (2002), Wayne et al. (2020) and Goldstein et al. (2011). For
the various ”loopholes” in Bell-type experiments, see Larsson (2014). I highly rec-
ommend the semi-popular articles of Mermin (1981, 1985).

You might also like