The Mathematics of Entanglement: Summer School at Universidad de Los Andes
The Mathematics of Entanglement: Summer School at Universidad de Los Andes
Contents
Lecture 1 - Quantum States
1.1 Probability theory and Tensor products
1.2 Quantum Mechanics . . . . . . . . . . .
1.2.1 Measurements . . . . . . . . . .
1.3 Mixed states . . . . . . . . . . . . . . .
1.4 Composite systems and Entanglement .
1.4.1 Partial trace . . . . . . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
3
3
4
4
5
5
6
11
11
11
13
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
14
14
15
16
16
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
17
17
17
18
18
19
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
Bell inequalities
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
23
23
23
24
25
26
.
.
.
.
.
27
27
28
28
28
28
39
Estimation
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
46
46
47
47
48
48
27 May, 2013
Quantum States
Lecturer: Fernando G.S.L. Brand
ao
Lecture 1
Entanglement is a quantum mechanical form of correlation, which appears in many areas, such
as condensed matter physics, quantum chemistry, and other areas of physics. This week we will
discuss a perspective from quantum information, which means we will abstract away the underlying
physics, and make statements about entanglement that apply independent of the underlying physical system. This will also allow us to discuss information-processing applications, such as quantum
cryptography.
1.1
Before discussing quantum states, we explain some aspects of probability theory, which turns out
to have many similar features.
Suppose we have a system with d possible states, for some integer d, which we label by 1, . . . , d.
Thus a deterministic state is simply an element of the set {1, . . . , d}. The probabilistic states are
probability distributions over this set, i.e. vectors in Rd+ whose entries sum to 1. The notation
Rd+ means that the entries are nonnegative. Thus, a probability distribution p = (p(1), . . . , p(d))
P
satisfies dx=1 p(x) = 1 and p(x) 0 for each x. Note that we can think of a deterministic state
x {1, . . . , d} as the probability distribution where p(x) = 1 and all other probabilities are zero.
n
Composition. Suppose we are given p Rm
+ and q R+ which correspond to independent
probability distributions. Their joint distribution is given by the vector
p(1)q(1)
p(1)q(2)
..
p q :=
p(1)q(n) .
..
.
p(m)q(n)
We have introduced the notation to denote the tensor product, which in general maps a pair of
vectors with dimensions m, n to a single vector with dimension mn. Later we will also consider
the tensor product of matrices. If Mn denotes n n matrices, and we have A Mn , B Mm then
A B Mnm is the matrix whose entries
are all
possible products of an entry of A and an entry
a11 a12
of B. For example, if n = 2 and A =
then A B is the block matrix
a21 a22
a11 B a12 B
.
a21 B a22 B
One useful fact about tensor products, which simplifies many calculations, is that
(A B)(C D) = AC BD.
We also define the tensor product of two vector space V W to be the span of all v w for
v V and w W . In particular, observe that Cm Cn = Cmn .
3
1.2
Quantum Mechanics
We will use Dirac notation in which a ket |i denote a column vector in a complex vector space,
i.e.
1
2
|i = . Cd .
..
d
The bra h| denotes the conjugate transpose, i.e.
h| = 1 2
d .
d
X
i i .
i=1
v
u d
uX
p
|i |2 .
kk2 = h|i = t
i=1
Now we can define a quantum state. The quantum analogue of a system with d states is the
d-dimensional Hilbert space Cd . For example, a quantum system with d = 2 is called a qubit. Unit
vectors |i Cd , where h|i = 1, are called pure states. They are the analogue of deterministic
states in classical probability theory. For example, we might define the following pure states of a
qubit:
1
1
0
1
, |+i = (|0i + |1i), |i = (|0i + |1i).
, |1i =
|0i =
1
0
2
2
Note that both pairs |0i, |1i and |+i, |i form orthonormal bases of a qubit.
1.2.1
Measurements
A projective measurement
P is a collection of projectors {Pk } such that Pk Md for each k, Pk = Pk ,
Pk Pk0 = k,k0 Pk and k Pk = I. For example, we might measure in the computational basis, which
consists of the unit vectors |ki with a one in the k th position and zeros elsewhere. Thus define
0
0
..
1
Pk = |kihk| =
.
.
.
0
0
Borns rule states that Pr[k], the probability of measurement outcome k, is given by
Pr[k] = h|Pk |i.
As an exercise, verify that this is equal to tr(Pk |ih|). In our example, this is simply |k |2 .
Example. If we perform the measurement {|0ih0|, |1ih1|} on |+i, then Pr[0] = Pr[1] = 1/2. If
we perform the measurement {|+ih+|, |ih|}, then Pr[+] = 1 and Pr[] = 0.
1.3
Mixed states
Mixed states are a common generalization of probability theory and pure quantum mechanics. In
general, if we have an ensemble of pure quantum states |x i with probabilities p(x), then define
the density matrix to be
X
=
p(x)|x ihx |.
x
(1.1)
1.4
We will work always with distinguishable particles. A pure state of two quantum systems is given
by a unit vector |i in the tensor product Hilbert space Cn Cm
= Cnm .
For example, if particle A is in the pure state |iA and particle B is in the pure state |iB
then their joint state is |iAB = |iA |iB . If |iA Cn and |iB Cm , then we will have
|iAB Cmn .
5
This should have the property that if we measure one system, say A, then we should obtain
the same result in this new formalism that we would have had if we treated the states separately.
If we perform the projective measurement {Pk } on system A then this is equivalent to performing
the measurement {Pk I} on the joint system. We can then calculate
Pr[k] = h|Pk I|i
=A h|B |i(Pk I)|iA |iB
= h|Pk |ih|i
= h|Pk |i
Just as there are joint distributions for which two random variables are not independent (e.g.,
...), there are quantum states which are cannot be written as a tensor product |i |i for any
choice of |i, |i. In quantum mechanics, this situation can even occur for pure states of the system.
For example, consider the EPR pair (we will see the reason for the name later):
|+ i =
.
2
We say that a pure state |i is entangled if, for any |i, |i, we have |i =
6 |i |i.
Entangled states have many counterintuitive properties. For example, suppose we measure the
state |+ i using the projectors {Pj,k = |jihj| |kihk|}. Then we can calculate
Pr[(0, 0)] = h+ |P0,0 |+ i = h+ | |0ih0| |0ih0| |+ i =
1
2
1
2
Pr[(0, 1)] = 0
Pr[(1, 1)] =
Pr[(1, 0)] = 0
The outcomes are perfectly correlated.
However, observe that if we measure in a different basis, we will also get perfect correlation.
Consider the measurement with outcomes
{| + +ih+ + |, | + ih+ |, | +ih + |, | ih |,
where we have used the shorthand | + +i := |+i |+i, and similarly for the other three. Then one
can calculate (and doing so is a good exercise) that, given the state |+ i, we have
1
Pr[(+, +)] = Pr[(, )] = ,
2
meaning again there is perfect correlation.
1.4.1
Partial trace
Suppose we have AB D(Cn Cm ). We would like a quantum analogue of the notion of a marginal
distribution in probability theory.
where {|ki} is any orthonormal basis on B. The operation trB is called a partial trace.
We observe that if we perform a measurement {Pj } on A, then we have
Pr[j] = tr(Pj I)AB = tr(Pk trB (AB )) = tr(Pk A ).
Thus the reduced state A perfectly reproduces the statistics of any measurement on the system A.
27 May, 2013
Quantum Operations
Lecturer: Matthias Christandl
Lecture 2
In this lecture we will talk about dynamics in quantum mechanics. First stating from measurements again, and then going to unitary evolutions and general quantum dynamical processes.
2.1
Measurements, POVMs
The
P converse of the above is also true: Whenever we are given a set of PSD matrices Qi 0
with i Qi = I, we can always find a projective measurement {Pi } on a larger system A B such
that
tr(Qi A ) = tr(Pi (A |0ih0|B )).
8
The generalized quantum measurements we obtain in this way are called positive operator-valued
measure(ment)s (POVMs). Note that since the Qi s are not necessarily orthogonal projections,
there is no upper bound on the number of elements in a POVM.
Example 1. Consider two projective measurements, e.g. {|0ih0|, |1ih1|} and {|+ih+|, |ih|}.
Then we can define a POVM as a mixture of these two:
1
Q0 = |0ih0|,
2
1
Q1 = |1ih1|,
2
1
Q2 = |+ih+|,
2
1
Q3 = |ih|.
2
P
It is clear that k Qk = I. One way of thinking about this POVM is that with probability 1/2 we
measure in the computational basis, and with probability 1/2 in the |i basis.
Example 2. The quantum state of a qubit can always be written in the form
= (~r) =
1
(I + rx x + ry y + rz z ) ,
2
0 i
1 0
0 1
.
, y =
, z =
i 0
0 1
1 0
Note that the Pauli matrices are traceless, so that the state has indeed trace one (its normalized).
We can the describe the state by a 3-dimensional vector ~r = (rx , ry , rz ) R3 . It turns out that
is PSD if, and only if, krk2 1. Therefore every quantum state of a qubit corresponds to a point
in a 3-dimensional sphere, called the Bloch sphere. A state is pure if, and only if, krk2 = 1.
Let us consider a collection of four pure statesP
{|ai ihai |}4i=1 that form a tetrahedron on the Bloch
sphere. Then, by symmetry of the tetrahedron, i |ai ihai | = I, so they form indeed a POVM.
FIXME: Insert tetrahedron figure.
2.2
Unitary Dynamics
Let |i be a quantum state and consider its time evolution according to the Schrodinger equation
for a time-independent Hamiltonian H. Then the state after some time t is given by
|t i = eiHt |i,
where we have set ~ = 1. The matrix U = eiHt describing the evolution of the system is a unitary
matrix, i.e. U U = U U = I.
Example 1. Ut = eit~e~ /2 with ~e R3 a unit vector and ~ = (x , y , z ) the vector of Pauli
matrices. We have
Ut (~r)Ut = (Rt~r),
where Rt rotates by an angle t around the axis ~e.
9
1
2
1 1
. Its action on the computa1 1
2.3
There are more general possible dynamics in quantum mechanics than unitary evolution. One
possibility is that we add an acilla state |0ih0|B to A and consider a unitary dynamics UAB :
A B A0 B 0 on the joint state. The state of the A0 B 0 system is
U A |0ih0|B U
(2.1)
Suppose now we are only interested in the final state of the subsystem A0 . Then
A0 = trB 0 U A |0ih0|B U ,
(2.2)
where we traced out over subsystem B 0 . We can associate a map to this evolution as
(A ) = 0A = trB 0 U A |0ih0|B U .
(2.3)
What are the properties of ? First it maps PSD matrices to PSD matrices.We call this property
positivity. Second, it preserves the trace. We say the map is trace preserving.
Another more interesting property is that even the map Id, where Id is the identity map on
an auxiliary space of arbitrary dimension, is positive. We call this property completely positivity.
An important theorem (sometimes called Stinespring dilation) is that the converse also holds:
Any which is compeltely positive and trace preserving can be written as
(X = trB 0 (U X |0ih0|B U ) ,
(2.4)
10
27 May, 2013
Quantum Entropy
Lecturer: Aram Harrow
3.1
Lecture 3
Shannon Entropy
In this part, we want to understand quantum information in a quantitative way. One of the
important concepts is entropy. But let us first look
P at classical entropy.
d
Given is a probability distribution p R+ , i pi = 1. The Shannon entropy of p is
H(p) =
pi log pi
(log is always to base to as we are talking about bits and units, convention limx70 x log x = 0).
Entropy quantifies uncertainty. We have maximal certainty for a deterministic distribution, e.g.
p = (1, 0, . . . , 0), H(p) = 0. The distribution with maximal uncertainty is p = ( d1 , . . . , d1 ), H(p) =
log d.
In the following we want to give Shannon entropy an operational meaning with help of the
problem of data compression. For this imagine you have a binary alphabet (d = 2) and you sample
n times independently and identically distribution from the distribution p = (, 1 ); we write
X1 , . . . , Xn i.i.d. {0, 1}. (Prob[Xi = 0] = Prob[Xi = 1] = 1 )
in the third. This implies that the standard deviation of S is smaller than 2n .
What does this have to do with compression? The
of possible n-bit
strings is
total number
(n/e)n
n
n!
n
n
|{0, 1} | = 2 . The number of strings with n 0s is n = (n)!((1)n)! = (n/e)n ((1)n/e)(1)n =
(1/)n (1/(1))n(1) where we used Stirlings approximation. We can rewrite this as exp(n log 1/+
(1 ) log 1/(1 )) = exp(nH(p)). Hence, we only need to store around exp(nH(p)) possible
strings, which we can do in a memory having nH(p) bits. (Note we ignored the fluctuations. If we
took them into account, we would only need additional O( n) bits.) This analysis easily generalises
to arbitrary alphabets (not only binary).
3.2
Typical
I now want to give you a different way of looking at this problem, a way that is both more rigorous
and will more easily generalise to the quantum case. This we will do with help of typical sets.
Again let X1 , . . . , Xn be i.i.d distributed with distribution p. The probability of a string is then
given by
11
Prob[X1 . . . Xn = x1 , . . . , xn ] = p(x1 )p(x2 ) p(xn ) where we used the notation (random variables are in capital letters and values in small letters)
p (n) = p p p (n times).
xn = (x1 , . . . , xn ) n
where = {1, . . . , d} is the alphabet and n denotes strings of length n over that alphabet
Note that
log pn (xn ) =
n
X
p
n V ar[log p(xi ] = nH(p) O( n)
i=1
where we used
E[log p(xi )] =
Let us now define the typical set as the set of strings whose
Tp,n, = {xn | log pn (xn ) + nH(p)| n}
Then
> 0
lim pn Tp,n, = 1
Our compression algorithm simply keeps all the strings in the typical set and throws away all
others. Hence, all we need to know the size of the typical set. This is easy. Note that
xn Tp,n, = exp(nH(p) n) pn (xn ) exp(nH(p) + n)
Note
1 pn (Tp,n, ) |Tp,n, | min pn (xn )
where the minimum is over all strings in the typical set. This implies
1 |Tp,n, | exp(nH(p) n)
which is equivalent to
log |Tp,n, | nH(p) + n
Exercise: Show this is optimal. More precisely, show that we cannot compress to nR bits for
R < H(p) unless the error does not go to zero. Hint: Use Chebycheff ienquality: Let Z be a random
variable P rob[|Z E[Z]| kSD[Z]] 1/k 2 Possible simplifications: 1) pretend all strings to be
typical 2) use exactly nR bits.
12
3.3
Quantum compression
13
28 May, 2013
Lecture 4
(4.1)
Observe that this has the property that repeated measurements always produce the same answer
(although the same is not necessarily true of generalized measurements).
For a pure state |i, the post-measurement state is
Pk |i
.
kPk |ik
(4.2)
Equivalently, we can write Pk |i = p|i, where |i is a unit vector representing the postmeasurement state, and p is the probability of that outcome.
4.1
Teleportation
Suppose that Alice has a qubit |iA0 = c0 |0i + c1 |1i that she would like to transmit to Bob. If they
have access to a quantum channel, such as an optical fiber, she can of course simply give Bob the
physical system A0 whose state is |i. This approach is referred to as quantum communication.
However, if they have access to shared entanglement, then this communication can be replaced
with classical communication (while using up the entanglement). This is called teleportation.
The procedure is as follows. Suppose Alice and Bob share the state
|+ iAB =
|0, 0i + |1, 1i
,
2
and Alice wants to transmit |iA0 to Bob. Then Alice first measures systems AA0 in the basis
{|+ i, | i, | + i, | i}, defined as
|0, 0i |1, 1i
2
|0,
1i
|1, 0i
| i =
2
| i =
meaning the outcome occurs with probability 1/4 and when it does, Bob gets |i.
One can show (calculation omitted) that outcome i (for i {0, 1, 2, 3}) corresponds to
1
(|i ihi | IB ) |iA0 |+ iAB = |i iA0 A i |iB ,
2
where {0 , 1 , 2 , 3 } denote the four Pauli matrices {I, x , y , z }. The 1/2 means that each
outcome occurs with probability 1/4. Thus, transmitting the outcome i to Bob allows him to apply
the correction i and recover the state |i.
This protocol has achieved the following transformation of resources:
1bit entanglement + 2 bits classical communication 1 qubit quantum communication.
As a sanity check, we should verify that entanglement alone cannot be used to communicate.
To check this, the joint state after the measurement is
3
A0 AB =
1X
|i ihi |A0 A i |ih|i .
4
i=0
B =
I
1X
i |ih|i = .
4
2
i=0
Teleporting entanglement. This protocol also works if applied to qubits that are entangled
with other states. For example, Alice might locally prepare an entangled state |iRA0 and then
teleport qubit A0 to Bob. Then the state |i will be shared between Alices system R and Bobs
system B. Thus, teleportation can be used to create shared entanglement. Of course, it consumes
entanglement at the same rate, so we are not getting anything for free here.
4.2
Suppose that Alice and Bob can freely communicate classically and can manipulate quantum systems under their control, but are limited in their ability to communicate quantumly. This class
of operations is called LOCC, meaning local operations and classical communication. It often
makes sense to study entanglement in this setting, since LOCC can modify entanglement from one
type to another, but cannot create it where it didnt exist before. What types of entanglement
manipulations are possible with LOCC?
One example is to map a pure state |iAB to (UA VB )|iAB , for some choice of unitaries
UA , VB .
A more complicated example is that Alice might measure her state with a projective measurement {Pk } and transmit the oucome to Bob, who performs a unitary Uk depending on the outcome.
This is essentially the structure of teleportation. The resulting map is
X
AB 7
(Pk Uk )(Pk Uk ).
k
One task for which we might like to use LOCC is to extract pure entangled states from a
noisy state. For example, we might want to map AB to |+ ih+ |. This problem is in general
15
called entanglement distillation since we are distilling pure entanglement out of noisy entanglement.
However, we typically consider it with a few variations. First, as with many information-theoretic
m , and
problems, we will consider asymptotic transformations in which we map n
AB to |+ ih+ |
seek to maximize the ratio m/n as n . Additionally, we will allow a small error (to be
formalized later) that goes to zero as n . Semi-formally, the distillable entanglement of is
nm
o
LOCC
ED (AB ) = lim max
: n m |+ ih+ |m .
n
n
4.3
The maximum distinguishing bias that any measurement can achieve between a pair of states ,
is
D(, ) = max | tr(M ( ))|.
{M,IM }
0M I
It turns out that D(, ) = 21 kk1 , where kXk1 is the trace norm, defined as kXk1 = tr( X X).
For this reason, the distance D(, ) is also called the trace distance.
Using this language, we can define ED properly as
nm
o
LOCC
ED (AB ) = lim lim max
: n m , km |+ ih+ |m k1 .
0 n
n
4.4
Entanglement dilution
Suppose we wish to create a general entangled state AB out of pure EPR pairs. As with distillation,
we will aim to maximize the asymptotic ratio achievable while the error goes to zero. Define the
entanglement cost
nm
o
Ec (AB ) = lim lim min
k
: LOCC, k(|+ ih+ |m ) n
AB 1
0 n
n
In general, Ec and ED are both hard to compute. However, if AB is pure then there is a simple
beautiful formula.
Theorem 4.1. For any pure state |iAB ,
Ec (|ih|AB ) = ED (|ih|AB ) = S(A ) = S(B ),
where S() = tr log .
16
28 May, 2013
Lecture 5
In the picture is the cover of a a book by Godel, Escher and Bach. You see it projects B, G or
E depending from where you project the light on. Is it possible to project any triple of letters in
this way? It turns out that the answer is no. For example, by geometric considerations there is no
way of projecting A everywhere.
The goal of this lecture is to introduce a quantum version of this problem!
5.1
Consider a set of n particles each with d-dimensions. The state lives in (Cd )n . We consider
different subsets of the particles Si {1, . . . , N } and suppose we are given quantum states in each
of this sets: Si . The question we want to address is whether they are compatible, i.e. does there
exist a quantum state {1,...,n} such that for all Si ,
trSic () = Si ,
(5.1)
5.1.1
Physical Motivation
This is a interesting problem from a mathematical point of view, but it is also a prominent problem
in the context of condensed P
matter physics and quantum chemistry. Consider a nearest-neighbours
Hamiltonian on a line H = i hi,i+1 , where hi,i+1 := hi,i+1 I{1,...,n} i,i+1 only acts on qubits i and
i + 1. A quantity of interest is the groundenergy of the model, given by the minimum eigenvalue
of H. We can write it variationally as
Eg = minh|H|i =
|i
min
{1,...,n}
tr({1,...,n} )
(5.2)
since the set of quantum states is convex and the extremal points are the pure states. Continuing,
Eg =
min tr({1,...,n} )
X
= min
tr hi,i+1 I{1,...,n}
{1,...,n}
= min
i,i+1
tr(hi,i+1 i,i+1 ).
(5.3)
Therefore
Eg =
min
{i,i+1 }compatible
X
i
17
tr(hi,i+1 i,i+1 ),
(5.4)
(5.5)
Clearly this inequality puts restrictions on compatible states. More interestingly, one can also use
results from the quantum marginal problem to give a proof of the inequality.
5.2
A particular case of the marginal problem is the following: given three quantum states A , B and
C , are they compatible? In this case it is that the answer is yes, just consider ABC = A B C .
But what if we require that the global state ABC is pure? I.e. we would like to have a pure
state |iABC such that
trAB (|ih|ABC ) = C , trAC (|ih|ABC ) = B , trBC (|ih|ABC ) = A .
(5.6)
Then just taking the tensor product of the reduced states is not an option any more.
Example 1: A = B = C = I/2 are compatible, with the GHZ state (|0, 0, 0i + |1, 1, 1i)/ 2
being a possible
extension. A = B = C = I/2 are compatible, with the GHZ state (|0, 0, 0i +
5.2.1
Warm-Up: 2 Parties
(5.7)
for orthogonal basis {|ei i} and {|fi i} o A and B, respectively. The numbers {si } are called Schmidt
values of |iAB . The reductions of |iAB are
X
A =
s2i |ei ihei |
(5.8)
i
18
and
A =
(5.9)
Therefore we see that the eigenvalues of A and B are equal and given by {s2i }.
Going back to the compatibility question, we then see from the discussion above that A and
B are compatible if, and only if, they have the same spectrum.
5.2.2
3 Parties of Qubits
Consider A , B and C each acting on C2 . Then since A = (A,1 , 1 A,1 ), the compatible region
is a subset of R3 . This have a simple algebraic characterization (shown in []):
B
C
A
max + max 1 + max
19
(5.10)
28 May, 2013
Monogamy of Entanglement
Lecturer: Aram Harrow
Lecture 6
Hij
<i,j>
1
n
X
1i<jn
6.1
Symmetric Subspace
Let Sn be the group of permutations of n objects. Note that it contains n! elements. Now fix D
and a permutation Sn . Let P act on (C D )n :
P |i1 i |in i = |i1 (1) i |i1 (n) i.
20
The symmetric subspace is defined as the set of vectors that are invariant under the action of the
symmetric group
Symn (C D ) = {|i (C D )n : P |i = |i Sn }.
Example:D = 2, n = 2
Sym2 (C 2 ) = span{|00i, |11i, |01 + 10i}
D = 2, n = 2
Sym2 (C 2 ) = span{|000i, |111i, |001 + 010 + 100i, |101 + 011 + 110i}
n
The general
P construction is as follows. For this define the type of a string x = (x1 , . . . , xn ) as
n
type(x ) = i exi , where ej is the basis vector with a 1 in the jth position. t = (t1 , . . . , td ) is a
type if t1 + t2 + td = n and the ti are natural numbers. For every type t the vector
1/2 X
n
|t i =
|xn i
t
where nt is the multinomial coefficient. Symn (C D ) = span{|t i} We can now compute the dimension of the symmetric subspace. Note that we can interpret this number
as the number of ways
in which you can arrange n balls into D buckets. There are n+D1
ways
of doing this, which is
n
therefore the dimension.
A useful way for calculations involving the symmetric subspace are the following two characterisations of the P
projector onto the symmetric subspace
1
1) sym = n!
P where the sum is over all permutations
R
sym
2) tr sym = d|ih|n where we integrate over the unit vectors in C D with the uniform
R
n+D1
n
D
measure d normalised
n
R to d = 1. Note that tr sym = dim Sym (C ) =
Example:
R n = 1: d|ih| = I/D
n = 2: d|ih|2 = Psym /(D(D + 1)/2) = (I + SW AP )/(D(D + 1))
We can prove 2) either by representation theory (using Schurs lemma) or by rewriting the
integral over unit vectors as an integral over Gaussian vectors and then using Wicks theorem to
solve the integral.
Note that sym projector onto the symmetric subspace, hence
a) |i Symn (C D ), sym |i = |i b) |i (C D )n , sym |i Symn (C D ) c) sym = sym
6.2
Application to Estimation
E(|h|i|
)
R
= sym . In fact, we will
In order to do this, we will use the continuous POVM {Q }, dQ
|
n . The normalisation constant can be worked out as follows
choose Q = c|ih
Z
n+D1
n
c|ih| = csym /
,
n
21
|
n .
|ih
hence Q = n+D1
n
We now use this to solve our estimation problem:
Z
2k
2k
E(|h|i| = |h|i|
Pr[|]
|
n |ih|n )
where P rob[|]
= tr(|ih
This equals
Z
where we used
tr |ih
n+k
sym
n+k+D1
n+k
n+k
n+k
|ih|
n+D1
=
n
d|ih|n+k .
22
n+D1
n
n+kD1
n
1 Dk/n
29 May, 2013
Lecture 7
7.1
Mixed-state entanglement
For pure states, a state is entangled if its not a product state. This is easy to check, and we can
even quantify the amount of entanglement (using Theorem 7.1) by looking at the entropy of one of
the reduced density matrices.
But what about for mixed states? Here the situation is more complicated.
Definition 1. Define the set of separable states Sep to be the set of all AB that can be written as
X
pi |i ihi |A |i ihi |B .
(7.1)
i
7.2
It is in general hard to test, given a state AB , whether Sep. Naively we would have to check
for all possible decompositions of the form in (7.1). So it is desirable to find efficient tests that
work at least some of the time.
One such test is the Positive Partial Transpose, or PPT, test. The partial transpose can be
thought of as (T id), where T is the transpose map. More concretely, if
X
XAB =
ci,j,k,l |iihj|A |kihl|B
i,j,k,l
23
i,j,k,l
i,j,k,l
This is still a valid density matrix and in particular is positive semidefinite (indeed, it is also in
Sep).
Thus, Sep implies PPT. The contrapositive is that 6 PPT implies 6 Sep. This
gives us an efficient test that will detect entanglement in some cases.
Are there in fact any states that are not in PPT? Otherwise this would not be a very interesting
test.
Examples
1. |+ iAB =
|0,0i+|1,1i
.
2
Then
1
(|0, 0ih0, 0| + |1, 0ih0, 1| + |0, 1ih1, 0| + |1, 1ih1, 1|)
2
1/2 0
0
0
0
0 1/2 0
= 1 SWAP.
=
0 1/2 0
0 2
0
0
0 1/2
A
|+ ih+ |TAB
=
This has eigenvalues (1/2, 1/2, 1/2, 1/2), meaning that |+ ih+ |AB 6 PPT. Of course, we
already knew that |+ i was entangled.
2. Lets try an example where we dont already know the answer, like a noisy version of |+ i.
Let
I
= p|+ ih+ | + (1 p) .
4
Then one can calculate min (TA ) = p2 +
1p
4
Maybe PPT = Sep? Unfortunately not. In D(C2 C3 ) (i.e. density matrices in which one
system has 2 dimensions and the other has 3) then all PPT states are separable. But for larger
systems, e.g. 3x3 or 2x4, then there exist PPT states that are not separable.
7.2.1
Bound entanglement
2. PPT is closed under tensor product. If AB , A0 B 0 PPT, then (AB A0 B 0 ) PPT. Why?
Because
T 0
A
AA0 B 0 0.
(AB A0 B 0 )TAA0 = TAB
Proof of Theorem 7.2. Assume towards a contradiction that PPT and ED () > 0. Then for
+
any > 0 there exists n such that n
AB can be transformed to | i using LOCC up to error . Since
n
PPT, is also PPT and so is the output of the protocol, which we call . Then TA 0 and
k |+ ih+ |k1 . If we had = 0, then this would be a contradiction, because is in PPT
and |+ ih+ | is not. We can use an argument based on continuity (details omitted) to show that
a contradiction must appear even for some sufficiently small > 0.
If is entangled but ED () = 0, then we say that has bound entanglement meaning that it
is entangled, but no pure entanglement can be extracted from it. By Theorem 7.2, we know that
any state in PPT but not Sep must be bound entangled.
Open question: A major open question (the NPT bound entanglement question) is whether
there exist bound entangled states that have a non-positive partial transpose.
7.3
Entanglement witnesses
Sep is a convex set, meaning that if , Sep and 0 1 then + (1 ) Sep. Thus the
separating hyperplane theorem implies that for any 6 Sep, there exists a Hermitian matrix W
such that
1. For all Sep, tr(W ) 0
2. tr(W ) < 0.
Example: consider the state = |+ ih+ |. Let W = I 2|+ ih+ |. As an exercise, show that
tr(W ) 0 for all Sep. We can also check that tr(W ) = 1.
Observe that an entanglement witness W needs to be chosen with a specific in mind. An an
exercise, show that no W can be a witness for all entangled states of a particular dimension.
25
7.4
CHSH game
One very famous type of entanglement witness is called a Bell inequality. In fact, these bounds rule
out not only separable states but even classically correlated distributions over states that could
be from a theory more general than quantum mechanics. Historically, Bell inequalities have been
important in showing that entanglement is an inescapable, and experimentally testable, part of
quantum mechanics.
The game is played by two players, Alice and Bob, together with a Referee. The Referee choose
bits r, s at random and sends r to Alice and s to Bob. Alice then sends a bit a back to the Referee
and Bob sends the bit b to the Referee.
R
r
s
a
Alice and Bob win if a b = r s, i.e. they want a b to be chosen according to this table:
r
0
0
1
1
s
0
1
0
1
desired a b
0
0
0
1
One can show that if Alice and Bob use a deterministic strategy, their success probability will
be 3/4. However, using entanglement they can achieve a success probability of cos2 (/8)
0.854 . . . > 3/4. This strategy, together with the payoff function (+1 if they win, -1 if they lose),
yields an entanglement witness, and one that can be implemented only with local measurements.
26
29 May, 2013
8.1
Lecture 8
Last lecture we looked at quantum states of three parties |iABC C2 C2 C2 . We say that if
A , B , and C are the reductions of |iABC then
B
C
A
max + max 1 + max .
(8.1)
h|A |i +
max
|i,k|ik2 =1
max
h|B |i
|i,k|ik2 =1
=
=
A ,B
A ,B
A ,B
= =
AB
max
= C
max ,
(8.2)
(8.3)
(8.4)
(8.5)
(8.6)
Likewise,
and
need picture... (add explanation later)
Next lecture we will see how the mathematics of representation theory is useful for generalizing
this result to higher dimensions.
27
8.2
8.2.1
Quantum Instrument
X
X
hi|B 0 U |0iB 0 A h0|B U |iiB =
Ei A Ei ,
i
(8.7)
8.2.2
Going back to the LOCC protocol, Alice first measurement can be modelled by a set of Kraus
operators {Ai1 }. Then Bobs measurement, which can depends on Alices outcome, will be given
by {Bi1 ,i2 }, and so on. In terms of a quantum operation a n-round LOCC protocol can be written
as
X
() =
(Ai1 ,...,in . . . Ai1 Bi1 ,...,in . . . Bi1 i2 )(Ai1 ,...,in . . . Ai1 Bi1 ,...,in . . . Bi1 i2 )
i1 ...,in
8.2.3
(8.9)
The general form of a LOCC operation (Eq. (8.9)) is daunting. It turns out that the whole picture
simplifies if we condition on measurement outcomes, which corresponds on only considering one of
the terms in the sum of Eq. (8.9). We call this operation Stochastic LOCC and it is an operation
which can only be implemented with some non-zero probability. A general form of SLOCC (for
three parties) is then
() = (A B C)(A B C) .
(8.10)
Note that p := tr(()) gives the probability that is implemented (with probability 1 p the
protocol fails and the state is transformed into something else).
28
(8.11)
The three following ones are the classes where only two parties are entangled, with representative
states
|+ iAB |0iC
(8.12)
and likewise for AC and BC.
The 5th class if the so-called GHZ class, with representative state
1
(|0, 0, 0i + |1, 1, 1i).
2
(8.13)
29
(8.14)
29 May, 2013
9.1
Lecture 9
de Finetti
(9.1)
(9.2)
(9.3)
(9.4)
(9.5)
D+n1
dp |hv ||ik |2 =
d|h
v ||ik |2
(9.7)
Z
D+n1
=
d|h|(idk |in )|ik |2
(9.8)
D1
Z
D+n1
=
tr(|ih| d|ih|n+k )
(9.9)
D1
D+n1
D+n+k1
D+n1
D+n+k1
=
/
tr(|ih|) =
/
D1
D1
D1
D1
(9.10)
30
We can now get a good lower bound from this by expanding the binomial coefficients into
factorials:
D+n1
D+n+k1
(n + 1) (n + D 1)
/
=
(1k/(n+1))D1 1kD/n.
(n + k + 1) (n + k + D 1)
D1
D1
Note that this bound is polynomial in n. This is tight. There exists, however, an improvement
to an exponential dependence in n at the cost of replacing product states by almost product states.
In order to conclude the proof of the quantum de Finetti theorem, we need to relate the trace
distance to the average we computed.
For this, we consider
the fidelity |h||i| between states |i and |i. If now |h||i| = 1 and
9.2
A surprising application of entanglement is quantum key distribution. Suppose Alice and Bob share
an EPR pair |i = 12 |00 + 11i, then the joint state of Alice Bob and a potential eavesdropper Eve
is |iABE s.th. trE |ih|ABE = |ih|AB it follows that |iABE = |iAB |iE
By measuring in their standard basis, Alice and Bob thus obtain a secret random bit r. They
can use this bit to send a bit securely with help of the Vernam one-time pad cipher: Lets call
Alices message m. Alice sends the cipher c = m r to Bob. Bob then recovers the message by
adding r: c r = m r r = m.
How can we establish shared entanglement between Alice and Bob? Alice could for instance
create the state locally and send it to Bob using a quantum channel (i.e. a glas fibre).
But how can we now verify that the joint state that Alice and Bob have after the transmission
is an EPR state?
Protocol 1) Alice sends halves of n EPR pairs to Bob
2) They choose randomly half of them and perform the CHSH tests. 3) They get key from the
remaining halves
There are many technical details that I am glossing over here. One is, how can you be confident
that the other halves are in this state? The de Finetti theorem! (the choice was permutation
invariant)
In order to make this applicable in actual implementations, one may use the exponential de
Finetti theorem (Renner) or the post-selection technique (Christandl, Konig, Renner).
Other issues: there may be noise on the line. It is indeed possible to quantum key distribution
even in this case, but here one needs some other tools mainly relating to classical information theory
(information reconciliation or privacy amplification).
31
30 May, 2013
10.1
Lecture 10
s
a
Alice and Bob win if a b = r s, i.e. they want a b to be chosen according to this table:
r
0
0
1
1
s
0
1
0
1
desired a b
0
0
0
1
Deterministic strategies. Consider a deterministic strategy. This means that if Alice receives
r = 0, she outputs the bit a0 and if she receives r = 1, she outputs the bit a1 . Similarly, Bob
outputs b0 if he receives s = 0 and b1 if he receives s = 1.
There are four possible inputs. If they set a0 = a1 = b0 = b1 = 0, then they will succeed with
probability 3/4. Can they do better? For a deterministic strategy this can only mean winning with
probability 1. But this implies that
a0 b0 = 0
a0 b1 = 0
a1 b0 = 0
a1 b1 = 1
Adding this up (and using x x = 0) we find 0 = 1, a contradiction.
Randomized strategies. What if Alice and Bob share some correlated random variable and
choose a deterministic strategy based on this? Then the payoff is the average of the payoffs of each
of the deterministic strategies. Thus, there must always be at least one deterministic strategy that
does at least as well as the average. So we can assume that an optimal strategy does not make use
of randomness. (Exercise: what if they use uncorrelated randomness? Can this help?)
Quantum strategies. Now suppose they share the state |+ i. Define
|0 ()i = cos()|0i + sin()|1i
|1 ()i = sin()|0i + cos()|1i
32
Observe that {|0 ()i, |1 ()i} is an orthonormal basis for any choice of .
The strategy is as follows. Alice and Bob will each measure their half of the entangled state in
the basis {|0 ()i, |1 ()i} for some choice of that depends on their inputs. They will output 0
or 1, depending on their measurement outcome. The choices of are
Alice
r=0
r=1
=0
= /4
Bob
s=0
s=1
= /8
= /8
1
As an exercise show that Pr[win] = cos2 (/8) = 21 + 2
> 3/4.
2
Another way to look at the quantum strategy is in terms of local observables. Alice and Bobs
strategy can be described in terms of the matrices
1 0
1 0
A0 =
A0 =
0 1
0 1
1 1 1
1
1 1
B0 =
B1 =
2 1 1
2 1 1
Given a state |i, the value of the game can be expressed in terms of
1
h|(A0 B0 + A0 B1 + A1 B0 A1 B1 )|i = Pr[win] Pr[lose] = 2 Pr[win] 1.
4
We can define a Hermitian matrix W 0 by
1
W = (A0 B0 + A0 B1 + A1 B0 A1 B1 ).
4
Then, for any Sep,
1
3
1
tr(W 0 ) 2 max Pr[win] 1 = 1 =
4
2
2
Define W = I2 14 W 0 . Then for all Sep, tr(W ) 0, while tr(W |+ ih+ |) = 12 < 0.
Thus Bell inequalities define entanglement witnesses; moreover, ones that distinguish an entangled state even from separable states over unbounded dimension that are measured with possibly
different measurement operators.
There has been some exciting recent work on the CHSH game. One recent line of work has
been on the rigidity property, which states that any quantum strategy that comes within of the
1
must be within 0 of the ideal strategy (up to some trivial changes). This
optimal value 21 + 2
2
is relevant to the field of device-independent quantum information processing, which attempts to
draw conclusions about an untrusted quantum device based only on local measurement outcomes.
(For more references see arXiv:1203.2976 and arXiv:1303.3081.)
33
10.2
Computational complexity
Figure 1: This figure is taken from the wikipedia article http://en.wikipedia.org/wiki/Clique (graph
theory). The 42 2-cliques are the edges, the 19 3-cliques are the triangles colored light blue and the
2 4-cliques are colored dark blue. There are no 5-cliques.
NP-hardness. It is generally very difficult to prove that a problem cannot be solved efficiently.
For example, it is strongly believed that 3-SAT is not in P, but there is no proof of this conjecture.
Instead, to establish hardness we need to settle for finding evidence that falls short of a proof.
Some of the strongest evidence we are able to obtain for this is to show that a problem is
NP-hard, which means that any problem in NP be efficiently reduced to it. For example, 3-SAT is
NP-hard. This means that if we could solve 3-SAT instances of length n in time T (n), then any
other problem in NP could be solved in time poly(T (poly(n))). In particular, if 3-SAT were in
P then it would follow that P = NP.
It is conjectured that P 6= NP, because it seems harder to find a solution in general than to
recognize a solution. This is one of the biggest open problems in mathematics, and all partial results
in this direction are much much weaker. However, if we assume for now that P 6= NP, then showing
a problem is NP-hard implies that it is not in P. And since thousands of problems are known to be
NP-hard1 it suffices to show a reduction from any NP-hard problem in order to show that a new
problem is also NP-hard. Thus, this can be an effective method of showing that a problem is likely
to be hard.
Theorem 10.1. Problems 1 and 2 are NP-hard for = 1/ poly(n, m).
We will give only a sketch of the proof.
1. Argue that MAX-CLIQUE is NP-hard. This is a classical result that we will not reproduce
here. Given a graph G = (V, E) with vertices V and edges E, a clique is a subset S V such
that (i, j) E for each i, j S, i 6= j. An example is given in Fig. 1. The MAX-CLIQUE
problem asks for the size of the largest clique in a given graph.
2.
1
See this list: http://en.wikipedia.org/wiki/List of NP-complete problems. The terminology NP-complete refers to
problems that are both NP-hard and in NP.
35
Theorem 10.2 (Motzkin-Straus). Let G = (V, E) be a graph, with maximum clique of size
W . Then
X
1
pi pj ,
(10.1)
1
= 2 max
W
(i,j)E
(i,j)E
Then
maxhx| hx|M |xi |xi = max
|xi
Defining pi = |xi
|2 ,
kxk2 =1
|xi |2 |xj |2 .
(i,j)E
4. We argue that
hSep (M ) = max ha, b|M |a, bi.
|ai,|bi
This is because Sep is a convex set, its extreme points are of the form |a, biha, b|, and the
maximum of any linear function over a convex set can be achieved by an extreme point.
5. Finally, we argue that maximizing over |a, bi is equivalent in difficulty to maximizing over
|a, ai.
What accuracy do we need here? If we want to distinguish a clique of size n (where there are n
1
vertices) from size n 1, then we need accuracy (1 n1
) (1 n1 ) 1/n2 . Thus, we have shown
that problem 2 is NP-hard for = 1/n2 .
36
30 May, 2013
Lecture 11
Last time we talked about SLOCC (stochastic LOCC), where we can post-select on particular
outcomes
Given a class of states that can be interconverted by SLOCC into |i other by SLOCC, C =
{|i : |i |i}, a result by Dur-Vidal-Cirac says that
C := {(A B C)|i/k . . . k : A, B, C SL(d)}
(11.1)
For three qubits there is a simple classification of all possible types of entanglement. Apart
from product states, and states with only bipartite entanglement, the two classes have the following
representative states:
1
|GHZi = (|000i + |111i)
(11.2)
2
and
1
|W i = (|001i + |010i + |100i)
(11.3)
2
The class of SLOCC operations forms a group:
G = {A B C : A, B, C SL(d)}.
(11.4)
(11.5)
11.1
What are the possible A , B , C compatible with a pure state |iABC CABC ? We say before
that this only depends on the spectra A , B and C , as one can always apply local unitaries and
change the basis.
For example, for the W class the set of compatible spectra is given by the equation A
max +
B
max + C
2.
max
Let us start with a simpler problem, namely given a state |iABC , does there exist a state in
G.|iABC with A = B = C = I/d? This is equivalent to
tr(A A) = tr(B B) = tr(C C) = 0,
(11.6)
(11.7)
(11.8)
Thus the norm of the state |ABC i should not change (to 1st order) when we apply an infinitesimal
SLOCC operation.
Let us look at
tA
ke etB etC |ABC ik =
hABC |etA etB etC |ABC i
t
t
= 2hABC |A I I + I B I + I I C|ABC i. (11.9)
So from Eq. (11.8), if |ABC i is the closest point to the origin in G.|ABC i, then A = B =
C = I/d.
What happens when there is no point in the class with A = B = C = I/d. That seems
strange, as it implies by the above that there is no closest point to the origin. But indeed this is
the case for the |W i class, for example. Consider
(0; 01/) (00 01/) (00 01/)|W i = |W i
(11.10)
and when goes to zero, one approaches the origin. However the limit is not in G.|W i.
In general, we have:
Theorem 11.1. (Kempf-Ness) The following are equivalent:
There exists a closest point to 0 in G.|ABC i.
There exists a quantum state in G.|ABC i with A = B = C = I/d.
G.|ABC i is closed.
The theorem says that there is no point in G.|W i which is maximally mixed.
How about if we look at the closure of G.|W i?
A fact is that for the closure of G.|ABC i, for every |ABC i, contains a unique closed orbit.
Corollary 11.2. There exists a quantum state in the closure of G.|i with maximally mixed reductions if, and only if, 0
/ G.|ABC i.
We saw before that 0 is in the closure of the |W i class. Therefore we cannot approximate (to
arbitrary accuracy) states in the |W i class by a state with maximally mixed reductions.
If we are given a class G.|i and we would like to show it does not contain the origin, we can
find a function which separates the sets. It turns out we can always choose a polynomial for that
P , such that P (0) = 0 and P (|i) 6= 0 for all |i G.|i. We can choose the polynomial P to be
G-invariant (i.e. P (|i) = P (g.|i) for all g) and homogeneous. The converse is also true, so we
find that:
Theorem 11.3. 0
/ G.|ABC i if, and only if, there exists a G-invariant homogeneous polynomial
such that P (0) = 0 and P (|ABC i) 6= 0.
This new characterization doesnt look particularly useful at first sight since we have to check
all homogeneous G-invariant polynomials. However it turns out that we only have to check a
finite number of polynomials since the set of G-invariant polynomials is finitely generated (ref). A
particular case of this result to three qubits state is that any G-invariant polynomial is a sum of
powers of the Cayleys hyperdeterminant.
Next lecture we will see how we can use representation theory to study the G-invariant polynomials.
38
30 May, 2013
Lecture 12
Today, I will tell you about bizarre things that can happen with entanglement of high dimensional quantum states. Recall from Fernandos lecture that
hSEP (M ) = max tr M
SEP
where {M, id M } are the yes/no outcomes of a POVM. He also showed that it is NP-hard to
compute this quantity exactly in general. So, here we want to consider approximations to this
quantity that we can compute easier.
For this we introduce approximations to the set of separable states based on the concept of
n-extendibility.
Definition 4. Let AB D(C dA CdB ). We say that AB is n-extendible if there exists a state
AB1 Bn D(CdA Symn (CdB )) s.th. AB = trB2 Bn AB1 Bn .
It turns out that the set of n-extendible states is a good outer approximation to the set of
separable states that gets better and better as n increases. But let us first check that the set of
separable states is contained in it, that is, that every
P separable AB is n-extendible. This can be seen
by writing the separable state AB in the P
form i pi |i ihi | |i ihi |. A symmetric extension is
then easily seen to be given by AB1 Bn = i pi |i ihi | |i ihi |n . The approximation statement
is then contained in the following theorem.
Theorem 12.1. If AB is n-extendible, then there is a separable state with 12 || ||1 nd .
The proof of this theorem is very similar to the proof of the quantum de Finetti theorem which
we did yesterday (in fact, you could adapt the proof as an exercise if you wish).
As a corollary it now follows that we can approximate hSEP (M ) by
hnext (M ) := max tr M .
next
d
.
n
The lower bound follows directly from the fact that the set of separable states is contained in
the set of n-extendible states (it even holds for all hermitian M without the restriction 0 M id.
For the upper bound, we use the theorem the observation that
1
max tr M ( ) = || ||1
0M id
2
and obtain
hnext (M ) = max tr M max tr M +
sep
next
39
d
.
n
We now want to see how difficult it is to compute hnext (M ). We rewrite hnext (M ) in the
form
hnext (M ) =
=
max
h|M idn1 |i
(12.1)
max
tr M trn1 |ih|
(12.2)
(12.3)
Hence, the effort to compute hnext (M ) is polynomial in dn+1 . In order to obtain an approximation to hsep (M ) we have to choose = d/n according to the corollary. Hence the effort to
approximate up to accuracy then the effort scales as dn/ .
Actually this is optimal for general M . In order to see why, we are going to employ a quantum
states known as the antisymmetric states (it is also known to the the universal counter example to
any conjecture in entanglement theory which you may have). The antisymmetric state comes in a
pair with the symmetric state:
The symmetric state is
d,2
id +F
sym
sym =
=
.
d(d + 1)/2
d(d + 1)
d,2
R
sym
It is indeed separable, because d(d+1)/2
= d|ih|2 .
The antisymmetric state is
anti =
id d,2
id F
sym
=
d(d 1)/2
d(d 1)
1
2 ||anti
||1
1
2
2. it is very extendible; more precisely, two copies anti anti are d 1-extendible.
Let us first see why 1. holds. For this note that for M = d,2
sym : tr M anti = 0. On the other
hand
1 1
tr M = tr(/2 + F /2) = + tr F .
2 2
In order to bound tr F note that
X
X
X
tr F (X Y ) =
hi, j|F (X Y )|i, ji =
hj, i|(X Y )|i, ji =
Xji Yij = tr XY.
i,j
i,j
ij
Hence
1 1X
+
pi tr F |i ihi | |i ihi |
2 2
i
1 1X
1
= +
pi |hi |i i|2 .
2 2
2
tr M =
40
(12.4)
(12.5)
anti =
.
2
2
1i<jd
X
1 X
sgn()|(1)i |(n)i.
d!
Sd
sgn() = (1)
n
X
= det(
|(i)ihi|).
i
Let us now verify that the Slater determinant extends anti , i.e. that
anti = tr3d |ih|.
This can be done by a quick direct calculation of the partial trace using the formula
X
tr3d |ih| =
id hi3 in ||ih| id |i3 in i.
i3 in
resulting in
X
1
|jk kjihjk kj|.
d(d 1)
j<k
Note that the extension we constructed was actually antisymmetric and not symmetric. But if
we take two copies of the antisymmetric state, the negative signs cancel out:
|i|i Symn (C d C d )
and we thus have the desired d 1 extension of anti anti .
41
31 May, 2013
LOCC distinguishability
Lecturer: Fernando G.S.L. Brand
ao
13.1
Lecture 13
Data Hiding
=
WAB
D(Cd Cd ).
d(d 1)
AB k1 = 12 .
It is k-extendable for k = d 1, and satisfies minSep 21 kWAB
Here the trace distance describes our ability to distinguish W and using arbitrary twooutcome measurements {M, I M } satisfying only 0 M I. However, since arbitrary measurements can be hard to implement, it is often reasonable to consider the smaller class of measurements
that can be implemented with LOCC.
Locality-restricted measurements
Define the LOCC norm to be
1
kAB AB kLOCC :=
2
max
0M I
{M,IM }LOCC
| tr(M ( ))|.
Define the 1-LOCC norm to be analogous, but with LOCC replaced with 1-LOCC. This stands
for one-way LOCC. This means that one party (by convention, Bob) makes a measurement,
sends the outcome to Alice and she makes a measurement based on this message. The resulting
measurements always have the form
X
M=
Ak B k
k
0 Ak I
for each k
0 Bk
X
Bk = I
for each k
42
Sep
1
1
+
kWAB
AB kLOCC kWAB
WAB
kLOCC ,
2
2
R
I+F
= d|i |ih|2 is separable.
since W + := d(d+1)
P
P
Observe that if {M, I M } LOCC then we have M = k Ak Bk and I M = k A0k Bk0
with each Ak , Bk , A0k , Bk0 0. Thus
0 M TA I.
(13.1)
We can then further relax
1
+
kWAB
WAB
kLOCC
2
=
max
tr(M (W + W ))
max
0M I
0M TA I
0M I
0M TA I
1
k(W + )TA (W )TA k1 .
2
P
To evaluate this last quantity, observe that if F = di,j=1 |i, jihi, j|, then F TA = d+ where
P
+ := |+ ih+ | and |+ i = 1d di=1 |i, ii. Then
TA
(W )
+ TA
(W )
=
=
I F
d(d 1)
TA
I +F
d(d + 1)
TA
I F TA
I d+
=
d(d 1)
d(d 1)
I + F TA
I + d+
=
d(d + 1)
d(d + 1)
13.2
This data hiding example raises the hope that a more useful version of the de Finetti theorem might
hold when we look at 1-LOCC measurements. Indeed, we will see that the following improved de
Finetti theorem does hold:
Theorem 13.1. If AB D(CdA CdB ) is k-extendable then
r
2 ln(2) log(dA )
.
min kAB AB k1-LOCC
Sep
k
43
(13.2)
1
1 p
k k2 :=
tr( )2 .
127
127
We will solve this problem using semidefinite programming (SDP), which means optimizing
a linear function over matrices subject to semidefinite constraints (i.e. constraints that a given
matrix is positive semidefinite). Algorithms are known that can solve SDPs in time polynomial
in the number of variables. The SDP for checking whether AB is k-extendable is to search for a
AB1 ,...,Bk satisfying
AB1 ,...,Bk 0
j
ABj = AB .
A)
The algorithm is to run the SDP for k = 4 ln(2)log(d
. If Sep then the SDP will be
2
feasible because is also k-extendable. The harder case is to show that the SDP is infeasible when
D(AB , Sep) . But this follows from Theorem 13.1.
k
The run time is polynomial in dA dk+1
B , which is dominated by the dB term. This is
f () = c .
If we had such a measure, the proof would be very easy.
log(dA ) E(A:B1 ...Bk )
by normalization
k
X
E(A:Bi )
by monogamy
repeating the argument
i=1
= kE()
44
Rearranging, we have E() log(dA )/k, and finally we use faithfulness to argue that D(, Sep)
A)
).
f ( log(d
k
Such a measure does exist! It is called squashed entanglement and was introduced in 2003
by our very own Matthias Christandl and Andreas Winter [quant-ph/0308088]. Normalization
and monogamy are straightforward to prove for it (and were proved in the original paper), but
faithfulness was not proved until 2010 [Brandao, Christandl, Yard; arXiv:1010.1750].
45
31 May, 2013
14.1
Lecture 14
Representation Theory
Given a group G a representation of G in U (V ), the set of unitaries over the vector space V = C
is a mapping g 7 U (g) U (V ) which is a homomorphism (...).
Definition 5. We say a representation is irreducible if for all such that
W V, U (g)|wig, q W
(14.1)
we have W = {0, V }.
(g), we say they are equivalent if
Definition 6. Given two representations g 7 U (g) and g 7 U
1
(g) for all g.
there exists a invertible A such that AU (g)A = U
We have the following important theorem
Theorem 14.1. For G a finite or Lie group, U (g) = U1 (g) . . . Ul (g) with U1 , . . . , Ul irreducible
representations. Analogously,
M
V =
Vi Cmi ,
(14.2)
=
iG
(14.3)
(14.4)
and
Then the action of the group in Vj is obtained by exponentiating the Lie algebra.
We can write Vj = Sym2j (C2 ) (C2 )2j and the action of the group in Vj as g 7 sym g
. . . gsym .
46
14.1.1
Schur-Weyl Duality
An important theorem in representation theory of SU (d) and the symmetric group is the so-called
Schur-Weyl duality.
Consider the following representation of SU (2): g 7 g . . . g. The associated vector space
is V n = (C2 )n . We can decompose this representation into irreps as
M
V (n) =
Vj Cmj .
(14.5)
j
14.1.2
Computing mj
M
n
=
Vj Cmj V1/2
j
M
n
=
Vj Cmj V1/2
j
Vj 0 C
(mn
+mn
)
j 0 +1/2
j 0 1/2
j0
= ....
(14.6)
2nh(1/2j/n)
(14.7)
n/2 j
n/2 j 1
with h the binary entropy.
But actually, Vj Cmj (C2 )n = span{(|01i|10i)(n/2j) |j, mi}, with Sn . The vectors
|j, mi C2j are equal to to
|j, mi = c(|0, . . . , 0, 1, . . . , 1i + permutations of 0s and 1s),
with j + m 0s and j m 1s.
47
(14.8)
14.2
Spectrum Estimation
Let us for a moment forget about representation theory and consider the problem of spectrum
estimation. In this problem we are given a source of quantum states which gives n copies of an
unknown density matrix: n . The goal is to perform a measurement which gives a estimate of .
Suppose we are only interested in estimating the eigenvalues of .
For a qubit state , the eigenvalues of we can be written as (1/2 + r, 1/2 r), with r
[0, 1/2]. The problem of estimating r was considered by Keyl and Werner [], who provided an
interesting connection of the problem to representation theory of SU (2). The measurement consists
of projecting the n copies into the spaces Vj Cmj introduced in the last section.
Informally, the claim is that with high probability j/m r. More precisely,
Pr(j) = tr(Pj n ) constenD(1/2+j/n||1/2+r) ,
(14.9)
where D(x||y) is the relative entropy of x and y, given by D(x||y) = x log xx log y +(1x) log(1
x) (1 x) log(1 y).
Let us now sketch the proof. We can compute
(h0, 1| h1, 0|) (|0, 1i |1, 0i) = (1/2 + r)(1/2 r).
(14.10)
tr(Pj
X
j
n
)
(1/2 r)n/2j (1 r)n/2+j (1/2 + r)j+m (1/2 r)jm
n(1/2 j/n)
m=j
X
(1/2 r)k
n
k=0
Noting that
14.2.1
(1/2r)k
k=0 (1/2+r)k
48
31 May, 2013
15.1
Lecture 15
Introduction
in this lecture, I will give a proof of the following theorem, first mentioned by Fernando, on the
way introducing useful properties of von Neumann and Shannon entropy.
Theorem 15.1. Let AB be k-extendible and M 0 a 1-LOCC measurement. Then there exists a
separable state such that
r
log dA
0
| tr M ( )| const
.
k
It is possible to swap the quantifiers with help of von Neumanns minimax theorem.
Pm
0
Recall that M 0 can be written in the
Pform M = i=1 Ai Bk for 0 Ai id and 0 Bi id.
Define the measurement M : M () = i tr Bi |iihi|. Note that AB extendible implies that there
exists a state AB1 Bn s.th. AB = ABi for all i. It is the goal to show that
id M (AB ) id M (AB )
for separable :
AB =
|i ihi | |i ihi |
Writing AB1 =
x
x px AB1 ,
x
px AB
1
AB1 x M (x )
x
x
AB1 A
B
15.2
Entropy
Recall that the Shannon entropy of a probability distribution p of a random variable X is given by
X
px log px = H(X)p .
H(p) =
x
P
When we have
joint
distributions
p(xy),
we
can
look
at
the
marginal
distributions
p(x)
=
y p(xy)
P
and p(y) = x p(xy) and their entropies. The conditional entropy of X given Y is defined as
X
p(y)H(X)p(x|y)
H(X|Y )p =
y
where p(x|y) = p(xy)/p(y). Writing the conditional entropy out explicitly we find the formula
H(X|Y )p = H(XY )p H(Y )p .
This gives us a beautiful interpretation of the conditional entropy: it is just the entropy of the joint
distribution minus the entropy of Y .
We can measure the correlation between two random variables X and Y by looking at the
difference between the entropy H(X) and H(X|Y )
I(X; Y ) = H(X) H(X|Y ) = H(X) + H(Y ) H(XY )
Note that this quantity is symmetric with respect to interchange of X and Y . This is known as
the mutual information and quantifies the by how much our uncertainty about X changes when we
are given Y (and vice versa, of course). It is also the amount of bits that you save by compressing
XY together as opposed to compressing X and Y separately.
The mutual information has a few nice properties:
I(X : Y ) 0
I(X : Y ) log |X| and I(X : Y ) log |Y |, where |X| denotes the number of symbols in X.
1
2 ln 2
2
|p(xy)
p(x)p(y)|
x,y
P
Let us now look at the quantum version of all this. I will use the notation that S(A) = S(A )
and we have a joint state AB , then S(A) = S(trB AB ). Note that it is not immediately clear
how to define the conditional entropy in the quantum case, since we cannot condition on a value of
system B. Luckily we had a second way of writing the conditional entropy and we are just going
to define the conditional quantum entropy as
S(A|B) := S(AB) S(B)
and the mutual information as
I(A : B) = S(A) + S(B) S(AB).
It has the following properties
50
S(A : B) 0
I(A : B) 2 log dA and I(A : B) 2 log dB (note the factor of two!)
(Pinskers inequality) I(A : B)
1
2 ln 2 ||AB
A B ||21
This last property looks like it could be useful in proving the theorem; but note that it all
cannot be that easy, because we know that we should use the 1-LOCC norm and not the trace
norm...In order to proceed, we need the conditional mutual information:
I(A : B|C) = S(A|C) + S(B|C) S(AB|C) = S(AC) + S(BC) S(ABC) S(C)
This formula is a little difficult to grasp and it is difficult to get an intuition for this quantity. The
conditional mutual information has a nice property, though, it satisfies the chain rule:
I(A : BC) = I(A : C) I(A : B|C)
which you can easily check. It is called the chain rule, in part, because we can iterate it: Lets
assume we have a k-extendible state and we measure all of the B systems. How much does A know
about the Bs? The chain rule gives us
I(A : B1 Bk ) = I(A : B1 ) + I(A : B2 Bk |B1 )
(15.1)
(15.2)
(15.3)
There are k terms and the sum is small than log d, hence there is a j such that I(A : Bk |B1 B2 Bk1
log d
k . This then immediately gives a proof of our theorem (see arXiv:1210.6367 for more details).
51